Sei sulla pagina 1di 150

Understanding the Zed-Meter Instrument (Draft)

Lightning Impulse Impedance of Transmission Tower Footings and Ground Electrodes


1015904

Understanding the Zed-Meter Instrument (Draft)


Lightning Impulse Impedance of Transmission Tower Footings and Ground Electrodes 1015904 Technical Update, December 2008

EPRI Project Manager F. Bologna

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 Kinectrics, Inc.

This is an EPRI Technical Update report. A Technical Update report is intended as an informal report of continuing research, a meeting, or a topical study. It is not a final EPRI technical report.

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, EPRI, and TOGETHERSHAPING THE FUTURE OF ELECTRICITY are registered service marks of the Electric Power Research Institute, Inc. Copyright 2008 Electric Power Research Institute, Inc. All rights reserved.

CITATIONS
This document was prepared by Kinectrics 800 Kipling Avenue Toronto, Ontario, Canada M8Z 6C4 Principal Investigators W. A. Chisholm E. Petrache This document describes research sponsored by the Electric Power Research Institute (EPRI). This publication is a corporate document that should be cited in the literature in the following manner: Understanding the Zed-Meter Instrument (Draft): Lightning Impulse Impedance of Transmission Tower Footings and Ground Electrodes. EPRI, Palo Alto, CA: 2008. 1015904.

iii

PRODUCT DESCRIPTION
Most utilities test transmission line tower ground resistance by isolating overhead groundwires; inserting low-frequency, battery-operated test equipment; taking readings; and restoring the connections. The resulting values are useful for power frequency grounding but less representative for lightning performance. Utilities can improve test productivity by switching to a Zed-Meter instrument test technique that uses off-the-shelf equipment and injects a safe, lightning-like impulse signal into the tower base. Results and Findings The Zed-Meter instruments method for testing transmission line grounds relies on the injection of a safe, transient pulse into the tower base. The pulse is similar to lightning, and it delivers the results that utilities need for improving the lightning performance of lines. An additional benefit of this approach is that the surge impedance of the lead wires, rather than the galvanic resistance of driven rods, provides the connection to ground. This means that the Zed-Meter test leads need not be grounded; they need only be laid on the surface of the right of way. In addition, the overhead groundwires need not be isolated from the tower. These are important advantages that save test time, especially in frozen soil or rocky areas. Challenges and Objectives The objective of this report is to help users of the Zed-Meter instrument to apply it more effectively and to understand technical issues that might be encountered in the field. Application, Value, and Use The Zed-Meter instrument was designed to measure the potential rise at the base of a fourfooting high-voltage or extrahigh-voltage lattice tower without having to encircle the entire tower in a large, expensive sensor for the tower current. This report shows that the reaction electrode that is used to push current into the tower also provides a value of surge impedance that can be used to establish the soil resistivity near the tower. Other waveform features from this test can yield more information about the tower surge response, transfer impedance to nearby equipment, and the soil resistivity in the top layer of the soil. EPRI Perspective EPRI members have collectively used every existing method to measure transmission line grounding. Each has its strengths but none is completely appropriate for transmission line groundingsome require expensive equipment, and others take too long to set up at each tower. What separates the Zed-Meter instrument from the existing technologies is the right combination of test speed and accuracy for its intended purposelightning protection.

Approach The goal of this report is to explain how to use the Zed-Meter instrument. It consists of the following sections: Section 1, Overview of the Zed-Meter Instrument Section 2, Zed-Meter Instrument Bench Tests Section 3, Zed-Meter Instrument Dipole Tests of Reaction Leads Section 4, Zed-Meter Instrument Tests on Transmission Towers Section 5, Running the Zed-Meter System Software Section 6, Typical Zed-Meter Instrument Results Section 7, Post-Processing Section 8, References Appendix A, Frequently Asked Questions Appendix B, Specifications Appendix C Troubleshooting Guide Appendix D, Hardware Evolution Appendix E, The Current Reaction Lead Appendix F, Modeling Zed-Meter Instrument Leads with NEC-4 Software

Keywords Lightning Surge impedance Test methods Transmission lines Zed-Meter instrument

vi

ABSTRACT
The Zed-Meter instruments method for testing transmission line grounds relies on the injection of a safe, transient pulse into the tower base. The pulse is similar to lightning, and it delivers the results that utilities need for improving the lightning performance of lines. An additional benefit of this approach is that the surge impedance of the lead wires, rather than the galvanic resistance of driven rods, provides the connection to ground. This means that the ZedMeter test leads need not be grounded; they need only be laid on the surface of the right of way. In addition, the overhead groundwires need not be isolated from the tower. These are important advantages that save test time, especially in frozen soil or rocky areas.

vii

CONTENTS
1 OVERVIEW OF THE ZED-METER INSTRUMENT ..............................................................1-1 What is the Zed-Meter Instrument? ...................................................................................1-1 How Does the Zed-Meter Instrument Differ from Conventional Instruments?...................1-3 How and Why Do the Zed-Meter Instruments Results Differ from Typical Resistance Measurements? ...................................................................................................................1-5 What Is the Basic Principle of Operation?............................................................................1-6 Labeled Diagram of Instrument............................................................................................1-7 Test Lead Arrangement .......................................................................................................1-9 Sequence of Zed-Meter Operations ..................................................................................1-9 Operating Temperature Ranges ........................................................................................1-10 2 ZED-METER INSTRUMENT BENCH TESTS .....................................................................2-1 Operational Test: Charging the Battery................................................................................2-1 Operational Test: Short Circuit.............................................................................................2-1 Connections and Process ..............................................................................................2-1 Results for High-Quality Short Circuit.............................................................................2-2 Results for Poor-Quality Short Circuit ............................................................................2-4 Operational Test: Fixed Resistance .....................................................................................2-5 Reference Result: 54- Calibration Resistor .................................................................2-5 Reference Result: 500- Resistor .................................................................................2-7 Practice Makes Perfect ........................................................................................................2-9 3 ZED-METER INSTRUMENT DIPOLE TESTS OF REACTION LEADS ..............................3-1 Function of Reaction and Potential Leads ...........................................................................3-1 Dipole Impedance Test Method ...........................................................................................3-1 Variation of Dipole Impedance with Lead Orientation and Length .......................................3-2 Dipole Test Results, 90-m Versus 130-m Lead Length .......................................................3-3 Grounding of Reaction and Potential Leads ........................................................................3-4 Reasons for Grounding the Current Reaction Lead .......................................................3-4 Reasons for Grounding the Remote Potential Lead ......................................................3-4 Troubleshooting the Zed-Meter Instruments Dipole Test Results ....................................3-5 Verifying a Safe Work Environment ...............................................................................3-5 Symptoms of Excessive Noise Level .............................................................................3-5 Symptoms of Problems in the Lead Layout ...................................................................3-5 Symptoms of Conductors Running in Parallel................................................................3-5 Symptoms of Problems in the Lead Terminations .........................................................3-6 4 ZED-METER INSTRUMENT TESTS ON TRANSMISSION TOWERS ...............................4-1 Connection Diagram for Impedance Test ............................................................................4-1 Preferred Type of Tower Connection ...................................................................................4-2

ix

Preferred Location of Tower Connection .............................................................................4-4 Orientation of Reaction and Potential Leads........................................................................4-7 Dealing with Obstructions When Laying Out Leads...........................................................4-10 Twists and Turns: The Meander Line...........................................................................4-11 Using Shorter Potential Leads .....................................................................................4-13 Vehicles in Proximity to Leads .....................................................................................4-15 Considerations for Guyed Towers......................................................................................4-16 Considerations for Twin Steel-Pole Towers .......................................................................4-18 Overhead Groundwire Connections...................................................................................4-20 Comparison Tests with and Without Overhead Groundwires ......................................4-20 Calculation of Overhead Groundwire Impedance ........................................................4-22 Unparalleling the Effect of Overhead Groundwires ......................................................4-22 Which Is the Correct Value to Use? .............................................................................4-23 Tests on Towers with High Noise Level .............................................................................4-23 Tests on Towers with Buried Counterpoise .......................................................................4-24 Test Lead Orientation for Lines with Counterpoise ............................................................4-24 Which Leg Has the Counterpoise Connection? ...........................................................4-25 5 RUNNING THE ZED-METER SYSTEM SOFTWARE .........................................................5-1 Overview ..............................................................................................................................5-1 Main User Interface..............................................................................................................5-3 Section 2 Elements ........................................................................................................5-4 Section 3 Elements ........................................................................................................5-7 Setting the Time Windows ...........................................................................................5-12 Setting the Alarm Thresholds .......................................................................................5-12 Pretrigger Interval.........................................................................................................5-12 Demo Mode........................................................................................................................5-13 How-To: Start the Measurement Process ..........................................................................5-14 How-To: Stop the Measurement Process ..........................................................................5-15 How-To: Save the Measurement Data...............................................................................5-15 Data File.......................................................................................................................5-15 Log File ........................................................................................................................5-16 How-To: Create Line Reports ............................................................................................5-16 6 TYPICAL ZED-METER INSTRUMENT RESULTS..............................................................6-1 Effects of Overhead Groundwire..........................................................................................6-1 Reasons for Steadily Increasing Impedance Values............................................................6-1 Low-Frequency Versus High-Frequency Resistivity ......................................................6-2 Effect of the Time Window, 500 ns Versus 1000 ns.......................................................6-3 Special Case: Towers with Isolated Overhead Groundwires and High Resistivity.........6-4 Reasons for Steadily Decreasing Impedance Values ..........................................................6-6 Typical Responses of Buried Horizontal Wires ..............................................................6-6

Reading Past the First Sweet Spot ................................................................................6-8 Reasons for High Impedance Values.................................................................................6-10 What Constitutes a High Reading? ..............................................................................6-10 Identifying Bad Connection to Current Leads ..............................................................6-11 Indications of Local Soil Resistivity from Dipole Test Results ......................................6-11 Reasons for Low or Negative Impedance Values ..............................................................6-11 Identifying a Bad Connection to the Potential Lead .....................................................6-12 Inserting Resistance in Series with Tower to Validate Connections ............................6-12 Indications of Local Soil Resistivity from Dipole Test Results ......................................6-12 7 POST-PROCESSING .............................................................................................................7-1 Organization of Data ............................................................................................................7-1 Converting Waveform Files to Excel Format........................................................................7-1 Producing Waveform Graphs in Excel .................................................................................7-1 Exporting Waveform Graphs from Excel to Portable Document Format..............................7-2 Establishing Confidence Level in Test Results ....................................................................7-2 Adaptive Filtering to Improve Confidence ............................................................................7-2 Offset Subtraction ..........................................................................................................7-2 Denoising Digital Artifact ................................................................................................7-3 Removing Noise Using Pretrigger Record .....................................................................7-4 Validating a Novel Test Lead Arrangement .........................................................................7-5 Validating Against What? ...............................................................................................7-5 NEC-2 for Simple Coupling Estimates ...........................................................................7-8 8 REFERENCES .......................................................................................................................8-1 A FREQUENTLY ASKED QUESTIONS .................................................................................. A-1 B SPECIFICATIONS ................................................................................................................ B-1 C TROUBLESHOOTING GUIDE ............................................................................................. C-1 D HARDWARE EVOLUTION ................................................................................................... D-1 E THE CURRENT REACTION LEAD ...................................................................................... E-1 Dipole Impedance Test Theory ........................................................................................... E-1 Theoretical Variation of Dipole Impedance with Height ...................................................... E-3 Theoretical Variation of Dipole Impedance with Soil Resistivity.......................................... E-4 Theoretical Variation of Dipole Impedance with Lead Orientation ...................................... E-5 Conductors Running in Parallel with Test Leads ................................................................ E-7 2003 Field Studies at American Electric Power .................................................................. E-9 2005 Current Reaction Lead Studies at the CN Tower in Toronto.................................... E-13 Dipole Test Results, 90-m Versus 130-m Lead Length .................................................... E-16 Dipole Test Results, 125-m Coaxial Cable Test Leads..................................................... E-16 Dipole Test Results, 300-m, 14-Gauge Solid Copper Wires............................................. E-19

xi

F MODELING ZED-METER INSTRUMENT LEADS WITH NEC-4 SOFTWARE...................F-1 NEC-4 Software for Detailed Calculations Using Inverse Fast Fourier Transform ..............F-1 Other Modeling Software .....................................................................................................F-4 Finding an Experienced Practitioner ....................................................................................F-4

xii

OVERVIEW OF THE ZED-METER INSTRUMENT


What is the Zed-Meter Instrument? The Zed-Meter instrument is a test instrument that measures the impedance of transmission line grounds by generating an internal, lightning-like signal that is applied to the base of a transmission tower. The lightning impulse response of the local tower grounding system is monitored by tracing the rise and fall of tower base voltage as a function of the injected current for a short time after the signal is triggered. Rather than conducting this test with a single, largeamplitude pulse that has safety and weight issues, the Zed-Meter instrument averages results from repeated applications of lower-amplitude pulses from an electric fence shock generator that has been certified safe (although uncomfortable) for human and animal contact. The Zed-Meter instrument is a smart instrument that has benefited from improvements in computer technology to reduce its size, weight, and cost. Field trials have been conducted at Public Service Electric and Gas Company, Duke Energy, Tennessee Valley Authority, Georgia Power, Hydro-One, Bonneville Power Administration, Eskom, Manitoba Hydro, and National Grid. Each utility in its turn has seen improved versions of the Zed-Meter instrument that have been smaller, lighter, less expensive, and better at analyzing, presenting, and storing results. The utility input in the development cycle distinguishes the Zed-Meter instrument from the incremental improvements that have been incorporated in standard earth resistance testers during the same period. In addition to the instrument itself, the Zed-Meter kit contains the following components: A short cable and specially adapted clamp that feed the high-frequency test current into the tower A pair of leads, usually coaxial cables of approximately 100-m length, that are used as highfrequency traveling wave antennas An optional pair of ground rods and connectors for grounding the traveling wave antennas Table 1-1 shows how the Zed-Meter kit has progressed from a list of suitable parts to a selfcontained instrument with full control over pulse application and measurement with a Panasonic Toughbook laptop computer that represents more than half the overall cost.

1-1

Table 1-1 Zed-Meter instrument evolution: prototype to production 2004 2006 2008

Digitizer Pulse generator Current sensors Cable reels Power Control and analysis

Tektronix 3054B oscilloscope Lab grade 200 V, 50 with 10-ns rise time Wideband current transducers, sensitivity 1 V/A, 10-ns rise time 90 m RG58 (2) 12 Ah battery, 117 V inverter Manual setup of oscilloscope; manual control of pulser; comma-separated values (.csv) files on floppy disc to Microsoft Excel software

Tektronix 2000 series oscilloscope, US$6000 Farm grade, safe (Underwriters Laboratories [UL] approved) with wave shaping circuit

Acute DS-1102 200 MS/s USB digitizers, US$900 each

90 m RG58 (2) 12 Ah battery, 117 V inverter Macro setup of oscilloscope; manual control of pulser; comma-separated values (.csv) files through network to Microsoft Excel software

90130 m (2) Internal rechargeable battery National Instruments Labview software control of digitizers, pulser, file transfer through USB, and wave analysis

1-2

How Does the Zed-Meter Instrument Differ from Conventional Instruments? The Zed-Meter instrument works strictly in the time domain, specifically about 100 ns to 2 s, whereas most conventional earth resistance test instruments use one or more fixed frequencies around 100 Hz, making them low frequency-domain tools. Conventional, general-purpose, three- or four-terminal earth resistance testers also generate internal signals and perform the same basic functions as the Zed-Meter instrument. However, these testers use a low-frequency signal that gives the resistance of many towers in parallel. This result can indicate the contribution of individual towers to the low overall resistance only when expensive supplemental sensors are placed around each tower leg and guy wire. The lowfrequency, three-terminal methods also require the insertion of metal probes into the ground, achieving sufficient depth of penetration to allow injection of test signals. Typically, a probe resistance of less than 10 k must be achieved. This is difficult in frozen soil, because the resistivity is about 100 times higher than unfrozen soil. In winter testing, probes must sometimes be pushed all the way through the frozen soil layer, which can be more than 1 m thick. Conventional clamp-on earth resistance testers generate internal signals at a medium frequency of approximately 2 kHz. The energy is magnetically coupled into a single ground lead through steel jaws that open and close. The power needed to drive test current into the grounding system can then be measured. If the multigrounded neutral system is in good condition with low impedance, most of the measured impedance comes from a local (single) ground rod. This method has limitations if there are two or more paths to ground on the same tower or if the connection from local tower to overhead groundwire (OHGW) or neutral is imperfect. In order to simulate what the Zed-Meter does, a standard earth resistance tester would have to be modified to produce much higher frequencies. This in itself is not unique. One earth resistance tester operated at a single frequency, 26 kHz, and a CIGRE working group recommended that 150 kHz would be a better choice [1]. However, a modified earth resistance tester simulating the Zed-Meter would have to read out a series of measured impedance values, one for every frequency in a series covering the range from 10 kHz to 4 MHz. It is possible to convert results between the time and frequency domains using Fourier transformation, as illustrated in Figure 1-1. A perfect impulse signal contains all frequencies. A lightning impulse has high-frequency and low-frequency roll-off. For the most important waveshape of the first negative downward return stroke, the associated time constants are approximately 1.2 s on the front and 50 s on the time to half value. This means that a suitable frequency-domain instrument for grounding tests would measure with several different Nyquist frequencies in the range of 0.5/120 ns = 4 MHz to 0.5/50 s = 10 kHz. Most of the interesting effects occur at the sine-wave frequency of 124 kHz that has the same peak current (I=31 kA) and peak rate of current rise (dI/dt=24 kA/s) as a median lightning flash. This characteristic frequency drops to about 80 kHz for very large currents because there is a strong correlation between peak current and rate of current rise.

1-3

x 10 200

-4

FFT So urce

Injected Impulse [V]

150 |Imp ulse|

100

50

0 0 0.5 1 1.5 Time [microsecon ds] 2 2.5 0 -1 10 10 10 Fr equ ency [MHz]


0 1

10

Zed-Meter instruments test signal in time domain

Zed-Meter instruments test signal in frequency domain

Figure 1-1 Time and frequency domain equivalents for Zed-Meter test signal

Before committing to the development of the Zed-Meter instrument, EPRI reviewed a wide range of other instruments, including the EPRI Smart Ground Multimeter, which is used mainly for substation grounding tests. The results of this evaluation are given in Table 1-2..
Table 1-2 Suitablity matrix for transmission tower ground impedance (Z) testers Measurement Approach Clamp-on impedance meter Cost US$10002000 Setup Time 1 min Signal 2 kHz Strengths Ease of use, size, battery life, wide range of Z Size, battery life, wide time scale, useful for soil Size, noise rejection, wide range of Z, does not require driving rods into frozen soil Size, battery life, noise rejection, wide range of Z, useful for soil Size, ease of use Limitations Must fit downlead; needs bond to parallel towers through overhead shield wires Wrong readouts; narrow range of Z centers at 50 125-m leads

Time-domain reflectometer testers EPRI Zed-Meter instrument

US$100010,000

15 min

1-ns step or pulse 1- to 3-s pulse, 200 V, 50 100 140 Hz

US$300010,000

15 min

Four-terminal earth resistance testers ABB 26-kHz meter EPRI Smart Ground Multimeter

US$400010,000

3060 min

Needs tower isolation from shield wires; 100m leads; many intermediate steps Inaccurate for Z>25; 70-m leads; no resistivity Seven leads up to 60 m, all terminated with rods to give low resistance

US$40,000 (estimated) US$100,000 (estimated)

15 min

26 kHz

60 min

500-Hz random square wave

Noise rejection, wide range of Z, useful for layered soil

1-4

Table 1-2 lists a mixture of frequency-domain and time-domain instruments. With its internal analysis software, the EPRI Smart Ground Multimeter provides impedance readouts in the frequency domain using a Fourier analysis of a time-domain series of step pulses. This is promising, but the analysis extends only from low frequency to 500 Hz; therefore, the Smart Ground Multimeter is not suitable for measuring the lightning surge impedance of transmission tower ground electrodes.
How and Why Do the Zed-Meter Instruments Results Differ from Typical Resistance Measurements?

There is an important physical difference between the lightning impedance of a transmission tower grounding system and the impedance of the same system at power frequency. The lightning surge is so rapid that the peak stress on insulators occurs before adjacent towers have had a chance to react and share the surge current. The two-way propagation time at the speed of light to the nearest pair of towers, 300 m away, is about 2 s. A test surge can also be injected and measured at a local tower before the signals can detect what is happening far away from the tower under test. This is the basic advantage of the Zed-Meter tester. In the frequency domain, the effect is described differently. The series inductance of the overhead groundwires has a higher inductive reactance as frequency increases. For a 300-m span with 300 H series inductance, the inductive reactance will be 18.8 at 10 kHz, 188 at 100 kHz, and so on. At high frequency, the surge impedance, ZC (300 in this example) appears in parallel with the local footing (see Figure 1-2 in the following subsection). The Zed-Meter test result indicates whether the tower under test is well grounded, but it reads a value that is higher than the power-frequency impedance of the OHGW system with multiple grounds or the continuous counterpoise system. In practice, the Zed-Meter test result tends to be a bit lower than the impedance of the isolated tower measured at low frequency. The lightning performance (number of flashovers) of transmission lines is related to the values of the tower footing resistances along the line length. Low-frequency methods for measuring the resistance of transmission tower ground electrodes lose accuracy when OHGWs are connected to the towers, and they do not provide a result that is related to lightning performance. Accurate measurement of each individual tower resistance, whether tested at low or high frequency, can be obtained only if one or more of the following conditions are true:
The towers are temporarily isolated from parallel connections to remote earth Sensitive current sensors can be placed around ground leads The measurement frequency is raised to approximately 150 kHz [1] If the safety of a temporary working ground system relies on having low power-frequency impedance, the Zed-Meter tester is not the right choice for proof tests. When a power system fault occurs, many towers can share the fault current if they are connected together by an OHGW. The 60-Hz impedance of a multigrounded OHGW system can typically be approximately 2 if each individual footing resistance is approximately 20 . The same effect occurs for continuous counterpoise electrodes. The Zed-Meter tester does not resolve these remote effects, and it is not reasonable to assume that any fixed reduction factor will relate the local Zed-Meter test result to the power frequency impedance for many similar towers in

1-5

parallel. This is a case in which a properly applied low-frequency earth resistance test instrument should be used.
What Is the Basic Principle of Operation?

The basic operating principle of the Zed-Meter instrument is the following: Two leads, typically 90125 m long, are laid in straight lines in different directions away from a tower leg. One lead is used for current injection, and the other is used to measure the remote potential rise. The angle between the leads should be at least 90. A lightning-like transient current is injected into the tower base through a current sensor. The potential rise at the tower base relative to the remote potential is measured. A time-varying impedance profile, Z(t), is derived from the potential rise response to the injected current. Analysis identifies the desired features of Z(t), which, simply stated, are the median and the standard deviation of the values taken between two programmed time intervals. The impedance measurement vector, Z(t), is valid after the effects of the tower surge response have rung down and before the effects of the adjacent towers, or the far end of the current injection lead, have time to affect the reading. As a quality control measure, the current into the reaction lead is also measured and compared with the current into the tower. These are initially different when the measurement is not useful, but they should stabilize after some time (about 200 ns) to the same, constant value if the measurement leads are correct. In 2005, a CIGRE working group issued a brochure that reviewed methods for measuring the earth resistance of transmission towers equipped with earth wires [1]. This study considered all the options reviewed for EPRI in the EPRI reports The EPRI Zed-Meter: A New Technique to Evaluate Transmission Line Grounds (1008734), Field Testing of EPRI Zed-Meter: Transient Impedance of Transmission Line Grounds (1010235), and Summary of Zed-Meter Field Tests: Transient Impedance of Transmission Line Grounds (1012314), which are shown in Figure 1-2 [2, 3, 4]. The most promising option for the development of a suitable tower footing resistance test in frozen soil uses impulse test methods, including a pair of propagation lines. These lines have been tried in various configurations such as the reflection method shown in Figure 1-2.

1-6

Figure 1-2 Equivalent circuit for voltage impulse test method

A good understanding of the role of the propagation lines in Figure 1-2 is helpful (see Section 3, Zed-Meter Instrument Dipole Tests of Reaction Leads). This understanding will lead to good decisions about the test lead length, orientation, and configuration in difficult situations on crowded or remote rights of way. One unique feature of the Zed-Meter tester is that the propagation lines, typically 90125 m long, need not be terminated in ground rods. The surge impedance, Z, of the lines couples them to ground potential rather than to a remote ground rod. The lines behave as a surge impedance for only the time that it takes an electromagnetic surge to propagate to the end and to return. The propagation speed is a fraction of the speed of light, c, or 300 m/s. At a typical speed of 0.6 c, the 125-m lead has a reflection disturbance at about 1.4 s. If the end of the propagation line has an open circuit, the returning voltage wave will be double the initial value, and the current wave will drop from its initial value to zero.
Labeled Diagram of Instrument

The Zed-Meter testers method for measuring the impedance of transmission tower grounds does not require disconnecting the ground wire, and it is ideally suited for evaluating the lightning response or footing impedance of transmission towers, including towers in frozen soil. Photographs and a block diagram of the instrument are shown in Figures 1-3 and 1-4.

1-7

External view Figure 1-3 Photographs of the Zed-Meter instrument

Internal components

Pulse Generator

Wave Shaping Circuit

12-V Power
Supervisor Circuit

100 MS/s Digitizer Two Channels

100 MS/s Digitizer Two Channels

USB-2 Interface to Laptop Computer

Figure 1-4 Block diagram of the Zed-Meter instrument

Figure 1-5 shows the Zed-Meter lead arrangement, with the impulse source and measurement equipment at the tower base.

1-8

Figure 1-5 Equivalent circuit for Zed-Meter instruments impulse test method in frozen soil

Test Lead Arrangement


The test lead arrangement in Figure 1-5 is symmetrical. With suitably short leads, there should be little difference in the results if the connections are reversed so that the current is injected into the left-hand lead and the potential is measured in the right-hand lead. In any test that involves straight leads, the leads should be reversed as a standard practice. Also, when the leads form a dipole antenna on the surface of the earth, a dipole or antenna impedance should be measured between the two wires. Generally, obtaining a constant value in the proof test of the propagation line impedance is a good first step to obtaining a reliable value of tower footing impedance. A proof test of the propagation line impedance should produce a signal that is initially of low amplitude and then rises to a constant level and stays at that level until the reflections from the end of the propagation lines arrive. Practically, there will often be some initial oscillations that damp down in approximately 100 ns if the connections are short and secure. The time after the current waves in each direction are equal and before the reflections arrive from the remote ends is used to establish the current flowing into the tower. The voltage rise in response to this current, measured over the same interval, is divided by the current to establish the value of the dipole surge impedance in the proof test.

Sequence of Zed-Meter Operations


Under control of the National Instruments Labview program running on the laptop computer, the supervisor circuit initiates a series of high-voltage pulses by controlling the supply voltage of the pulse generator. The wave-shaping circuit flattens the crest of the pulses and sharpens the front time to make a rectangular current pulse (see Figure 1-5) between the two current sensors.

1-9

The injected currents, I1 and I2, are fairly constant regardless of the value of the tower footing resistance, RT, and the parallel OHGW impedance, ZC. The currents in each lead of the pulse generator are converted to voltage by a pair of current transducers. If there are no connections linking the Tower Lead and Current Lead terminals in Figure 14, there is no current flow. If there is a resistive connection, the currents should be equal. Two 2-channel digitizer modules running at 100 megasamples per second simultaneously measure the following:

Current I1 in the tower lead Current I2 in the current lead Voltage rise V relative to the tower lead

Under the control of the Labview program, waveforms from the digitizer are transferred to the laptop computer using the USB interface cable. The waveforms are 8000 points long, corresponding to a time of 80 s. The quiet period before each impulse is recorded with the use of a pretrigger. The pretrigger duration is presently fixed at approximately 1.5 s or 150 points. The waveforms are accumulated by the Labview program, and the process is repeated a specified number of times to collect average values of each of the three parameters. After the averaging is completed, two vectors are generated: 1) the voltage rise V relative to the tower lead, divided sample by sample by the current I1 in the tower lead, and 2) the voltage rise V relative to the tower lead, divided sample by sample by the current I2 in the current lead. Two preprogrammed time intervals are used to calculate median impedance values. The first time interval starts after the tower oscillations have damped down and ends when the reflection from the remote end of the current reaction lead arrives back at the test point. This interval is valid for both grounded and ungrounded lead configurations and also for most practical transmission line span lengths. The second time interval starts after the current in the reaction lead has settled down to a new, stable, large value and normally ends when reflections arrive from adjacent towers, at about 2 to 3 s. If the span length is short (<100 m), the measured impedance declines continuously in this period and the fitting of an L/Z time constant in postprocessing is necessary. The second interval evaluation is meaningful only for grounded test lead configurations that achieve a ground resistance of less than approximately 1000 . As a quality check, the standard deviation of the impedance over these intervals is reported on the graphic interface of the Labview program. Some user-programmable limit checks are also executed to update a visual display using stoplight (red, yellow, and green) icons.

Operating Temperature Ranges


Currently, the Zed-Meter tester performs well in ambient temperatures from -20C to +40C. There is interest in extending the temperature range at both extremes. For example, typical grounding inspection and remediation work in Canada is conducted only in winter because of better access over the terrain (see Figure 1-6). Additional specifications of the Zed-Meter tester are provided in Appendix B, Specifications.

1-10

Figure 1-6 Transport of the Zed-Meter instrument in freezing conditions

1-11

2
ZED-METER INSTRUMENT BENCH TESTS
As with any instrument, it is best to become familiar with the capabilities and operation of the Zed-Meter instrument in a comfortable indoor environment before taking the instrument into the field.

Operational Test: Charging the Battery


The Zed-Meter instrument has an internal battery that provides power to the pulse generator and other circuits. This battery is charged by connecting the ac adapter to a suitable source. The lightemitting diode indicators show the state of the charging system and whether the instrument is switched on.

Operational Test: Short Circuit


The Zed-Meter instrument has internal testing and calibration features to establish that the equipment is in good working order and connected correctly. Because it uses very fast transient signals, the Zed-Meter instruments measurements of such common references as a short length of wire or a 500- resistor provide results that correspond to the true inductive or capacitive response.
Connections and Process

The Zed-Meter instrument has three terminals: voltage (potential measurement), tower, and current. The surge energy is applied between the tower and current reaction lead terminals. All three terminals use Bayonet Neill Concelman (BNC) female connectors.

Figure 2-1 Zed-Meter instrument connection for short circuit or resistance tests

2-1

The metal collars, not the shielded internal connections, on the three BNC connectors are energized. This means that there is an irritating shock if a finger is placed between the outer metal surfaces of the tower lead and current lead BNC connectors. This arrangement is necessary to take advantage of the large radius of the sheath of a typical coaxial cable to provide lightweight, flexible cables with low surge impedance. A coaxial cable is connected across the two current outputs in the short circuit test. A BNC-type tee should be used to provide a parallel path to connect the voltage lead to the current lead terminal. The connection from the tower lead to the current lead shorts out the pulse generator. The current is limited to about 2.5 A by the internal impedance of the pulse generator and the pulse-shaping circuitry. The high current output also shortens the pulse duration. The instrument is operated to accept several waveform samples, typically 16, and to store the waveform. The interface should be adjusted to display the test currents in each channel, and the test currents should be compared with the results in Figure 2-2.

Currents into short circuit

Long-time voltage response to short circuit

Figure 2-2 Zed-Meter instruments currents and voltage for short circuit: poor operating conditions

Results for High-Quality Short Circuit

The Zed-Meter instruments reading of a short-circuit loop of 1 m should settle to a value less than 1 in 1 s, with a short circuit current level that does not damage the pulse-shaping circuit or overload the current monitoring channels. In a properly operating Zed-Meter instrument, the currents in Figure 2-2 should be identical, and the measured voltage should be free of noise. Records from a poorly operating Zed-Meter instrument highlight several possible defects. Each current waveform had a dc offset, corresponding to about +250 mA. The injected current is the difference between this offset and the peak value, giving 2250 mA for I1 and 2450 mA for I2. The 200-mA difference in peak levels recorded in each channel is considerablealmost 10%where the current transducers should agree to within 1%.

2-2

The measured voltage had a bias of approximately -5 V, which is easily addressed in processing. However, Figure 2-2 also shows a peak-to-peak noise level of 6 V, which is excessive. The tail of wave also showed evidence of quantization. Post-processing of the short-circuit waveform using Microsoft Excel software can also eliminate some of the internal noise shown these records. The overall result is a negligible change in voltage for a change in current of 2000 mA, which is satisfactory even with these defects; however, the point-by-point values of impedance are meaningless in view of the high noise level on the voltage signal. When the Zed-Meter instrument is set up and running correctly, the averaged currents and voltage for a short circuit across the output should be similar to the waveforms in Figure 2-3.

Current

Voltage

Figure 2-3 Zed-Meter currents and voltage for short circuit: correct operating conditions

The current injected into the short circuit has a well-controlled 10-90% rise time of about 130 ns, with minimal overshoot. The voltage in response to this current has a pulse duration that matches the time that the current is rising. This indicates a nearly pure inductive response. For the waveform shown, the rate of change of current dI/dt is 2.6 A / 0.16 s or 16 106 A/s. The peak voltage in response to this change in current can be used to compute the inductance of the shortcircuit cable system. In this case, the inductance is given by 30 V/ (16 106 A/s, or 1.9 H. This is equivalent to about 2 m of connection length. Figure 2-4 shows the impedance versus time plot from the waveforms in Figure 2-3.

2-3

Figure 2-4 Zed-Meter instruments impedance display for short circuit: correct operating conditions

Before the trigger, the ratio of voltage to current has no meaning. Just after the pulse, the impedance in Figure 2-4 peaks at just greater than 100 , then falls within 200 ns to a value less than 1 . The local standard deviation, taken over a 100-ns sample interval, also falls with time and stabilizes at a level of approximately 0.02 . The Zed-Meter instrument can sometimes be used to measure transmission towers with low footing impedance of less than 5 . The noise floor of the measurement system is approximately 0.7 in the sweet spot that starts at about 500 ns and continues indefinitely in Figure 2-4.
Results for Poor-Quality Short Circuit

The Zed-Meter instrument produces and measures high-frequency pulse energy that is intended to flow through radio frequency connectors and coaxial cables. One factor in successful measurements is getting this energy from the instrument into the transmission tower under test without introducing signal losses. Poor-quality connectors and connections introduce undesired series and shunt impedance into the signal path. This leads to undesirable signal oscillations that take a long time to damp out, and it can also raise the minimum impedance that the instrument can measure.

2-4

It is useful to conduct the short circuit test using the clamps and cables that will actually be used to attach the tower lead to the tower. If excessive oscillation (>100 ns) on the voltage or current measurements is noted in the results, the connection method should be reevaluated.

Operational Test: Fixed Resistance


The Zed-Meter instruments reading of a fixed resistance should reach the low-frequency value in about 10 s, depending on the characteristics of the resistor. The better the wiring configuration, the faster the Zed-Meter will settle to this value. It is recommended that high-quality low-inductance resistors be purchased or constructed for calibration of the Zed-Meter instrument. The value should be in the range of 10500 . One possibility is standard, matching 50- or 75- resistors, which typically have a constant impedance (or low-voltage standing wave ratio) at frequencies up to 1 GHz. Alternately, a lowinductance resistor can be constructed by soldering 10 or more metal film resistors in parallel. Generally, wire-wound resistors will have internal inductance that will make them unsuitable for the calibration process. The selected resistor should ideally be mounted in an adapter with a pair of insulated BNC male bulkhead connectors, using only the ground leads.
Reference Result: 54- Calibration Resistor

Bench measurements of a Zed-Meter instrument, using a 1.3-s pulse width, were conducted on a 54- resistor mounted in a small coaxial box. The setup and test results are shown in Figure 2-5.

2-5

Measured voltage, 105 V after 100 ns

Test setup (2006 version)

Measured current, 2 A

Calculated Z(t) and standard deviation over 16 samples

Figure 2-5 Zed-Meter instrument test setup and results for a 54- calibration resistor

In this case, the voltage across the resistor stabilized to a constant value with a time constant, , of about 50 ns. This time constant, multiplied by the measured resistance of 54 , gives the inductance of the measurement loop, consisting of two 1-m wire leads, as shown in the test setup. The inductance works out to about 2.7 H, a little higher than the expected 1 H/m for straight leads. Figure 2-5 also shows a plot of the calculated impedance profile, Z(t), from each of the measured currents. Using logarithmic scales, it is also possible to show the standard deviation of the overall result. The standard deviation is a function of several factors, including the inherent noise in the measurement, the degree of agreement between the two current sensors, and the settling time of the wiring. In this case, the standard deviation is initially high, but it falls below 1 at about 200 ns, and it remains less than 1 to the end of the pulse waveform at 1200 ns.

2-6

The time period at which the standard deviation of the reading falls below a threshold and remains relatively constant is called a sweet spot with high signal-to-noise ratio. In original iterations of the Zed-Meter instrument, the sweet spot was identified by eye, sometimes with different results, depending on which eye was used. Modifications to the instrument have focused on making the sweet spot as wide as feasible, consistent with retaining a rapid test time. On the bench, the sweet spot should be quite long.
Reference Result: 500- Resistor

The reference resistor test is set up by fitting coaxial cabletobanana jack and binding post adapters to each of the two current outputs. The calibration resistor is attached between the ground (black) terminals of the binding posts. A BNC tee is used to provide a parallel path to connect the potential measurement lead to the current reaction lead terminal using a short length of coaxial cable, terminated in BNC male connectors. Alternately, a third coaxial to banana jack/binding post adapter can be used with a short length of standard wire, connecting the ground (black) end of the potential measurement binding post to the ground (black) binding post of the current reaction lead. The use of a coaxial cable and BNC tee is preferred. The voltage and current signals should be averaged for 16 impulses. The results can be stored as text (.txt) files for analysis. For example, Figure 2-6 shows the measured voltages and currents for a 500- wirewound resistor.

Figure 2-6 Measured voltage and currents into wirewound 500- resistor showing oscillations

The standard deviation of the measured impedances in Figure 2-7 was approximately 10 when both signals were considered, and the standard deviations reduced to 13 when they were computed individually. It took about 500 ns for the oscillations in the currents to damp down sufficiently to start the sweet spot of the record. This is a function of the choice of resistor value

2-7

(500 ) and its internal capacitance of about 50 pF, giving a resonant frequency of about 17 MHz with the 2-H wire loop.

Short-term response with standard deviation

Detail of long-term response, showing current transformer droop

Figure 2-7 Zed-Meter instrument measurement of wirewound 500- resistor on two time scales

The detail of the long-term response of the Zed-Meter instruments records in Figure 2-7 shows two defects. First, the impedance seems to increase with time. This is a result of the lowfrequency limitations of the selected current transformers. They exhibit a droop, or signal loss, of about 0.1% per microsecond. This means that, compared to the initial value, the measured current drops by about 8% at a time of 80 s. For a constant applied source voltage provided by the Zed-Meter instrument, a decrease in measured current corresponds to an increase in impedancefrom 480 to 530 for one current transformer and from 500 to 545 for the other. Each current transformer has a slightly different droop that can be compensated in software. Second, each current transformer indicates a slightly different value of current, leading to a 20- difference in the measured value of the reference resistor. The transient impedance of the tested resistor at late time varied from 480 to 530 using Z1, and it was 3% higher using Z2. It is expected that these two readings will be within 1% of each other. The software process for correcting the droop and scale factors for the Zed-Meter instrument is under development. When comparing the results in Figure 2-7 with those in Figure 2-5, remember that both setups have approximately the same lead inductance. The L/R time constant is far less of a factor for the 500- resistance than for the 54- resistance, and the lead inductance dominates the response of the short-circuit test.

2-8

Practice Makes Perfect


The Zed-Meter instrument can be connected to the tower and current leads in a number of different ways. This section has shown that the type of wire, the type of connectors, and the instrument settings can all affect the quality of the results. A familiarization period on the bench, testing impedance of resistors with good and bad wiring practice, is helpful to establish good habits in the field. This is especially important when measuring towers with low impedance. If the connection leads are too long, the Zed-Meter instrument will simply report the dynamic impedance of the leads themselves. It is possible to compensate for sloppy wiring practice near the tower by extending the length of reaction leads and analyzing the Z(t) values at later times. However, it is better to start with a good waveform.

2-9

3
ZED-METER INSTRUMENT DIPOLE TESTS OF REACTION LEADS
Function of Reaction and Potential Leads
The Zed-Meter instrument relies on the fact that the surge impedance of an insulated wire, laid close to the ground, is constant and has a value in the range of 400700 . Two propagation lines are used. The first propagation line is a current reaction lead. The impulse source is placed between the tower base and this lead. The current injected into the tower base will be a faithful copy of the current launched down the reaction lead after some initial oscillations up and down the tower have decayed away. If a tower has a resonant structure with guy wires, it can take a bit longer for this to occur, and the propagation lines might have to be extended to 125 m or more. The second propagation line is the potential lead. The tower base potential is measured by this insulated wire, which is also coupled to ground by its surge impedance. This lead impedance plays less of a role in the accuracy of the measurement because the input impedance of the measuring circuit is high. However, any ac or high-frequency noise picked up by this horizontal antenna must be rejected by the measurement circuit. This is done by averaging many impulses, with the expectation that the noise signals are not correlated to the test wave. If the surge impedance of the current reaction lead is too high (as can be the case over highresistivity grounds like snow and ice), less current is injected, and the potential rise at the tower base becomes too low. In addition, if there is a high noise level on the potential reference lead, more impulses must be averaged. In severe cases, potentials on the ungrounded leads might be too high (>50 V) to be handled without personal protective equipment. The configuration and selection of reaction and potential lead layouts has had an extensive period of development. Appendix D, Hardware Evaluation, summarizes the options that have been considered.

Dipole Impedance Test Method


Although it adds to the field test time, it is recommended that a dipole test be conducted on the leads every time, before taking a measurement of the tower footing impedance. The dipole impedance of the two test leads can be measured well, and it should be approximately constant whether the leads are oriented at 90 or 180. The results make the most sense if both leads run in straight lines and both leads have the same type of termination, either driven thin rods or open circuit. The dipole test setup connects the Zed-Meter instruments voltage (potential measurement) terminal to the current reaction lead using a BNC-type tee and a short coaxial cable. The current lead terminal is also connected to the current reaction lead. The central terminal feeds the potential lead (see Figure 3-1).

3-1

90-125 m

90-125 m

Figure 3-1 Connection diagram for dipole impedance test

It should be possible to reverse the red and blue leads in Figure 3-1 and still obtain the same impedance result. This is worth checking, especially in areas in which there is considerable induced noise on the cables or in cases in which it was easier to ground one cable. Induced noise will probably be measured on the pretrigger record of the digitizers. Ideally, everything before the pulse fires should read zero. In practice, the meter often picks up currents of a few mA resulting from AM radio broadcasts. Averaging multiple test shots should reduce this interference to a low level compared to the 400-mA test signal. If the dipole impedance values fluctuate from one test series to another, or if there is a large difference when the leads are reversed, it might be appropriate to increase the number of samples being averaged or to terminate the far ends in ground rods if they are floating in the first test series. Depending on the vegetation, the reaction and potential reference leads can be near the surface of the soil, or they can be suspended a meter or more off the ground. A lead close to the ground will have lower and more constant surge impedance as well a- a slower propagation time. These factors will give a higher quality measurement. For that reason, the coaxial cable should be placed as close to the ground as possible. If possible, it is best to walk back along the lead to force it close to the ground. If this is not feasible, the lead length should be increased.

Variation of Dipole Impedance with Lead Orientation and Length


The angle between the two test leads can vary from 180 to 45 without making much change in the dipole impedance. Therefore, the lead orientation will not have a major effect on the results if 3-m leads running parallel to one anohter are avoided.

3-2

Lead lengths depend on soil resistivity (see Appendix E, The Current Reaction Lead, for details). In order to have a measurement with a period of constant injected current, lead lengths should be increased in areas with higher soil resistivity or in areas where the lead must be draped in vegetation well above the earth surface. Observe the following guidance for required length for the current reaction lead:

Use a minimum 90-m lead length for soil types with resistivity 100 < < 500 m Lead lengths of approximately 125 m should be considered in any of the following circumstances: High soil resistivity ( >500 m) is anticipated. Soil is frozen. It is not feasible to lay the lead close to the ground. The tower has guy wires. The tower is isolated from the OHGWs by insulators. A configuration with two or more possible issuesfor example, a guyed tower on frozen soilmight work better with lead lengths of 150 m.

Dipole Test Results, 90-m Versus 130-m Lead Length


The bench tests in Section 2 show some important aspects about Zed-Meter instrument testing. When measuring low values of resistance, the effects of series test lead inductance take some time to damp out before the sweet spot with constant voltage and current enables a valid calculation. Measuring resistances of 500 can also lead to high-frequency oscillations. The advantage of a 130-m cable over a 90-m cable was evaluated in a test series in freezing conditions, which would be the worst case, with extremely high surface resistivity in the frozen soil and snow layer.
Pulse G enerator Measures current to the left wire Insulated 90 m Wire Me asure s curr ent to the right wire Insul ated 90-m Wire

Snow and/o r Frozen S oil La yer

Pulse G enerator Measures current to the l eft wire Insulate d 130 m Wire Me asure s cur rent to the right wire Insulated 1 30-m Wire

Snow and/or Frozen Soil L ayer

Figure 3-2 Comparison of 90-m and 130-m RG-58 coaxial cable laid on surface of frozen soil

3-3

Because it is pointless to drive ground rods into frozen soil, the leads were left floating. This meant that the traveling wave reflections from the ends of the cables were marked by clear reductions or reversals of current. In Figure 3-2, the sweet spot of constant and equal current in each leg starts at about 500 ns in all cases. The suitable evaluation time continues to 800 ns for the 90-m leads, corresponding to a propagation velocity of 0.75 c. The sweet spot increases to 1100 ns for the 130-m leads, also corresponding to 0.75 c. This is a worthwhile gain for an extra minute or two of walking time in each direction.

Grounding of Reaction and Potential Leads


Test experience has shown that, in most cases, there is no need to ground the ends of either the current reaction lead or the reference potential lead. Both wires are earthed through their surge impedance to ground, which is typically about 500 . It can be fairly difficult to drive a temporary rod into the local soil to obtain a resistance lower than 500 . If the termination resistance is higher or lower than the wire surge impedance, a negative or positive current reflection coefficient will appear at the end of the current reaction lead. This means that, after two-way travel time along the wire, the current will drop or increase to a new value. Several reflections can occur before the current settles to a new, constant value given by the pulse generator output voltage divided by the ground termination resistance.
Reasons for Grounding the Current Reaction Lead

The initial reading of impedance is taken while the test current is constant and limited by the surge impedance of the current reaction lead. If the current reaction lead is terminated in a ground rod, the test current will settle down to a new value after a few round trips of the traveling waves. It is then possible to take a second reading of impedance in a different, slower time window. The degree of agreement between the two values, one in the 0.51 s range and another in the 25 s range, will help to characterize the soil better and thus improve modeling. In order to obtain a useful second reading, the resistance of the remote termination of the current reaction lead must typically be <1000 . It can be useful to use a standard or clamp-on earth resistance tester to ensure that the driven rod meets this specification.
Reasons for Grounding the Remote Potential Lead

The main reason for grounding the remote end of the remote potential lead is to reduce the effects of electrostatic pickup. In some cases, continuous applications of the surge pulse can also cause the lead to charge up, leading to an offset in the measured signal that must be corrected in software. If the static charge builds up to an excessive level, it can overload the voltage channel and cause an over-range indication. The remote end of the potential lead should be grounded to preserve signal integrity if the measured potential relative to the tower exceeds about 10 Vrms or if the tower has an impedance of less than about 5 .

3-4

Troubleshooting the Zed-Meter Instruments Dipole Test Results


Verifying a Safe Work Environment

The first problem that can occur when setting up a dipole test beneath a transmission line is that there can be excessive voltage on the leads. Good practice is to treat the coaxial cables as live until they are proven safe with a voltage reading. If the ac or dc voltage exceeds local safety limits (such as 50 Vrms), the test will require special care, including the use of insulated rubber gloves and covers rated for grounding tests by the local utility safety practices.
Symptoms of Excessive Noise Level

If the induced noise on the cables is a considerable fraction of the test voltage (200 V) or the circulating currents are on the order of 50 mA, there are two possible problems. The first problem is that the sum of the signal and noise will exceed the automatically selected voltage or current scale. This over-range condition should be recognized by the instrument software, but waveforms should be examined carefully for clipping. Another symptom of clipping is a very low standard deviation in the measurement results. If clipping is noted, it might be necessary to restart the instruments auto-range function (see How-To: Start Measurement Process in Section 5) in order to establish the most appropriate internal measurement scales. The second problem with noise on the cables is easier to fix. If there is considerable noise on the leads but it is not correlated with the Zed-Meter instruments free-running impulse source, taking multiple records and averaging them will reduce the noise level within seconds. General experience has shown that 16 waveform averages are sufficient for reducing noise beneath transmission lines 765 kV. The presence of a nonzero signal amplitude in the pretrigger period before the step of voltage and current is an indication that the averaging is not sufficient for the local noise level.
Symptoms of Problems in the Lead Layout

The dipole test should indicate the same current in each leg. The currents and voltage should rise quickly, stabilize within at most 500 ns, and remain relatively constant for at least another 300 ns. This will result in a measured surge impedance value with a low relative standard deviation from the instrument readout.
Symptoms of Conductors Running in Parallel

If one or both of the leads cross areas of different resistivity, there can be one or more perturbations in the current records. For example, leads that run in low-resistivity soil for some distance and then follow gravel construction roads have had changes in impedance of more than 100 . Running the lead over obstructions can also introduce bumps in the dipole results. The reasons for these problems relate to changes in the surge impedance of the cable as a function of high gravel compared to typical soils or extra height above ground. The remediation is fairly simple: reroute the lead so that it has consistent height and resistivity. Rerouting the lead should introduce fairly gentle curves, if possible. However, if a 90 bend is necessary to maintain constant height and soil type, the numerical modeling for this project suggests that the effect of the bend will be much less than the effect of different height or soil.

3-5

Symptoms of Problems in the Lead Terminations

It is unlikely that the driven rods at each end of the dipole will have the same resistance values. This means that some asymmetry in the leg currents can occur after the return of reflections from the terminated ends. Because this is beyond the sweet spot, the differences can be ignored. In fact, it can be useful to exploit this effect deliberately by leaving one of the two ends open to ensure that the return time is marked clearly in the analysis.

3-6

4
ZED-METER INSTRUMENT TESTS ON TRANSMISSION TOWERS
Although it is not strictly necessary to conduct a dipole test before testing a tower, the leads must always be checked with a voltmeter for standing potential before connecting the Zed-Meter instrument. This is done by measuring the voltage potential between the tower and the leads using a handheld voltmeter as shown in Figure 4-1.
Voltmeter

>90 Meter Lead

Figure 4-1 Connections to measure voltage potential on leads

If the potential exceeds 50 Vrms, most utilities call for the use of insulating gloves or other countermeasures. The Zed-Meter instrument itself can generate good results even if the induced pickup exceeds 100 V, because the current transducers are dielectrically isolated from the leads.

Connection Diagram for Impedance Test


The basic layout of the current reaction and remote potential leads was described in Section 3, Zed-Meter Instrument Dipole Tests of Reaction Leads, along with a recommended method for validating the impedance of the configuration. Deviations from the recommended practice, 125-m leads in two directions, 180 apart along the right of way, are described in Orientation of Reaction and Potential Leads in this section. When a good-quality result is obtained in the dipole test for the leads, say 400700 with 20- standard deviation, the Zed-Meter is ready for connection to the tower (see Figure 4-2).

4-1

90-125 m

90-125 m

Figure 4-2 Connection of Zed-Meter to reaction lead, remote potential lead, and tower leg

Some interesting things should happen when the dipole test is changed to a tower test. First, the current injected into the tower should be about twice the current into the dipole. The surge impedance of one leg of the dipole is twice that of the full dipole. This impedance is typically much greater than the impedance of the tower under test. Second, the potential rise of the tower base relative to the ground reference potential lead can initially be quite high, but it soon settles down to a relatively constant value. The Zed-Meter instrument is fast enough to see the effects of currents traveling up and down tower legs and guy wires. This also means that the connection of the tower lead to the base of the tower steel should be as short as possible, and it should make a good electromagnetic connection at high frequency. It might be necessary to initiate the auto-range function of the Zed-Meter to adjust to the new signal levels (see How-To: Start Measurement Process Section 5). This is done by changing the line name or structure number when starting a test. The software uses a change in either field to initiate the auto-range sequence.

Preferred Type of Tower Connection


A number of methods have been explored for making a good high-frequency connection from the Zed-Meter instrument to the transmission tower. Figure 4-3 shows the following three examples:

An alligator clip with 0.5-m wire lead and banana plug A citizens band (CB) radio antenna clamp and coaxial cable (intended for truck mirror mount) A welding clamp fitted with a swaged BNC female connector

4-2

2003 (American Electric Power), alligator clip to lattice tower legdefect: overshoot

2004 (Public Service Electric and Gas), CB radio clamp-on antenna mount

2008 (Manitoba Hydro), modified welding clamp Figure 4-3 Improvements in radio frequency connection of Zed-Meter to transmission towers, 20032008

4-3

Generally, the best results are obtained when the two measured currents are in close agreement and not oscillating too much after the initial rise. Figure 4-3 shows that the alligator clip is poor in this respect, with 100% overshoot; the CB radio antenna clamp is good, and the modified welding clamp is quite good. Although the high-frequency performance of the CB antenna radio mount is theoretically better than the modified welding clamp and its cost is much lower, in practice the device could be loosened and tightened only about 50 times before it broke in half. The welding clamp (see Figure 4-4) is likely to give longer service life and is recommended as the best choice at this time.

Connection to lattice tower Figure 4-4 Close-ups of modified welding clamp with swaged BNC female connector

Connection to stud on steel pole

Preferred Location of Tower Connection


Numerical simulations have suggested that the height of current reaction plays a role in the quality of the results. The effect of lead height on the propagation velocity and impedance of the reaction lead was detailed in Section 3, Zed-Meter Instrument Dipole Tests of Reaction Leads. A dipole test establishes whether the leads are long enough to compensate for this effect. The numerical simulations also found a second factor. The height and length of the connection from the instruments tower lead to the tower plays a considerable role in the quality of the results. This seems to be more important than the height of the reaction lead further away from the tower. An example of a desirable, tight connection, with short leads and close to the ground, is shown in Figure 4-5.

4-4

Figure 4-5 Preferred Zed-Meter instrument connection to lattice tower

Figure 4-6 shows that, when the wiring is tight, there is little difference in the results with a connection to the tower 0.1 m off the ground.
Sim# I1-C-L01-GR-GW-50 40 Vmeas / I meas Measured Impedance: Vmeas / I meas [Ohm] 40 Vmeas / I meas Measured Impedance: Vmeas / I meas [Ohm] 35 30 25 20 15 10 5 0 Sim# I1-C-L1-GR-GW-50 BIS

=50 m h=0.1 m: 5 h=1 m: 4

35 30 25 20 15 10 5 0

Median 450-650 ns = 5 0 0.5 1 1.5 Time [microseconds] 2 2.5

Median 450-650 ns = 4 0 0.5 1 1.5 Time [microseconds] 2 2.5

Sim# I1-C-L01-GR-GW-1000 100 Measured Impedance: Vmeas / I meas [Ohm]


100

Sim# I1-C-L1-GR-GW-1000 BIS

=1000 m h=0.1 m: 46 h=1 m: 43

80 70 60 50 40 30 20 10 0 0 0.5 Median 450-650 ns = 46 1 1.5 Time [microseconds] 2 2.5

Measured Impedance: Vmeas / I meas [Ohm]

90

Vmeas / I meas

90 80 70 60 50 40 30 20 10 0 0 0.5

Vmeas / I meas

Median 450-650 ns = 43 1 1.5 Time [microseconds] 2 2.5

Reaction lead 0.1 m above ground

Reaction lead 1 m above ground, except 0.1 m at tower

Figure 4-6 Effect of reaction lead height on measured impedance with 90-m leads using good (tight) wiring practice as a function of soil resistivity and height above ground h

4-5

In contrast, Figure 4-7 shows simulations for a detailed electromagnetic model of a lattice tower. These simulations were performed for three values of soil resistivityextremely low (50 m), high (1000 m), and extremely high (20,000 m).

Bottom leg of lattice tower geometry in NEC-4 model

Sim# I1-C-L01-GR-GW-50 40 Vmeas / I meas [Ohm] [Ohm] 35 30 25 20 15 10 5 Median 450-650 ns = 5 0 0 0.5 1 1.5 Time [microseconds] 2 2.5 0 0 0.5 35 30 25 20 15 10 5 40

Sim# I1-C-L1-GR-GW-50 Vmeas / I meas

meas

meas

Measured Impedance: V

Good: 5 Loose: 9

Measured Impedance: V

meas

=50 m

/I

/I

meas

Median 450-650 ns = 9 1 1.5 Time [microseconds] 2 2.5

Sim# I1-C-L01-GR-GW-1000 100 V [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0.5 Median 450-650 ns = 46 1 1.5 Time [microseconds] 2 2.5
meas

Sim# I1-C-L1-GR-GW-1000 100 /I


meas

V [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0.5

meas

/I

meas

meas

meas

Measured Impedance: V

Good: 46 Loose: 43

Measured Impedance: V

meas

=1000 m

/I

/I

meas

Median 450-650 ns = 43 1 1.5 Time [microseconds] 2 2.5

Sim# I1-C-L01-GR-GW-20000 500 V [Ohm] 400


meas

Sim# I1-C-L1-GR-GW-20000 500 /I


meas

V [Ohm] 400

meas

/I

meas

meas

meas

Measured Impedance: V

Good: 261 Loose: 219

200

Measured Impedance: V

meas

=20000 m

/I

300

/I

meas

300

200

100

100

0 Median 450-650 ns = 261 -100 0 0.1 0.2 0.3 0.4 0.5 Time [microseconds] 0.6 0.7 0.8

0 Median 450-650 ns = 219 -100 0 0.1 0.2 0.3 0.4 0.5 Time [microseconds] 0.6 0.7 0.8

Good (tight) wiring practice

Poor (loose) wiring practice

Figure 4-7 Effect of wiring practice at tower base on measured impedance with 90-m reaction leads

4-6

As Figure 4-7 shows, there is an extensive and large disturbance in the Z(t) profiles during the sweet spot (about 450650 ns) when using poor wiring practice, with a connection 1 m above the ground. This effect is greatest when the soil resistivity is low; the variation in Z(t) seems to have a constant amplitude. In practice, the differences between clamp locations are somewhat less than those suggested by simulations. Even so, minimizing the length of the tower lead is probably as important as using a high-quality clamp with low impedance at high frequency for obtaining satisfactory results.

Orientation of Reaction and Potential Leads


The normal procedure is to run the current reaction lead in one direction along the right of way and to run the remote potential lead in the other direction (see the right side of Figure 4-8). The calculation of tower base voltage with this configuration gives an estimate of tower-base potential that is closest to the true situation of a vertical lighting flash to the tower top. It is also the most practical orientation in a typical right of way because permission to run the leads through adjacent properties is not necessary. If access to the right of way along one direction is restricted, perhaps by a road, the alternative of running the current reaction lead at right angles to the line direction, 90 to the potential lead, can also be used (see the left side of Figure 4-8). This configuration is more common for grounding tests at a single, fixed frequency because it reduces steady-state mutual coupling effects among leads. The effect of lead orientation is most noticeable when testing towers with low footing impedance. The differences amount to 12 , which is a considerable fraction of a 5- result but can be neglected if the tower impedance is measured to be 25 .

4-7

Current reaction lead 90 to line; potential lead along line (toward)


Sim# I1-A-L01-GR-GW-50 40 Vmeas / I meas Measured Impedance: Vmeas / I meas [Ohm]

Current reaction lead along line (away); potential measurement along line (toward)
Sim# I1-C-L01-GR-GW-50 40 Vmeas / I meas Measured Impedance: Vmeas / I meas [Ohm] 35 30 25 20 15 10 5 0

35 30 25 20 15 10 5 0

Median 450-650 ns = 6 0 0.5 1 1.5 Time [microseconds] 2 2.5

Median 450-650 ns = 5 0 0.5 1 1.5 Time [microseconds] 2 2.5

Z(t) in 50 m soil, leads at 90

Z(t) in 50 m soil, leads at 180 along right of way

Figure 4-8 Numerical modeling of Zed-Meter instrument result with 90 and 180 lead orientation

In some cases, it is desirable to run the leads at right angles to one another. A preferred alternate configuration calls for the current lead to run at right angles to the OHGW direction and the voltage lead to run along the right of way. Figure 4-8 shows that there is a slight increase of about 1 when the leads run at 90, compared to the reference case. This amounts to an increase of about 20% for a well-grounded tower with a 5- impedance. Measurements with test leads at each orientation were conducted on a pair of towers, Manitoba Hydro DC 1/2 tower 1966 and DC 3/4 tower 1966. The impedances were low, and the best standard deviation values of just under 1 were achieved from 500900 ns (see Figure 4-9).

4-8

Figure 4-9 Typical Zed-Meter instrument result: 2.9 for Manitoba Hydro guyed-V DC 3/4 bipole tower 1966

The effect of lead orientation was tested in the presence of induced dc voltages that were surprisingly low, considering the line voltage of 500 kVdc (see Tables 4-1 and 4-2).
Table 4-1 Effects of lead orientation on Zed-Meter instrument results for DC 1/2 tower 1966 Leads at 180 Normal Reversed Leads at 90 Normal Reversed

Impedance () Median

2.64

2.49

2.13 2.1

2.14

2.04

2.50

2.32 2.5

2.59

Table 4-2 Effects of lead orientation on Zed-Meter instrument results for DC 3/4 tower 1966 Leads at 90 Connect to Tower Steel Normal Reversed Leads at 180 Connect to Tower Base Normal Reversed

Impedance () Median ()

4.39

4.03 3.7

3.31

3.27

3.45 3.5

3.09

2.93 3.0

2.71

3.11

4-9

The results of this test series confirm the numerical modeling in Figure 4-8. The measured impedance with leads in the reference configuration (180 along right of way) was about 20% lower than the results with leads at 90 in both cases. In addition to changes in lead orientation and position, the influence of the connection to the base plate rather than the tower steel was checked. A connection to tower steel, 1 m above the base plate (see Figure 4-10), gave results that were about 0.2 higher than the desired connection with tight wiring practice to the base plate of the tower.

Typical tower base

Typical overall view, DC 1/2 tower 1967

Figure 4-10 Detail of preferred Zed-Meter instrument connection to tower

High-voltage dc lines have static voltage at ground level that might influence test procedures in ways that are different from ac electrostatic fields. These tests were needed to establish whether the static pick-up was low enough for normal handling of the leads and proper operation of the equipment.

Dealing with Obstructions When Laying Out Leads


The first obstruction that will be noted when laying out potential and current reaction leads at the preferred 180 angle is that one of the leads can run past a tower leg that is not being connected to the Zed-Meter instrument. The lead should be offset from the tower by a distance of >1 m. The tower will tend to reduce the measured potential rise, so it is better if the current lead runs past the unenergized leg. To check the influence, however, it is good practice to take measurements with leads reversed, by interchanging the coaxial connectors for the current and voltage leads. The two results should be in good agreement; the degree by which that they differ indicates the uncertainty in the selected test layout.

4-10

Twists and Turns: The Meander Line

In some situations (for example, where the leads are oriented at 90 to the line direction) it can be impossible to continue laying the leads beyond the edge of the transmission line right of way. In these cases, simulations show that a meander configuration with one or more 90 bends still gives good results. Possible meander configurations are shown in Figure 4-11, and the resultant waveforms are shown in Figures 4-12 and 4-13.

ZZ1

ZZ2

ZZ3

ZZ4

Figure 4-11 Four possible meander configurations of current reaction and remote potential leads
40 Measured Impedance: V / Imeas [Ohm] meas 35 30 25 20 15 10 5 0 ZZ1-L01-GR-GW-50: 5 ZZ2-L01-GR-GW-50: 4 ZZ3-L01-GR-GW-50: 5 ZZ4-L01-GR-GW-50: 5

0.5

1 1.5 Time [microseconds]

2.5

Figure 4-12 Effect of meander-line configurations ZZ1 to ZZ4 on Zed-Meter instrument impedance profile Z(t) for 50 m soil resistivity, lattice tower 4-11

100 Measured Impedance: Vmeas / I meas [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0.5 1 1.5 Time [microseconds] 2 2.5 ZZ1-L01-GR-GW-1000: 47 ZZ2-L01-GR-GW-1000: 48 ZZ3-L01-GR-GW-1000: 47 ZZ4-L01-GR-GW-1000: 46

Figure 4-13 Effect of meander-line configurations ZZ1 to ZZ4 on Zed-Meter instrument impedance profile Z(t) for 1000 m soil resistivity, lattice tower

Table 4-3 shows the errors introduced when the potential leads (ZZ1), current leads (ZZ3), and both leads (ZZ2 and ZZ4) have the meander or zigzag layout. Even for regions with high soil resistivity, where the inherent measurement uncertainty is high, all four zigzag patterns give results that are practically the same as the reference configuration.
Table 4-3 Computed Zed-Meter instrument results (450650 ns) for 100-m meander configurations Soil resistivity Reference configuration 50 m 5 1000 m 46 20,000 m 261

ZZ1 ZZ2 ZZ3 ZZ4

5 4 5 5

47 48 47 46

262 252 261 248

4-12

Using Shorter Potential Leads

If one lead must be shorter than the other, it makes sense that the short lead should be the one used to provide the remote zero potential reference. If the current reaction lead is shortened, the entire measurement process can be compromised. An early return of the reflection from the far end can occur before the initial oscillations have damped out, leading to a sweet spot of zero length. If, instead, the tower base potential rise is measured relative to a nearby ground, a large fraction of the total voltage rise will still be recorded. Numerical simulations have confirmed that, if one lead must be shorter than the other, it is better that the short lead be used to measure potential. The potential falls off rather quicklyfor example, to a level of about 40% at a distance that is one diameter away from a four-leg footing. This can even be compensated exactly if the equivalent radius of the ground electrode is known. It a short potential lead is used, it should be terminated as well as possible in a ground rod. Four configurations were tested in the numerical models with NEC-4, using a 100-m current reaction lead. The reference case has a 100-m potential lead at 180, oriented along the transmission right of way. A tight wiring geometry was assumed at the tower base, reflecting the extra care that should be taken in all other measurement aspects when compromising in one aspect. The results in Figure 4-14 are quite encouraging. As expected, the potential measured on the shortened lead is less than that on the 100-m long reference, but it is not reduced by much. The most sensitive case is when the soil resistivity is high. It would be feasible to derive a correction table as a function of soil resistivity for short potential leads that could be applied, for example, to the 25-m case relative to a 100-m case, or even a case in which a vertical lightning flash strikes the same tower and grounding system. However, in order to apply this table, one would need to measure the soil resistivity. At present, this measurement is a time-consuming process. Additional development of the Zed-Meter instruments dipole test could lead to a successful method that separates the effects of soil resistivity from the effects of lead height over ground. This could eventually supply the missing link needed to implement this improvement.

4-13

Soil resistivity and resultant Impedance


40

Reference: 100 m potential lead

Results with 75 m, 50 m, and 25 m

40 V
meas

/I

meas

Measured Impedance: Vmeas / I meas [Ohm]

35 30 25 20 15 10 5 Median 450-650 ns = 5 0

Measured Impedance: Vmeas / I meas [Ohm]

35 30 25 20 15 10 5

C-PL75-L01-GR-GW-50: 5 C-PL50-L01-GR-GW-50: 5 C-PL25-L01-GR-GW-50: 5

50 m All: 5

Median 450-650 ns 0 2.5 0 0.5 1 1.5 Time [microseconds] 2 2.5

0.5

1 1.5 Time [microseconds]

100 V Measured Impedance: Vmeas / I meas [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0.5 Median 450-650 ns = 46 1 1.5 Time [microseconds] 2 2.5
meas

100 /I
meas

Measured Impedance: Vmeas / I meas [Ohm]

90 80 70 60 50 40 30 20 10 0 0 0.5

C-PL75-L01-GR-GW-1000: 46 C-PL50-L01-GR-GW-1000: 45 C-PL25-L01-GR-GW-1000: 43

1000 m 100m 46 75 m 46 50 m 45 25 m 43

Median 450-650 ns 1 1.5 Time [microseconds] 2 2.5

500 V Measured Impedance: Vmeas / I meas [Ohm] 400


meas

/I

500
meas

Measured Impedance: Vmeas / I meas [Ohm]

400

C-PL75-L01-GR-GW-20000: 257 C-PL50-L01-GR-GW-20000: 238 C-PL25-L01-GR-GW-20000: 202

20,000 m 100m 261 75 m 257 50 m 238 25 m 202

300

300

200

200

100

100

0 Median 450-650 ns = 261 -100 0 0.1 0.2 0.3 0.4 0.5 Time [microseconds] 0.6 0.7 0.8

0 Median 450-650 ns -100 0 0.1 0.2 0.3 0.4 0.5 Time [microseconds] 0.6 0.7 0.8

Figure 4-14 Effect of shortened remote potential lead on Zed-Meter instrument result for three soil resistivities

4-14

Vehicles in Proximity to Leads

At a wood-pole gulf-port structure (Manitoba Hydro CW8 tower 219), a comparison test evaluated sensitivity to local conducting objects. In this case, a large utility vehicle (a Chevrolet Suburban) was parked on the current reaction lead. Table 4-4 shows all the test results for this case.
Table 4-4 Results of Zed-Meter instrument tests on CW8 tower 219 Impedance, 400920 ns Leads Normal Trial 1 Trial 2 Trial 1 Leads Reversed Trial 2 Median (Std Dev)

Tower 219 Tower 219, 90

12.5 13.2

12.8 12.4

12.5 12.0

12.9
12.4

12.6 (0.2) 12.4 (0.5)

These are all highly consistent results, even for the test in which the Suburban was parked on the test lead (marked in bold font). It was apparent where the vehicle was located, indicated by a extra peak in the Z(t) profile (see Figure 4-15).

Results without truck on lead

Results with truck parked on reaction lead

Figure 4-15 Comparison of Zed-Meter instrument test results for CW8 tower 219, 90 lead orientation

The vehicle caused an extra peak of more than 100 at 100 ns in Figure 4-15, but the median impedance settled down to the same value, 12.4 , as in the other tests. Nevertheless, this is not a recommended practice. By way of historical interest, one of the first methods evaluated to park the test current was to use the capacitance of the line truck rather than the constant surge impedance of the reaction line. Unfortunately, there was not enough current flow into the truck to obtain a valid measurement.

4-15

Considerations for Guyed Towers


Guyed transmission structures, such as the Georgia Power 230-kV structure shown in Figure 416, have a resonant response associated with the distance up the tower and out to each of the individual guy anchors. The guy anchors can provide some or most of the grounding for the tower, depending on how many there are and how deep the guy anchors are driven. Each guy wire is a straight, simple path connected to the tower, leading to an electromagnetic response that is poorly damped.

Figure 4-16 Guyed H-frame 230-kV tower 72 on Georgia Power CedartownPortland 230-kV line

The Zed-Meter instruments test waves will also bounce back and forth several times before oscillations decay to acceptable levels. This response is illustrated in Figure 4-17 for two conditions, one before installing the radial electrodes and a second test afterward to see the performance improvement.

4-16

Tower base potential, injected current and measured impedance Guyed H-frame 230-kV tower 73. Georgia Power Cedartown-Portland Median, 500800 ns: 8.0 untreated; 3.8 with four 38-m buried radial crowfoot electrodes in parallel

Figure 4-17 Typical Zed-Meter instrument results for H-frame tower with 16 guy wires

Figure 4-17 illustrates that the guyed structure has a tower base voltage that takes almost 500 ns to stabilize, and the effects of the tower resonant frequency are even seen to some extent in the measured currents. In this test series, relatively high soil resistivity led to a fairly rapid propagation velocity along the current reaction lead. This reduced the sweet spot for valid measurements from both directions, leaving a fairly narrow 300-ns window from 500800 ns. The current reaction lead was 90 m long, and it would have been helpful with this tower type to use a longer, 125-m length, which would have extended the reflection from the far end of the reaction lead to about 1100 ns. The leads in this case should be laid from the tower base, not from one of the guy wires. The leads should bisect the guy wires, and because the oscillation time is longer, a minimum length of 125 m should be used. The test results in Figure 4-17 showed that the Georgia Power grounding specification, calling for installation of four buried, 38-m radial wires, was effective in reducing the surge impedance of the tower by a factor of two. However, we could have ensured greater confidence in the results if we had used a longer, 125-m current reaction lead to extend the duration of the sweet spot. This is especially important when measuring towers with low footing impedance, because the voltage swings below zero volts give negative values to the Z(t) vector before the beginning of the sweet spot.
4-17

Considerations for Twin Steel-Pole Towers


Each leg of a twin steel-pole tower such as the one illustrated in Figure 4-18 will have slightly different impedance. The lightning response will be the parallel combination of the two impedances, adjusted for the effects of their mutual resistance. Measurement of the lightning response with the Zed-Meter instrument should eventually give the same impedance. The question about how long to wait has a relatively complicated answer. There is a smooth path from one leg, up to the cross-member, across to the adjacent pole, and back down to the adjacent concrete base. This smooth path means that the tower is susceptible to resonance at a frequency that is related to the total path length. The pole height (11 m) and separation (6 m) can be scaled from the tower photo, using the length of the 12-unit insulator string. The total path length from foundation to foundation works out to 28 m, which has a travel time at the speed of light of 94 ns. Figure 4-18 shows that the calculated impedance profile has an oscillation with a period of about 90 ns, which persists to about 700 ns.

4-18

Tower base and ground lead

Overall view

4.2 with denoising process

5.9 on other pole, longer 300-m leads

Figure 4-18 Typical Zed-Meter instrument measurement results for each leg of twin steel-pole tower, Manitoba Hydro S65R tower 67

One way to compensate for problems with persistent oscillations caused by simple or guyed towers is to extend the length of the current reaction lead. This in turn will extend the currents pulse width past the time when the tower oscillations have damped down. Although it is not feasible in every case, it was quite simple to lay out a 300-m reaction lead on the frozen snow in winter and to repeat the measurement in Figure 4-18 to obtain better results at a later time after 800 ns.

4-19

Overhead Groundwire Connections


OHGWs generally have constant surge impedance, with a good estimate for individual wires given by Z=60 ln (2h/r), where h is the OHGW height (in meters) and r is the wire radius (also in meters). Lines with a single OHGW have a surge impedance that is one-half the value of the wire in a single direction. Lines with two or more OHGWs have some mutual coupling among wires that tends to increase the value slightly from the rough estimate of the surge impedance of a single wire, divided by the number of OHGWs heading away from the tower top. If the measured footing impedance is less than about 20 , the effect of four parallel OHGWs (two in each direction) can generally be ignored. For higher resistance values, a correction can be calculated from the overall impedance of the OHGW, given by the OHGW diameter, height above ground, and separation. These values should be recorded at the tower under test, for use as described in the following subsection, Comparison Tests with and Without Overhead Groundwires. Some utilities insulate their OHGWs from towers using one or more disc insulators. This is done to improve AM re-radiation characteristics, to limit loop flow current, or for other technical reasons. In these cases, the transmission tower itself appears as a capacitance in parallel with the footing resistance. After the tower is charged up to the test potential, it does not exert any more influence. This process takes a time roughly equal to four times the tower height, divided by the speed of light. In this case, the influence on the Zed-Meter instruments measurement occurs early in time, and the case should be treated in the same way as the oscillations of a guyed tower.
Comparison Tests with and Without Overhead Groundwires

The twin H-frame towers (Manitoba Hydro S65R 16 and 17) were tested with and without OHGWs attached to the pole bonds.
Table 4-5 Zed-Meter instrument test results, twin H-Frame with and without overhead groundwires Tower 17 With OHGW No OHGW Tower 16 With OHGW No OHGW

Impedance (), offset correction Impedance (), denoising process

4.98 5.10

4.81 4.19

4.20 4.43

2.58 2.67

2.17 2.65

The Zed-Meter instrument measures the parallel combination of the footing impedance and the OHGW surge impedance in two directions. With the typical 1- standard deviation in each result, there is no significant difference in the readings with or without OHGWs. In this case, with low footing impedance, the high parallel impedance of the twin OHGWs of about 150 would make a negligible change in the expected result of the parallel combination of the two values.

4-20

Tests with and without a connection were also conducted on a lattice tower with a single OHGW. The isolation was carried out live-line, as shown in Figure 4-19. A specialized highreach bucket truck was used to work between the phases of this double-circuit line in high wind conditions.

Isolation with bucket and boom between live phases

High wind conditions and blowing snow

Figure 4-19 Isolation of overhead groundwires and Zed-Meter instrument testing on Tower RY 6-7 #11

This setup would not have been feasible on a line with a traditional double OHGW because the upper phase would block access to the outboard OHGW from the inside, and the center phase would block access from the outside using a vertical lift. The test results with and without OHGW connection are given in Table 4-6.
Table 4-6 Results of Zed-Meter instrument tests on lattice towers with and without single overhead groundwire connection Impedance (), 400910 ns Trial 1 Trial 2 Trial 3 Trial 4 Median (Std. Dev)

RY 6-7 #11 without OHGW RY 6-8 #11 with OHGW RY 6-7 #10 with OHGW RS 51 #10 with OHGW

1.33 1.17 1.70 6.51

1.54 0.44 1.61 7.16 1.06 0.62 8.22 1.35 0.89 6.55

1.4 (0.1) 1.1 (0.4) 1.2 (0.5) 6.8 (0.8)

4-21

There was no significant difference between the test results on RY 6-7, tower 11, with and without the OHGWs. Other test results were obtained on steel pole towers as shown in Table 4-7.
Table 4-7 Results of Zed-Meter instrument tests on steel poles with and without single overhead groundwire connections Impedance () 400910 ns Trial 1 Trial 2 Trial 3 Trial 4 Trial 5 Median (Std. Dev)

RS 51 #15 with OHGW RS 51 #15 without OHGW RS 51 # 16 with OHGW RS 51 #16 without OHGW

1.56 1.70 6.47 7.02

1.35 1.84 6.22 7.13

1.98 1.97 6.56 7.24

1.87 2.06

1.89

1.9 (0.3) 1.9 (0.2) 6.5 (0.2) 7.1 (0.1)

The presence or absence of the OHGW had a negligible effect on the test result for tower 15 because its footing impedance was low relative to the surge impedance of the OHGW in parallel. The median result without OHGW for tower 16 was 8% higher, and a 3% increase would be expected from theory.
Calculation of Overhead Groundwire Impedance

If desired, the effect of the parallel impedance of OHGWs can be corrected in the Zed-Meter instruments result. Equation E-2 gives the surge impedance of a single OHGW in one direction. To obtain ZGW, this impedance is divided by 2 at the tower top because effectively there is a wire in each direction. With Zaa equal to 543 for a 7-mm radius wire at a 30-m height, ZGW will be 272 . With two OHGWs in each direction, the single-wire impedance of Zaa=60 ln (2h/r) is divided by 4 and then corrected for the mutual impedance, Zab. This makes use of Equations E-3 and E-4 along with the knowledge that the currents and voltages in each OHGW will be equal. For a 5-m separation between OHGWs, the mutual impedance Zab is 149 , and the total surge impedance of 0.25(Zaa+Zab) results in ZGW of 173 .
Unparalleling the Effect of Overhead Groundwires

After the value of the surge impedance of all OHGWs in parallel, ZGW, has been obtained, the true impedance of the footing is calculated from Equation 4-1:

Z footing =

1 1 Z Meas 1 Z GW

Equation 4-1

If the measured impedance is greater than the calculated impedance of OHGW, it implies that there is no conductive path from tower to the OHGW, and it also indicates that the tower is quite poorly grounded.

4-22

Which Is the Correct Value to Use?

The Zed-Meter instrument reading accurately reflects the response of the footing in parallel with the OHGWs. This is the same value that is used to calculate the tower potential rise in lightning calculations. For decisions about which towers would benefit from treatment, the Zed-Meter instruments value with OHGW in parallel should be used. The estimate of critical current for each tower should be based on the insulation level, the Zed-Meter instruments result, and the coupling coefficient from OHGW to phase conductors. Most sophisticated programs for lightning analysis consider the effects of the OHGWs in parallel. If the Zed-Meter instruments result will be entered in these programs, tower by tower, the recorded dimensions of OHGW height, separation, and radius should be used to unparallel the result.

Tests on Towers with High Noise Level


At one tower, Manitoba Hydro S65R 63, all the test leads near the steel-pole towers were found to have high noise levels of up to 80 Vrms. The high noise levels led to considerable remaining dc offset in the averaged waveforms. The induced potentials depended on lead orientation. Table 4-8 lists the test results.
Table 4-8 Typical variations in Zed-Meter instrument test results, S65R towers Tower 63 63, other pole 68 69 70

Offset (V) Impedance () with offset correction Impedance () with denoising process

+62 4.47 4.89

+7 6.01 6.78

-5 6.29 6.82

+8 8.13 8.66

-6 5.52 5.97

-26 2.07 2.66

+24 1.26 1.76

+14 3.15 3.49

+30 2.95 3.36

Tower 69 was re-tested with 300-m leads of solid, insulated, 14-gauge wire rather than 125-m RG-58 coaxial cable sheath (see Figure 4-20). The improvement in standard deviation was significant with the denoising process and the longer leads, as noted for tower 69.

4-23

1.3- impedance with offset correction

1.8- impedance with denoising process

Figure 4-20 Zed-Meter instrument measurement results for twin steel pole, S65R tower 69, with 300-m leads

Tests on Towers with Buried Counterpoise


When the soil resistivity is high, a continuous buried or surface wire is sometimes used to connect the bases of transmission towers. This wire can have a finite length and be terminated in a ground rod, or it can continue all the way to the next tower. It provides a path with low impedance at power system frequency, but the connection has high inductance and surge impedance to transient lightning currents and Zed-Meter instrument test signals. One cross-calibration series on towers with continuous counterpoise connection to a substation was carried out at Duke Energy. The Zed-Meter instruments impedance values were, on average, 2.5 times as high as the reference values of low-frequency resistance obtained using an EPRI Smart Ground Multimeter tester.

Test Lead Orientation for Lines with Counterpoise


Calculations showed little influence of parallel conductors above perfectly conducting ground on the surge impedance of the reaction leads. The presence of the counterpoise does not affect the injection of test current into the tower and grounding system. After the current is launched into the counterpoise, it propagates at a speed of about 0.3 c, compared to a speed closer to 0.6 c in the Zed-Meter instruments leads on the earths surface. The problem is that the counterpoise takes on the same tower potential that we are trying to measure against the remote potential lead. Any coupling of potential from the counterpoise system to the remote potential lead reduces the measured voltage, giving an incorrect (low) estimate of the actual impedance. For this reason, if it is at all feasible, the Zed-Meter instruments test leads should be oriented at right angles to the counterpoise to reduce coupling.

4-24

Which Leg Has the Counterpoise Connection?

In some grounding test methods that measure the current splits into each tower leg, the readings for one leg are quite low relative to the other three. In some cases, the readings of the three ungrounded legs can even be negative. When taking the parallel impedance of all values, this should still lead to a positive result. A feature of this method is that, if one leg has a low resistance, it is a good indication that it still has an intact connection to buried counterpoise. Simulations were carried out to evaluate whether the Zed-Meter instrument can also conduct this function. Results for 1000-m soil are most relevant because this is the type of soil in which buried counterpoise would normally be effective. Two burial depths, 0.6 m and 1 m, were simulated with 38-m length in each direction from a single tower leg, running along the line right of way. With the recommended lead orientation at 90 to the direction of the counterpoise, the results in Figure 4-21 show that the two results do deviate, especially after the 450650 ns sweet spot used to read out the impedance values. If the Zed-Meter instruments test leads must be laid along the right of way, where they are more closely coupled to the counterpoise, there is a much more notable effect. In this case, if the test leads are run along the right of way, the leg that is known or suspected to have a counterpoise connection should be the one that is tested.

4-25

Leg does have counterpoise


100 Measured Impedance: Vmeas / Imeas [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0.5 1 1.5 Time [microseconds] 2 2.5 CA1-06-38-B-L01-GR-GW-1000: 24 CA1-10-38-B-L01-GR-GW-1000: 24

Leg does not have counterpoise


100 Measured Impedance: Vmeas / I meas [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0.5 1 1.5 Time [microseconds] 2 2.5 CA2-06-38-B-L01-GR-GW-1000: 21 CA2-10-38-B-L01-GR-GW-1000: 20

Preferred leads for counterpoise test, 180 apart, 90 to right of way


100 100 Measured Impedance: Vmeas / I meas [Ohm] 90 80 70 60 50 40 30 20 10 0 0 0 0.5 1 1.5 Time [microseconds] 2 2.5 0 0.5 1 1.5 Time [microseconds] 2 2.5 Measured Impedance: Vmeas / I meas [Ohm] CA1-06-38-C-L01-GR-GW-1000: 27 CA1-10-38-C-L01-GR-GW-1000: 25 90 80 70 60 50 40 30 20 10 CA2-06-38-C-L01-GR-GW-1000: 11 CA2-10-38-C-L01-GR-GW-1000: 11

Reference leads, 180 apart, running along right of way Figure 4-21 Effect of counterpoise connection on tested leg using the Zed-Meter instrument

4-26

5
RUNNING THE ZED-METER SYSTEM SOFTWARE
Previous sections have described the Zed-Meter instrument and presented results from several versions of the instrument. The contents of those sections will remain essentially fixed, because the principles of applying a safe transient pulse to the tower base with a pair of long leads placed on the ground are what define the concept of the Zed-Meter instrument. Software, as opposed to hardware, is subject to continuous improvement. This section, which can change with each improved version, describes the operation of the existing Zed-Meter instrument using a laptop computer and the Labview program.

Overview
The program is launched by double-clicking the icon on the computer screen, or it can run automatically, depending on the installation. On startup, the instrument checks that its componentsa pair of digitizers and a special controllerare available for interrogation on the USB port. An interface screen is then presented. This allows the user to enter the specific data for the tower to be tested. Test results can all be given a common line name, in order to be grouped later into a line report. After the data are entered, the user should press the Start button. The instrument will apply its test impulses, transfer the results to the laptop, and perform some analysis. It will select an appropriate sweet spot in the records, record the median impedance and standard deviation, and read these out to the user in a large display. The user has the capability to save the data, repeat the test, or exit. Figure 5-1 shows the overall block diagram from the Labview program, and Figure 5-2 shows the details of the acquisition control loop that actually applies the sequence of pulses.

5-1

Figure 5-1 Block diagram of the main loop in the Labview program

Figure 5-2 Block diagram of the data acquisition control loop in the Labview program

5-2

Main User Interface


The main user interface can be separated into the following three sections (see Figure 5-3):
Section 1. This section contains the display selection control, which allows the user to navigate between the meter and graph displays (see Section 2). Section 2. This section shows the initialization screen and then toggles between the meter and graph displays. Section 3. This section contains the menu for the measurement and data storage screens.

Figure 5-3 Three sections of the user interface

5-3

Section 2 Elements

Section 2 is composed of three main elements:


The initialization screen The meter display The graph display

Initialization Screen

When the software is launched, the first active display is the initialization screen in Section 2. The initialization screen contains a progress bar, which indicates the status of the software and hardware initialization.

Figure 5-4 Initialization screen in section 2

5-4

Meter Display

After initialization, the second display is the meter display (see Figure 5-5). The meter display is designed to display the test parameters entered by the user (1); an impedance graph (2); the calculated impedance (3); which also includes the total measurements taken and the impedance standard deviation; and two limit indicators (4). The top limit indicator warns the user if the measured impedance exceeds the target impedance. The bottom limit indicator warns the user if the impedance standard deviation has exceeded a set limit.

Figure 5-5 Meter display in section 2

5-5

Graph Display

The graph display (see Figure 5-6) allows the user to view the voltage, current, and impedance measurements (1). The graph selector (2) allows the user to select specific measurements to view.

Figure 5-6 Graph display in section 2

5-6

Section 3 Elements

Section 3 contains the menu for the measurement and data storage processes (see Figure 5-7), which consists of the following options:
Start. Starts the measurement process (see How-To: Start the Measurement Process in this section.) Stop. Stops the measurement process (see How-To: Stop the Measurement Process in this section.) Save Data. After a successful measurement, this control can be used to save the data (see How-To: Save the Measurement Data in this section.). Line Report. Generates individual reports based on the line names stored in the log file (see How-To: Create Line Reports in this section). Data Viewer. Allows saved data files to be viewed graphically (see Data Viewer in this section). Setup. Allows the user to set parameters for the Zed-Meter instrument (see Configuration (Setup) Window in this section). Help. Displays the help file. About. Displays the software information and version. Exit. Closes the software.

Figure 5-7 Menu in section 3

5-7

Data Viewer

Press the Data Viewer button on the menu to access this function. The data viewer (see Figure 5-8) allows the user to view saved data files. The user interface for the data viewer is similar to that of the graph display. A graph exists to display the saved data, and the user can also select a specific type of data to view by using the graph selector. The data file path indicator, located to the right of the graph selector, points to the directory of the saved data files. Left-click the rectangular graph display to change plot characteristics.

Figure 5-8 Data viewer

The user can also manipulate the data that is displayed on the graph. The graph palette (in the lower left corner) allows the user to move or zoom in and out of the graph. The graph palette provides the following tools:
The cursor movement tool (left button) moves the cursor on the display. The zoom tool (middle button) zooms in and out of the display. The planning tool (right button) moves the plot around on the display. The plot legend (upper right corner) allows the user to edit the plot characteristics, such as plot colors.

To load a saved data file, follow these instructions: 1. 2. 3. 4. Press the Data Viewer button. Press the Open button on the Data Viewer interface. Select the appropriate data file and click the OK button to view the file. Use the graph selector to view specific data.

5-8

Configuration (Setup) Window

Press the Setup button on the menu to access this function. The configuration window allows the user to set critical parameters for the Zed-Meter instrument. The configuration window consists of a category selection box, which allows the user to select a specific parameter to edit, and the following three buttons:
The OK button allows the user to save changes made in the configuration window. The Default button sets all parameters to the default state. The Cancel button exits the configuration window without saving any changes. The following parameters for the Zed-Meter instrument are controlled by the configuration options: Measurement parameters File path parameters Impedance standard deviation limit Target impedance Measurement parameters. The measurement parameters (see Figure 5-9) have three inputs. The first input, Number of Pulses, sets the total number of pulses (current and voltage) to be measured by the Zed-Meter instrument, which are then averaged. Setting a higher pulse count helps to remove unwanted noise. The Impedance Measurement Start (ns) and Impedance Measurement Stop (ns) inputs set the time interval on which the median impedance and impedance standard deviation are calculated.

Figure 5-9 Measurement parameters

5-9

File path parameters. The Data File Path field (see Figure 5-10) allows the user to specify a folder to which all data files, including line reports, are saved when the Save Data button is pressed.

Figure 5-10 File path parameters

To set the data file path, follow these instructions: 1. Click the folder icon to the right of the Data File Path field (see Figure 5-10). 2. Navigate to the preferred folder (see Figure 5-11). 3. Click the Current Folder button (see Figure 5-11) to accept the folder as the save path.

5-10

Figure 5-11 Pop-up screen for setting data file path

Impedance Std Deviation Limit. This screen (see Figure 5-12) allows the user to set a limit for the impedance standard deviation. If the calculated impedance standard deviation is higher than the limit specified by the user, it can indicate a connection problem between the Zed-Meter instrument and the structure under test.

Figure 5-12 Impedance standard deviation limit

5-11

Target impedance map. This screen (see Figure 5-13) allows the user to set a target impedance for a specific line voltage. The target impedances specified in the map are then used by the ZedMeter instrument as impedance limits. To save changes to the map, press the Save button; pressing the OK button has no effect on the target impedance map.

Figure 5-13 Target impedance map

Setting the Time Windows

The lead configuration should be set up so that the minimum sweet spot duration is 200 ns. The default time window of 500700 ns is practical in most cases because it allows for the decay of oscillations in guyed towers, yet ends well before reflections are expected back from the end of the current reaction lead. The time window start and end points can be adjusted in the Measurement Parameters window of the configuration menu (see Figure 5-9) after viewing the test waveforms.
Setting the Alarm Thresholds

Many utilities have internal standards for the low-frequency ground resistance to be achieved with an isolated tower. This value can be used as a guide to the high-frequency resistance limit. However, the lightning impulse response of the majority of tested towers has been lower than the low-frequency value. This means that a tower that achieves a 10- impedance in a valid ZedMeter instrument test could very well have a low-frequency resistance of 20 . Alarm thresholds can be set in the Target Impedance Map screen (see Figure 5-13)
Pretrigger Interval

The pretrigger interval is currently fixed at 1500 ns. This allows a few cycles of AM radio interference to be recorded as input to the denoising process that might be necessary.

5-12

Demo Mode
If any of the three hardware devices cannot be detected, the software enters Demo Mode (see Figure 5-14). In this mode, the user does not have the ability to acquire and save measurement data. In order to resume normal operation, the user must check all hardware connections to the computer and restart the software.

Figure 5-14 Demo mode

5-13

How-To: Start the Measurement Process


To start taking measurements, ensure that the hardware is properly connected and the software is not running in demo mode. 1. Press the Start button in the menu (see Figure 5-7). The Test Parameters window displays (see Figure 5-15).

Figure 5-15 Test parameters window

2. Enter the necessary information in the Test Parameters window (see Figure 5-15): Operator name. Select the name of the current user. Date and time. The date and time are entered automatically, based on the computer system time. Line name. Enter or select the line name, a location-specific parameter based on the line. Structure type. Enter or select the structure type, a location-specific parameter based on the structure. Structure number. Enter or select the structure number, a location-specific parameter based on the structure. Line voltage and target impedance. Select a proper line voltage for the line being measured. The corresponding target impedance is automatically selected based on the line voltage. This information is gathered from the target impedance map, which is accessed through the configuration (setup) window.
5-14

Comments. Enter a brief description of the lead layoutfor example 90 degree, voltage on ROW or 180 degree, cross to ROW for counterpoise. 3. Click the OK button to start the measurement. 4. If the test parameter line name or structure type is different from the previous input, the software performs an auto range (see Figure 5-16). This is where the proper hardware settings for the digital storage oscilloscope are set.

Figure 5-16 Zed-Meter instrument auto range

5. Monitor the Pulse Count indicator on the meter display (see Figure 5-5). 6. After the Pulse Count indicator reaches the number of pulses set in the Measurement Parameters screen of the configuration window (see Figure 5-9), the software calculates and displays the line impedance and the standard deviation of the impedance. If the line impedance or the standard deviation of the impedance is too high, the limit indicators light up. To view a graphical display of the measured data, select the Graph option in the display selector (see Figure 5-3).

How-To: Stop the Measurement Process


To stop a measurement that is currently in progress, press the Stop button in the menu (see Figure 5-7). The Stop button is active only when a measurement is taking place. Otherwise, this button is disabled.

How-To: Save the Measurement Data


Data can be saved only after a successful measurement. Otherwise, the Save Data button is disabled. The data are saved in a text (.txt) file. The file name is composed of the line name, followed by an underscore, and then the tower number. To save the current measurements, press the Save Data button. The software creates a data folder and a log folder inside the folder (data file path) that you specified in the configuration window (see Figure 5-10).
Data File

The data folder contains the text files, in tab-delimited format, of data for each of the lines measured. These files can also be opened in Excel for further analysis. The data file (see Figure 5-17) contains several lines showing the identification data, followed by the sweet spot start and stop times. The measured value of impedance and its standard deviation follow on the next lines. Next is a heading with the time (ns), remote voltage, lead current, structure current, lead impedance, and structure impedance, followed by 8000 readings starting at -1500 ns.

5-15

Figure 5-17 Example of data file for test of National Grid line 4VW_234 structure

Log File

When measurements on a particular line are finished, pressing the Line Report button (see Figure 5-7) causes the software to create a second text file (see Figure 5-18), the log file. The log folder contains the log file, in tab-delimited format, for all the lines measured. The difference between the log file and the data file is that the information collected in the log file is reserved for creating line reports. The log file can also be opened in Excel for further analysis.

Figure 5-18 Example of a log file

How-To: Create Line Reports


Line reports can be created only if a log file is present. 1. Press the Line Report button (see Figure 5-7). 2. If a log file is present, the software creates a line reports folder and the necessary line report files. The line reports folder is located inside the folder (data file path) that you specified in the configuration window (see Figure 5-10). 3. When the Line reports created prompt (see Figure 5-19) displays, click OK.

5-16

Figure 5-19 Line report created prompt

4. A line report (see Figure 5-20) is generated for each line that was measured. These files are quite similar to the log file, except they are for each individual line. The line reports can also be opened in Excel for further analysis. 5. After the waveforms have been reviewed, you can change some of the default settings for analyzing the results.

Figure 5-20 Sample line report

5-17

6
TYPICAL ZED-METER INSTRUMENT RESULTS
The s on the end of the section title, Results, is deliberate. The impedance of a ground electrode is not a constant value. The Zed-Meter instruments impedance profile can decline with increasing time, suggesting that the grounding system has long buried wires (counterpoise). In the more common case, Z(t) can increase slightly with time as a result of the changes in resistivity of common earth materials and also because the ground has capacitance that takes some time to charge up. The time window selected for analyzing the impedance profile will affect the result in each case. Experience has shown that a good sweet spot of 500 to 800 ns can be achieved in most tests, and this is used as a reference in the production Zed-Meter instrument. However, additional information can gained by evaluating the impedance at different time intervals in postprocessing. For this reason, at least one waveform file should be saved for every test.

Effects of Overhead Groundwire


The test result should increase by a small amount when the OHGWs are disconnected at the tower top. The high (170250 ) surge impedance of all the OHGWs appears in parallel with the low (530 ) impedance of the footing. The amount of the increase in the test result depends on the value of footing impedance. For well-grounded towers, the increase should usually be less than 10% when OHGWs are isolated. For this reason, and also to save test time, isolation of the OHGW is not recommended as part of the normal test process. The OHGW heights and separation should be recorded if a correction for the parallel impedance will be made after the measurement. On towers with insulated OHGWs, there is often a strong benefit to carrying out conventional low-frequency measurements of resistance, using an oblique (4560) orientation of the potential profile in the three-point test. This lead geometry gives the low-frequency tower resistance, R, the local average resistivity, , and effective ground electrode perimeter, /R, with the same effort usually given to obtaining only the resistance value. Tests of this sort formed an important component of the Zed-Meter instruments cross-calibration program at several utilities.

Reasons for Steadily Increasing Impedance Values


In many of the test results, there is a slight trend for the impedance profile with a wide sweet spot to increase with time. The median impedance in the time range from 500 ns to 700 ns forms the reference. Nearly all tower and lead configurations produce good results in this window. It takes some extra time to lay out the extra lead length needed to obtain an impedance profile that extends to more than 1000 ns, especially on soil with high resistivity. However, when this is done, processing of the results in two separate time windows will often yield different impedance values.

6-1

Low-Frequency Versus High-Frequency Resistivity

The first physical reason for an increase in impedance with increasing time is related to the role of the dielectric response of the soil. Many natural soil materials have relative dielectric permittivity values, r, in the range of 5 to 10. Especially for soils with high resistivity (>1000 m), the initial impedance of the soil will be capacitive, and it will charge up with a time constant to the eventual resistive value observed at lower frequency. Figure 6-1 gives one indication of how much difference there should be when Zed-Meter instrument values are compared to low-frequency measurements, based on the difference in measured soil resistivity at 100 Hz and 100 kHz.

Figure 6-1 Expected ratio of Zed-Meter instrument readings to low-frequency readings of compact electrodes

If the soil around the tower is sand, the Zed-Meter instruments result should be 7085% of the low-frequency value. The actual readings will show a large change with moisture content, meaning that readings taken days or hours apart can be quite different if there has been recent rain. The factors relating Zed-Meter test results to low-frequency values are much larger for other types of soil. Till is soil smeared into place by glaciers, containing a wide range of particle sizes from boulders (>256 mm) and pebbles through sand to silt (0.06250.004 mm). Till retains moisture much better than sand and supports hardwood forest or farming. For towers in this common soil type, Zed-Meter instrument readings should be 5070% of the values found in lowfrequency measurements at 100 Hz.

6-2

Clay has the smallest particle size (<0.0020.004 mm), and its resistivity has the strongest frequency dependence. Zed-Meter instrument readings for towers in clay should be 1033% of the values obtained with low-frequency tests. Because clay resistivity is very low compared to sand or till, both readings will often be <10 . In practice, most utilities take some care to avoid placing towers in clay soils to ensure mechanical stability of the foundations. Overall, cross-checks on isolated towers, plotted together in Figure 6-2, confirm that the ZedMeter instruments impedance results tend to be well below the low-frequency measurements.

Figure 6-2 Comparison of the Zed-Meter instruments impedance with independent measurement at low frequency

The graph of results from a total of six weeks of cross-calibration tests in Figure 6-2 confirms the expectations raised from fixed-frequency tests of soil samples in Figure 6-1. The majority of the test points lie above the y=x line. For compact electrodes, such as tower foundations without radial counterpoise, the Zed-Meter instruments impedance is less than that found at low frequency.
Effect of the Time Window, 500 ns Versus 1000 ns

Some of the test results in the cross-calibration program had Z(t) records that allowed analysis at two different periods in a wide sweet spot, extending all the way from 500 ns to 1000 ns. Figure 6-3 shows that, for towers with impedance <3 , there was little change in Z(t) with time. Above this level, the impedance at 500 ns increased by a factor of two when the Z(t) profile was evaluated at 1000 ns.

6-3

Figure 6-3 Comparison of Zed-Meter instrument results at 500 ns and 1000 ns for compact electrodes with no extensive buried wire

The Z(t) results for 1000 ns are much closer numerically to the low-frequency measurements. The power-law fit has an exponent that is nearly linear, and the regression coefficient is better than the relation for the 500-ns values. The change in impedance from 500 ns to 1000 ns is also interesting because models based on material tests predict that this change should occur much more slowly. Two points are circled in Figure 6-3. For these two towers, reference tests showed that one of the tower legs had an intact counterpoise connection to a nearby substation. For these distributed grounding systems, the Zed-Meter instruments impedance was considerably higher than the low-frequency value at both 500 ns and 1000 ns. Valid Zed-Meter instrument results at both 500 ns and 1000 ns can usually be obtained by laying out current reaction leads of suitable length. This is now recognized as a desirable goal in the test procedure.
Special Case: Towers with Isolated Overhead Groundwires and High Resistivity

At Tennessee Valley Authority, the 500-kV transmission system is operated with insulated OHGWs to reduce power loss. This simplifies the measurement of tower resistance using lowfrequency methods, and in fact, Tennessee Valley Authority maintains an extensive database of these results. During cross-calibration, some of the towers with the highest values of resistance were re-measured with standard and Zed-Meter instrument methods. When there is no OHGW connection, the parallel impedance in the order of 200 will not appear as an upper limit to the measurement. The tower itself will appear as a capacitance, calculated from the tower surge impedance and travel time, in parallel with the footing impedance (see Figure 6-4).

6-4

Figure 6-4 Schematic Zed-Meter instrument circuit for transmission lattice tower with insulated overhead groundwires

Figure 6-5 shows the Zed-Meter instruments Z(t) profiles for the two towers with the highest impedances in the test program. For these isolated and poorly grounded towers, both show increasing impedance with time in the range of 500 to 1000 ns.

Tower 307: low-frequency resistance 163

Tower 309: low-frequency resistance 208

Figure 6-5 Impedance profiles for isolated 500-kV towers in area of high soil resistivity

The sweet spot for these measurements starts rather early, at about 200 ns. The R-C time constant of the tower capacitance (1 nF) and the Zed-Meter instruments impedance is about 100 ns, and this effect is seen early in the wave. However, the rise in impedance from 500 ns to 1000 ns is a characteristic of the soil response, not the tower. The sweet spot of the Z(t) profile for tower 309 also ends rather early, at 500 ns. The reason that the standard deviation of the measurement in tower 307 stays low after the reflection from the current reaction lead is that the termination resistance in this case provided a reasonable match to the current reaction lead surge impedance.

6-5

Reasons for Steadily Decreasing Impedance Values


In the bench tests for the Zed-Meter instrument, it was noted that the meter responds to a short circuit cable across its terminals with a reading that is initially high and then falls to a low value. It was possible to estimate the lead inductance (typically L= 1 or 2 H) from the time constant, t=L/Z. The Zed-Meter instrument uses this capability to measure the effect of the inductive response of ground electrodes, which will typically be 1 H per meter of length, in addition to the resistive response. This is reported as an impedance profile, Z(t), that starts with a high reading and falls with time.
Typical Responses of Buried Horizontal Wires

It is practical and useful to measure the impedance of newly installed buried wires with the Zed-Meter instrument. The wiring connection that is normally used for the dipole test is adapted in Figure 6-6 for this purpose.

Distributed
Figure 6-6 Wiring layout for Zed-Meter instrument measurement of distributed electrode impedance

A test of this sort was carried out on a set of four 38-m buried wires, radiating outward from the tower base. The impedance profile is shown in Figure 6-7.

6-6

Figure 6-7 Typical Zed-Meter instrument measurement of 38-m buried wire with 26- low-frequency resistance

The initial impedance of about 16020 is typical of the surge impedance of counterpoise systems. In this case, the impedance of the four wires in parallel falls relatively quickly, reaching a value of 47 at 500 ns and 29 at 1000 ns. The low-frequency resistance between this fourradial-crowfoot electrode and the tower was measured to be 26 with an earth resistance tester. The impedance result can be modeled as an L/R time constant of 300 ns to give an effective inductance of 8 H. This is a good match to a rough estimate given by the inductance of each 32-m wire (32 H) divided by the number of wires (4) in parallel. In 10 of 15 cases in which Zed-Meter instrument tests were carried out on towers that were known or proven to have a distributed electrode, the impedance at 500 ns was greater than the impedance at 800 or 1000 ns. Figure 6-8 also shows that the Zed-Meter instrument result at 500 ns was greater than the low-frequency value in 12 of the 15 cases, by as much as a factor of four.

6-7

Figure 6-8 Comparison of Zed-Meter instrument results with low-frequency resistance measurements for distributed electrodes with long buried wires

As a reminder, results for compact electrodes in Figure 6-2 were consistently above the y=x line, corresponding to Zed-Meter instrument impedances that were less than the low-frequency values, and the results for a later time (see Figure 6-3) showed an increasing impedance with time. Thus, a declining impedance profile in a Zed-Meter instrument result is a characteristic of a ground electrode with considerable inductance in series with the low-frequency resistance. Again, it is stressed that there is some advantage to taking the extra time to lay out sufficient reaction and remote potential cable lengths, oriented at right angles to the run of the counterpoise wires, to obtain a wide sweet spot of 5001000 ns. It may be tempting to use one or both of the counterpoise leads as alternatives to the coaxial cables laid above the ground. Some advice: Do not try this unless you are willing to accept the slow convergence of your Sommerfeld integral terms when you do your post-processing analysis.
Reading Past the First Sweet Spot

The two impedance profiles in Figure 6-5 showed that it can be feasible to obtain impedance readings after the return of the reflection from the terminated end of the current reaction lead. In these measurements, the question about what happens at a later time could not be answered because the test pulse width was 1200 ns. Since these tests were performed, the Zed-Meter instruments hardware has been modified to provide a much longer pulse width.

6-8

Although the initial Zed-Meter instrument reading is not influenced by the termination (open or grounded) of the current reaction lead, this is not true for the readings taken at later times. The potential at the tower base must be adjusted for the influence of the test pulser voltage minus the tower base potential rise at a later time. The steps to this process for an isolated tower include the following: 1. Estimation or measurement of the potential applied to the remote ground rod. For a high resistance relative to the tower, this will be approximately 200V. 2. Estimation of the effective radius of the temporary grounding probe for the current reaction lead. For example, a driven rod with a 0.005 m radius and a 0.3 m length is equivalent to a hemisphere of radius a= 0.055 m in uniform soil. 3. Calculation of the influence voltage, in this case 200 V (a/d), where d is the distance from tower base to the temporary ground probe. This is about 90 mV when d=125 m. Several factors can lead to a poor-quality low-frequency (late-time) measurement with the ZedMeter instrument. If the tower current falls much below 100 mA, the typical tower base potential rise will be on the order of 2 V. It can require extra averaging to obtain a suitable waveform if the noise on the test leads is considerable. It can be desirable to reduce the probe resistance by wetting the soil in order to increase test current if this technique is to be exploited. However, this changes the shape of the electrode and spoils the estimation of resistivity. Finally, the soil resistivity estimate is quite sensitive to changes in the humidity of the upper soil layer. All of these factors argue for using the correct toola three-point low-frequency earth resistance tester with an oblique probe geometry, rather than the Zed-Meter instrumentto obtain a late-time (low-frequency) impedance value. If the tower has connections to OHGWs, the Zed-Meter instruments impedance profile will be subject to the same issues that affect low-frequency grounding. There will be a continuous decline in the impedance as more and more adjacent towers take up their shares of test current. In Figure 6-9, the initial Zed-Meter reading of about 10 falls to about 6 at 10 s from this effect.

6-9

Figure 6-9 Extended impedance profile (to 10 s) obtained with grounded current reaction lead

It becomes increasingly complicated to compensate the tower base voltage for the influence voltages that arrive, delayed in time, from the adjacent towers.

Reasons for High Impedance Values


There are two main reasons why the Zed-Meter instruments result can be high. First, there can be a bad connection to diagnose if the right sort of current flow is not occurring in the test leads. Second, the impedance can be initially low and then rise to a value close to the impedance of the OHGWs. This is a characteristic of a poorly grounded tower situated in an area where the underlying resistivity of the rock or soil is high.
What Constitutes a High Reading?

The product of the lightning surge current with the Zed-Meter instruments reading gives a good estimate of the voltage that will appear on the transmission tower. The voltage across the insulation will be a notable fraction of this potential, because the same coupling between test leads and nearby conductors also affects the OHGWs and phase conductors. With a median lightning surge peak current of 31 kA, a 25- reading will give a tower base voltage rise of 775 kV. Factoring in typical coupling, the peak voltage from this stroke can be withstood by an insulator with 0.65-m dry arc distance, typical of a string of four discs. If instead, the line is designed to withstand a severe current of 155 kA , which will be exceeded about 1.5% of the time, the potential rise and necessary insulation length to avoid backflashover will also be five times as high, calling for 22 discs.

6-10

With typical transmission line reliability requirements that increase with voltage level, most utilities adopt limits for low-frequency resistance in the range of 25 . For a compact electrode, such as the four foundations of a transmission tower, the corresponding Zed-Meter result is in the range of 11 to 16 , depending on the time window selected for analysis. If the 25- resistance has been achieved with the use of extensive buried counterpoise, then the Zed-Meter instruments result will be higher and the lightning performance will be worse than expected.
Identifying Bad Connection to Current Leads

A current waveform with a lot of oscillation in the time range 0200 ns followed by settling to a low value of a few mA indicates that there is a bad connection to the tower or current reaction lead at the Zed-Meter instrument. With a 120-m lead length, a sudden drop in the current after 10001400 ns indicates that the current reaction lead is not connected to its terminating ground rod. This might have been deliberatefor example, in frozen soiland is not a problem unless it results in excessive induced potential on the lead.
Indications of Local Soil Resistivity from Dipole Test Results

The surge impedance of the current reaction and remote potential leads in the dipole test is a function of two variablesthe height of the leads over the soil and the resistivity of the soil. If the leads are laid on the earth surface, the following are some telltale signs that the underlying resistivity is high:

A surge impedance of more than about 600 . An early arrival of the reflection from the end of the current reaction lead. For a 120-m lead, the two-way speed at the speed of light, c, is 800 ns, so an early reflection arrival (0.8 c) for this lead length would be at 1000 ns. A reflection from the end of a grounded current reaction lead that drops the current almost to zero, behaving in nearly the same way as when the lead is ungrounded. If some or all of these factors are noted in the dipole test, it is reasonable to expect a high impedance reading from the Zed-Meter instrument when it is connected to the tower and to reset the automatic scaling to adjust to this possibility.

Reasons for Low or Negative Impedance Values


Some measurements with the Zed-Meter instruments have given results that were less than 1 . As it turns out, most of these were obtained on isolated towers in areas of clay soil that had typical low-frequency resistivity of 5080 m. These towers push the limits of what can be achieved with the Zed-Meter instruments test method. For example, having a 1-m lead from the meter to the tower adds an inductance of about 1 H. The L/Z time constant of this simple circuit works out to a negligible 100 ns with a 10- impedance, but it becomes a notable 1000 ns for a tower that truly has a 1- footing impedance.

6-11

With the low resistance also comes poor damping of oscillations. Many of the Z(t) profiles for well-grounded towers swing above and below zero, sometimes all the way through the desired sweet spot for analysis. The Zed-Meter instrument is programmed to report a zero value when the median is negative, but the trainer should be alert to understand why this has happened.
Identifying a Bad Connection to the Potential Lead

If the instrument indicates an impedance of less than 1 , there are a few issues to double check. First, the connection of the potential lead to the instrument might be loose. It might be prudent to repeat a measurement with the remote potential lead deliberately disconnected or with the reaction leads reversed.
Inserting Resistance in Series with Tower to Validate Connections

When low impedance with a large relative standard deviation is reported, a double check is warranted. One method that ensures confidence in the reading is to insert a calibration resistance such as 10 or 50 in series into the tower lead, using a pair of BNCtobanana plug adapters or a dedicated fixture. A repeat of the measurement should lead to a value that is 10 or 50 higher than the initial reading. If the measurement of the tower impedance with 10- or 50- resistance in series indicates the nominal calibration value, the tower is well grounded. Some care should be taken to ensure that the comment field points out the validation, lest the resistance value be taken as correct in the log and summary files.
Indications of Local Soil Resistivity from Dipole Test Results

The indicators of high soil resistivity in the dipole test are reversed to indicate the possibility of low local soil resistivity, which can help to confirm a low impedance. Low soil impedance is suggested by a dipole test result that has the following indications:

A surge impedance of less than about 450 . A late arrival of the reflection from the end of the current reaction lead. For a 120-m lead, the two-way speed at the speed of light, c, is 800 ns; therefore, a late reflection arrival (0.6 c) for this lead length would be at 1330 ns. A reflection from the end of a grounded current reaction lead that leads to a notably increased current.

6-12

7
POST-PROCESSING
Organization of Data
The log files and line reports are helpful for organizing records, but post-processing is done on the waveform files. As shown in Figure 7-1, the waveform file consists of a header followed by 8000 records of time (ns), voltage rise, currents, and impedances.

Figure 7-1 Example of data file for Zed-Meter instrument test of National Grid line 4VW_234 structure

It is recommended that post-processing be done in Excel, which offers a suitably rich set of options and is widely available in the electrical utility industry. Other programs, such as Matlab software from The MathWorks, Inc., can offer a richer set of resources in the signal processing toolboxes that are interesting to PhD-level engineers, but they are also susceptible to the common problem of torturing data until it confesses, whether it is innocent or guilty.

Converting Waveform Files to Excel Format


The waveform files can be accessed with Excel (File, Open, All files). The Text Import Wizard will start automatically. The waveform files can be imported into the native Excel format by clicking the Finish button on the first screen of the Wizard. The description of the waveform files in Figure 7-1 can be used to identify the columns.

Producing Waveform Graphs in Excel


Generally, variables should be plotted against the vector of time values (ns) found in the first column (A).

7-1

Exporting Waveform Graphs from Excel to Portable Document Format


Generally, graphs exported from Excel have poor resolution and look bad when imported into technical papers and reports. One way to improve quality is to print the graph as an Adobe portable document format (.pdf) document, selecting high-resolution options. The conversion process produces a scalable document that satisfies most publication quality requirements.

Establishing Confidence Level in Test Results


For identifying the sweet spot of the waveform, a plot of the lead and structure impedances, along with the local standard deviation of 32 values, 16 from each vector, has been used in many of the plots in this report. These graphs have also used a logarithmic axis for the impedance in order to see the three variables with good resolution on the same page.

Adaptive Filtering to Improve Confidence


The analyst has good confidence in the Zed-Meter instruments result if the standard deviation is a small fraction of the result. With an inherent instrument standard deviation of about 1 , this means that measurements of towers with low impedance might need to be improved with postprocessing. Offset subtraction and noise removal procedures have proved to be useful in some situations.
Offset Subtraction

The first set of post-processing duplicated the intended functions of the Labview software in the prototype instrument. The offset for each voltage and current signal was calculated from the average of the last 2000 points in the 8000-sample (80-s) record. This value was then subtracted from every point. The sample-by-sample values of impedance were given by the measured potential, V, divided by I1 or by I2. The standard deviation of 15 samples using both Z1 and Z2 values was also computed. This processing assumes that there is a quasi-static offset that does not change much over the 80s record length. The Zed-Meter instrument configuration used in this study does not provide a pretrigger record to conduct the offset process on the values before the trigger point. Preliminary work with oscilloscopes used about 25% pretrigger with a 10,000-sample record length [2, 3, 4]. When the offset is computed from the points near the front of wave, uncertainty about the change in noise signal over the 80-s record length is reduced to uncertainty about the change over the 2-s measurement interval. Now that the Zed-Meter instrument is automatically set up to use pretrigger values, these are the preferred data that are used to compute the necessary offset corrections for each averaged signal. Tests at one tower in Manitoba showed a high noise level on the potential lead, with up to a 30-V offset in the tail of wave to be subtracted from every point. At this tower, the Zed-Meter instrument test was repeated four times, twice on one pole of the H-frame tower and twice on the other (see Table 7-1). Test 3 used 125-m coaxial lead wires, and test 4 used 300-m leads of 14-gauge wire.

7-2

Table 7-1 Typical variations in Zed-Meter instrument test results, S65R tower 67 Tower 67, Trial Number 1 2 3 4

Offset (V) Impedance (), Offset Correction Impedance (), Denoising Process

-30 3.79 4.17

-11 3.78 4.31

+6 5.17 5.43

-18 5.73 5.90

There was less than one standard deviation difference among the results with offset correction. Table 7-1 also shows the effects of the digital artifact denoising process of about 3 Vp-p.
Denoising Digital Artifact

One Zed-Meter instrument prototype had internal digital noise problems that could have compromised a test campaign. The noise problem in this case was that, in the pretrigger record and again at the tail of the wave, the voltage should be constant and should average to zero. What was found was a repetitive pattern that was quantized into discrete levels, with an 8-sample period as shown in Figure 7-2.

Figure 7-2 Tail-of-wave noise pattern in Manitoba Hydro Zed-Meter instrument trials

This pattern was fitted by adapting each value to give a minimum error on the tail of the wave, where the signal should be low. The repetitive pattern was extended along the entire vector of voltage values and subtracted. Figure 7-3 shows the degree of improvement that can be achieved with this relatively simple process.

7-3

Z(t) before denoising

Z(t) after denoising

Figure 7-3 Improvement in measurement uncertainty with 16-sample denoising, Manitoba Hydro S65R_067

The facts that the repetitive pattern was synchronized to the sample rate, that it had discrete levels corresponding to quantization limits, and that it was not eliminated by averaging suggest some combination of internal software and hardware problems that were later resolved. For usercustomized versions, the following suggestions to eliminate the noise should again be checked: Are portions, rather than the entire vector of voltage values, being averaged? Is the dynamic range of the digitizer (expressed as number of effective bits versus sample rate) deficient at high sample rates? Is there internal coupling from the potential lead to the digitizer hardware? The good experience with denoising can also be extended to cases in which, despite averaging, there is an interference signal in the pretrigger record. The same principlesidentifying and fitting the noise record in Excel and subtracting the fitted noise signal from the original record can achieve success in these cases.
Removing Noise Using Pretrigger Record

The use of a pretrigger period of 1500 ns can show currents and voltages that are nonzero, even after averaging many test impulses. One example of interference with a period of about 500 ns (frequency of 2 MHz) was recorded during tower impedance tests at the National Grid in the United Kingdom. Figure 7-4 shows that peak-to-peak currents of about 4 mA, with approximately zero mean, were found on one tower.

7-4

Line ZK, tower 1011 (ZK_1011)

Line 4VW, tower 2341 (4VW_2341)

Figure 7-4 Pretrigger records for Zed-Meter instrument impedance tests at National Grid

The complex nature of the noise in the ZK_1011 record will be difficult to fit with a simple set of sine and cosine wave functions. However, the 1100-kHz pretrigger noise in the 4VW_2341 record has a relatively simple form that could be fitted by a sine term and a linear term, using the Excel Solver add-in to minimize an error sum by adjusting the frequency, phase, and amplitude of the correction function.

Validating a Novel Test Lead Arrangement


It has proved to be helpful to conduct numerical analyses of the frequency and time domain response of Zed-Meter instruments test lead configurations. In particular, the calculations have given increased confidence that the rising impedance profiles in high-resistivity soil have a sound physical basis. Other features of the test waves, such as the effect of lead direction and orientation, have also been tested. The use of additional lead length to optimize the sweet spot of the Z(t) profile was optimized with the calculation results. Also, the need to have a short and direct connection to the tower leg was an output of numerical analysis. When field test results are returned with a note or sketch showing what was done to obtain a result, it might be necessary to determine a correction factor for the result. This would normally be done by a specialist in time-domain electromagnetics, using one of the standard modeling programs, such as a version of the Numerical Electromagnetics Code (NEC).
Validating Against What?

The standard in-line test lead configuration uses current reaction and remote potential leads that run along the right of way, parallel to the OHGWs and phase conductors. The leads are oriented at 180. Changes to the lead orientation and geometry have generally been evaluated against the benchmark result. It is reasonable to ask how well the Zed-Meter instrument actually simulates the response of the tower to real lightning. This breaks into two separate questions: what happens for a stroke to midspan, and what happens for a stroke to the top of the tower.

7-5

Modeling the lightning channel above the tower is feasible, but it calls for the adoption of an engineering model of the channel that accurately reproduces the nearby electric and magnetic fields. This is a rather specialized domain, and the results can depend to some extent on the choice of return stroke model. An interim step is to evaluate the response of the tower and OHGW to a current injected at midspan as shown in Figure 7-5.

Figure 7-5 Reference result: impulse injection into midspan

Qualitatively, after noting that there will be a time delay of about 400 ns for the pulse to reach the tower base, there is reasonable agreement between results. For a soil resistivity of 50 m, the voltages in both cases stabilize at about 2 V in the time range from about 200 ns to 700 ns after the initial peak. The result is about 18 V for the case of 1000-m soil. With 20,000-m soil, the voltage no longer reaches a constant value. Instead, it rises continuously, to a level of 100 V at 900 ns. Effectively, the tower is acting as a lumped capacitance in this case, and the ground electrode has no effect on the resulting voltage. (See Figure 7-6.)

7-6

Pulse injection at midspan

Zed-Meter instrument injection at tower base


14 Vtower-grd.ref Seg.No.110

50 m

Sim# Pulse Injection GW-50 BIS 6 Vtower-grd.ref Seg.No.1 5

12 10

4 3

V o lta g e[V ]

8
1 0 -1 -2

V o lta g e [V ]
0 0.5 1 1.5 Time [microseconds] 2 2.5

6 4 2 0 -2

0.5

1 1.5 Time [microseconds]

2.5

1000 m
25 V 20
tower-grd.ref

30 Vtower-grd.ref Seg.No.110 25
Seg.No.1

20 15 V oltage [V ]
0 0.5 1 1.5 Time [microseconds] 2 2.5

15 Voltage [V]

10

10 5 0 -5 -10 -15

-5

0.5

1 1.5 Time [microseconds]

2.5

20,000 m

150 Vtower-grd.ref Seg.No.1


120 Vtower-grd.ref Seg.No.110

100

100 80

Voltage [V]

60 Voltage [V]

50

40 20 0 -20 -40

-50

0.5

1 1.5 Time [microseconds]

2.5

-60

0.5

1 1.5 Time [microseconds]

2.5

Figure 7-6 Tower base potential to remote ground for midspan and Zed-Meter instrument current injection

7-7

NEC-2 for Simple Coupling Estimates

The NEC-2 computer program has some notable advantages for analyzing coupling. It is freely available on the Internet, and it is supported by the amateur (ham) radio community. Typically, it is used to calculate the impedance and radiation pattern of antennas at one or a narrow band of frequencies. The antenna geometry is described by a file with a list of start and end points, wire radius, and other factors. Lossy soil can be modeled, using the Sommerfeld-Norton model, but one limitation in calculations is that the antenna segments cannot penetrate into the lossy ground. For time-domain calculations, computations must be performed for large numbers of points covering the entire frequency spectrum of interest. The results must then be transformed into the time domain, using an inverse Fourier transform function in Excel, Matlab, or another program. The NEC-2 program comes with a specific case study for Beverage antennas. This case can be modified easily to match the characteristics of the Zed-Meter instruments current reaction and remote potential leads. For evaluating coupling in cases with buried wires, the NEC-4 computer program must be used as described in Appendix F. The Zed-Meter system computations in this document have been carried out in the frequency range of 195.3 kHz to 100 MHz with the increment step of 195.3 kHz. This corresponds to the time range of 05.12s with a 5-ns sampling interval.

7-8

8
REFERENCES
1. L. E. Juhlin and Members of Working Group C4.2.02 (formerly 36.02), Methods for Measuring the Earth Resistance of Transmission Towers Equipped with Earth Wires, CIGRE Brochure 275, June 2005. 2. The EPRI Zed-Meter: A New Technique to Evaluate Transmission Line Grounds. EPRI, Palo Alto, CA: 2004. 1008734 3. Field Testing of EPRI Zed-Meter: Transient Impedance of Transmission Line Grounds. EPRI, Palo Alto, CA: 2005. 1010235. 4. Summary of Zed-Meter Field Tests: Transient Impedance of Transmission Line Grounds. EPRI, Palo Alto, CA: 2006. 1012314. 5. Guide for Transmission Line Grounding: A Roadmap for Design, Testing, and Remediation. EPRI, Palo Alto, CA: 2004. 1002021 6. S. Bourg, B. Sacepe, T. Debu. Deep earth electrodes in highly resistive ground: frequency behaviour, 1995 IEEE International Symposium on Electromagnetic Compatibility. August 1418, 1995, pp. 584589. 7. N. Theethayi, R. Thottappillil, M. Paolone, C. A. Nucci, F. Rachidi. External impedance and admittance of buried horizontal wires for transient studies using transmission line analysis. IEEE Transactions on Dielectrics and Electrical Insulation, Vol. 14, No. 3, pp. 751761, June 2007. 8. J. Ma and F. P. Dawalibi. Influence of inductive coupling between leads on ground impedance measurements using the fall-of-potential method. IEEE Transactions on Power Delivery, Vol. 16, No. 4, pp. 739743, October 2001. 9. A. Semlyen. Ground return parameters of transmission lines: an asymptotic analysis for very high frequencies. IEEE Transactions on Power Apparatus and Systems, Vol. PAS100, No. 3, pp. 10311038, March 1981. 10. N. Theethayi, A. Rakov, R. Thottappillil. Responses of airport runway lighting system to direct lightning strikes: comparisons of TLM predictions with experimental data. IEEE Transactions on Electromagnetic Compatibility, Vol. 50, No. 3, Part 2, pp. 660 668, August 2008. 11. M. Darveniza. A practical extension of Rusck's formula for maximum lightning-induced voltages that accounts for ground resistivity. IEEE Transactions on Power Delivery, Vol. 22, No. 1, pp. 605-612, January 2007.

8-1

A
FREQUENTLY ASKED QUESTIONS
1. What is the basic principle of operation? A shock from an electric fence pulser is modified and applied to the tower base. The other end of the pulser is attached to a cable that runs along the ground. The cable acts like a traveling wave antenna, providing a temporary path for the return current. The tower base potential is measured between the tower and a second cable, running 90 to 180 away from the first cable. A computer and digitizers capture and store the fastchanging voltages and currents and process them into a value of impedance. 2. What instrument checks and calibration procedures do I need to do before going out to the field, and how do I do them? The Zed-Meter instrument should be operated with a short circuit on the terminals, using the tower clamp, and also with a small calibration resistor inserted in series. 3. How is the Zed-Meter instrument connected to a structure? How do I check that my connections are right? The Zed-Meter instrument should be connected to the structure using a welding clamp or antenna ground clamp that provides a low impedance at high frequency. Rust should be removed to make the connection. If the current into the tower does not equal the current into the reaction lead after some initial oscillations, lasting no more than 200 ns, or if it does not stabilize to a constant value for about 7001000 ns, then there is probably something wrong. 4. What lead lengths should I use and why? How does soil resistivity impact my choice? Leads should be at least 90 m long. Longer leads give a wider sweet spot in the measurement, which is hoped to be constant from 500700 ns after the initial pulse. Coaxial cable (RG58 or RG59) is sold in 500-ft (152-m) lengths at most large hardware stores, so this length is a good choice for most situations. The cables should always be pulled all the way off their wind-up reels. Longer 125-m or 150-m leads are needed in frozen soil, which has a high resistivity; in rocky areas, which also have high resistivity; in grassy areas, where the cable can run well above the ground surface in tall grass; and on towers with guy wires. 5. What external influences will affect the measurements and how? The presence of power-frequency, static, and radio broadcast current and voltage on the test leads is common, and the instrument will average many pulses to reduce their influence. Test results will vary (increase) slightly if the overhead groundwires (OHGWs) are disconnected.

A-1

6. How do I orient the measurement leads for various electrode and structure types? What precautions are there? For towers that do not have buried radial wires or continuous counterpoise as supplemental electrodes, the two leads should normally be oriented along the right of way, 180 apart, and stood off from any nearby metal structures by at least 1 m. For lines that do have counterpoise, the best results will be obtained if the test leads run at 90 to the line direction. This is a good configuration for all towers. Especially if the remote ends of the cables are not terminated in short ground rods, they should be treated as live until proven safe by measuring the cable-to-tower potential with a voltmeter. Test cables with the upper limit for bare-hand work (50 V) might not spark when teased against the tower steel, so this check is inconclusive. 7. How do I orient the leads when obstructions such as fences or buildings are in close proximity? The Zed-Meter instrument results are relatively insensitive to lead layout, as long as the leads are kept away from obstructions that might be well grounded. One suitable rule is the one-diameter goalfor example, keep the leads about 5 m away from a vehicle or shed or 1 m away from a tower leg foundation or fence. In cases where the choices are limited, perform a dipole test of the two leads to confirm that the impedance of the leads is constant and that the duration of the test signal is sufficiently long. 8. What measurement lead height above vegetation can be tolerated before the results are compromised? How can I tell my results have been compromised? In theory, the Zed-Meter instrument can work even if the test leads are suspended several meters off the ground. Some problems can occur when this is done. The waves travel faster, so the leads need to be longer. It can be difficult to maintain a constant height. The lead should be close to the ground as they approach the Zed-Meter instrument location at the tower leg. A dipole test should be recorded to confirm that the impedance of the leads is constant and that the duration of the test signal is sufficiently long. 9. Can a structure without a shield wire be measured, and if so, what results can I expect if that structure is in high resistivity soil? The Zed-Meter instrument is designed from the ground up to report the impedance of the local tower, in parallel with the surge impedance of the shield wire. If the soil resistivity is quite high (over 5000 m) and the tower base has a small surface area and geometric radius, the tower resistance can be much less than the surge impedance of the OHGWs, which is typically 170250 . This will give a rising Z(t) profile and a result with a high standard deviation.

A-2

10. Must the measurement leads be grounded? And will this affect the results? The leads do not need to be grounded except when there are high induced potentials and currents. Tests with and without grounding give the same results in the 500-1000 ns sweet spot when a sufficient number of test pulses are averaged. If the test leads are grounded with resistance of 2000 or less, it might be possible to obtain a Z(t) profile in the time range from 2000 to 10,000 ns that shows the influence of adjacent towers. 11. How do the impedance (Z) results differ from typical resistance measurements and why? This depends on the type of grounding system. If there are buried wires, the value of Z should be much higher than the low-frequency value, especially at the critical measurement time of 500 to 700 ns. The impedance profile will fall with increasing time. If, instead, there are only the foundations of the tower and a few local ground rods, Z will be less than the low-frequency resistance, and Z will increase as time moves from 500 ns to 1000 ns. In the first case, the series inductance of the grounding system gives a result with an L/R time constant that is initially high and falls to the level value. In the second case, the capacitance of the soil behaves in the opposite way, initially low and rising to settle at the level value. 12. What does zed stand for? The normal symbol for impedance is Z. In some countries, such as Canada and South Africa, this letter is pronounced zed. The original concept was developed in New Brunswick and refined through projects managed by EPRI staff of South African origin who also pronounce the letter zed rather than the American zee. Zed-Meter is a registered trademark of EPRI.

A-3

B
SPECIFICATIONS
Tables B-1 and B-2 list the physical and electrical specifications of the Zed-Meter instrument. Table B-3 describes the software outputs. Table B-4 lists the accessories that are supplied, and Table B-5 lists recommended additional accessories. Figure B-1 is a photograph of the instrument.
Table B-1 Zed-Meter instrument physical attributes Parameter Specification

Power supply Power consumption Mass External dimensions Operating time before charging Operating conditions
Table B-2 Zed-Meter instrument electrical outputs Parameter

100240 V, 5060 Hz 100 W maximum 30 kg 450 300 200 mm (width height depth) 18 hours, depending on ambient temperature -20 to 40C, 5% to 95% relative humidity non-condensing

Specification

Internal communication USB wireless support Internal signal system

USB 2.0 (3 devices) No Electric fence pulse generator meeting Underwriters Laboratories (UL) and Canadian Standards Association (CSA) standards; proprietary square pulse-shaping circuit 500 V open circuit, 1 A short circuit; 250 V at 0.5 A into 500 typical <10 ns <50 ns with typical leads >20 s 100 MS/s, four channels simultaneous 8 bits (0.4%) 80,000 points (80 s), 4 channels 200 MHz 8 ns (CM-100-M) 50% per ms -20 to 40C, 5 to 95% relative humidity non-condensing
B-1

Square pulse output Square pulse rise time Square pulse settling time Square pulse duration Digitizer sample rate Digitizer resolution Digitizer memory Digitizer bandwidth Current sensor rise time Current sensor droop Operating conditions

Table B-3 Zed-Meter system software outputs Parameter Specification

Software platform Log file

Labview 8.1 for Windows XP operating system Text (.txt) file for all tested towers, consisting of the following fields: date, line name, tower number, impedance (), voltage, current (mA), comments Text (.txt) file for every tower with the same line name, consisting of the following fields: date, line name, tower number, impedance (), voltage, current (mA), comments Text (.txt) file for every tested tower, consisting of the following fields: File name Operator name Date and time Line name Structure (tower) type Structure (tower)number

Line report

Data file

Line voltage Target impedance Comments Impedance measurement start (ns) Impedance measurement stop (ns) Ground impedance 8000 rows of the following data: time (ns), remote voltage (V), lead current (mA), structure current (mA), lead impedance (), structure impedance ()

B-2

Table B-4 Supplied accessories Item Quantity

50-cm RG58 or RG59 coaxial cable terminated in BNC connector Sheet metal welding clamp or truck mirror-style portable CB antenna mount with BNC connector 10 , 100 , and 500 calibration resistors
Table B-5 Recommended accessories Item

1 1 1

Quantity

Voltmeter Insulated gloves and covers in accordance with local safe work practices for grounded equipment Opaque jacket to shield screen and instrument from direct sun Ground spikes and hammers Cable reels for 120 m of coaxial cable

1 2 1 2 2

B-3

Figure B-1 The EPRI Zed-Meter Instrument

B-4

C
TROUBLESHOOTING GUIDE
Table C-1 lists troubleshooting information for initial operation; Table C-2, for dipole tests; and Table C-3, for tower impedance tests.
Table C-1 Initial operation Symptom Cause Correction

Power switch is off No power The battery is discharged Battery is damaged from excessive heat or cold USB cable is disconnected or loose Weak battery Weak spark gap Overheated pulse generator No pulse output Failed supervisory circuit

Turn power switch on so that red light is illuminated. Connect charger and confirm that charging light is on. Replace battery. Disconnect and reconnect USB cable. Wait for computer to indicate that it has found new hardware for three different devices before restarting program. Charge battery. Replace CG90L spark gap. Place meter in shade to cool to less than 40C. Restart instrument and check that all three USB devices are found. Disconnect supervisory circuit from pulse generator. Check independent operation of pulse generator with four D-cell batteries. Tighten connection. Replace CG90L spark gap. This is normal. Internal or external short circuit. This is normal. Oscillations can be reduced by using shorter leads. Test leads are too long or connections are lossy. Reduce lead length. Use proper radio frequency connectors or splices, not alligator clips. Conduct current transformer gain calibration process. This is normal. The resistor has small parasitic inductance and capacitance, and the test leads are not perfect. Conduct current transformer droop calibration process.

Battery does not charge Computer does not recognize instruments

Inconsistent pulse output

Loose connection Weak spark gap

Pulser makes noise

Like a mouse-click, once per second High-pitched whine Unequal for a short time (<100 ns)

Measured currents into test resistor are unequal

Unequal for >50 ns but eventually become equal Unequal for 300 ns or more

Measured impedance of fixed resistor changes with time

Changes in the time scale of 0 to 100 ns Changes from 1 to 20 s

C-1

Table C-2 Dipole test Symptom Cause Correction

Low reading, <200

Connection to voltage lead open Signal clipping

Restore connection using BNC tee and short coaxial cable. Restart instrument or reset autorange feature. Restore connection. Walk over lead to push it closer to ground level. Remove connection to tower. Re-route leads, zigzag if necessary. Increase number of samples averaged. Restart instrument or reset autorange feature. Connect coaxial cable terminations directly to instrument if oscillations persist for more than 300 ns. Use coaxial cable rather than bare wire.

High reading, >1000

Connection to current lead open Variable height of lead over ground Test lead still connected to tower Test leads run across terrain with different soil resistivity High noise level on test cables

Uneven reading with high relative standard deviation

Very low standard deviation Initial oscillations in voltage, current, and impedance

Signal clipping Sloppy wiring using banana plugs and wires

Early change in dipole impedance at 700 ns Drop in dipole impedance at time corresponding to 0.50.7 c Increase in dipole impedance at time corresponding to 0.80.9 c

Test leads too short

Add extensions to test leads, make sure all cable pulled off reels, snake cable in large zigzag pattern, if necessary. This is normal.

Test leads grounded in lowresistivity soil Test leads open on frozen or highresistivity soil

This is normal.

C-2

Table C-3 Tower impedance test Symptom Cause Correction

Low reading, < 0.2

Connection to voltage lead open Signal clipping

Restore connection using BNC tee and short coaxial cable. Restart instrument or reset autorange feature. Restore connection. Repeat dipole test and adjust leads to obtain satisfactory result. Walk over lead to push it closer to ground level. Move tower clamp as close as possible to base of selected leg. Re-route leads, zigzag if necessary. Increase number of samples averaged. Insert10- resistance in series with tower lead and repeat; result should be 10 higher. Restart instrument or reset autorange feature. Adjust sweet spot if these persist for more than 500 ns. Add extensions to test leads, make sure all cable pulled off reels, snake cable in large zigzag pattern, if necessary. Take readings in two sweet spots, near 500 ns and 1000 ns. Take readings in two sweet spots, near 500 ns and 1000 ns.

High reading, >1000 Uneven reading with high relative standard deviation

Connection to current lead open Unsatisfactory lead impedance in dipole test Variable height of lead over ground Test lead connected high up on tower leg Test leads run across terrain with different soil resistivity High noise level on test cables Low tower impedance (<5 )

Constant reading with very low standard deviation Initial oscillations in voltage, current, and impedance Early change in impedance at 700 ns Rising impedance profile with time, 5001000 ns Falling impedance profile with time, 5001000 ns

Signal clipping Tower surge impedance in parallel with ground electrode Test leads too short

Soil permittivity Inductance of distributed ground electrodes, such as counterpoise

C-3

D
HARDWARE EVOLUTION
The Zed-Meter instrument has had an extensive period of development, starting with feasibility studies hosted by American Electric Power in 2003 and continuing to the present day. This appendix documents the equipment configuration of each version.
Generation 1

Figure D-1 illustrates a low-cost, first-generation Zed-Meter instrument.

Figure D-1 Low-cost version of Zed-Meter instrument hardware, 2006

D-1

Generation 2

After experiencing difficulties in negotiating a suitable contract for commercialization, EPRI decided to develop the Zed-Meter instrument in-house. Figure D-2 is a block diagram of the second-generation Zed-Meter instrument, and Figure D-3 is an internal photograph.

Figure D-2 Block diagram of EPRI Zed-Meter instrument, 2007

D-2

Figure D-3 Internal view of EPRI Zed-Meter instrument prototype, 2007

D-3

E
THE CURRENT REACTION LEAD
The current reaction lead plays an important role in the success of Zed-Meter instrument testing. If the lead is correctly selected and placed on the ground, a constant current is fed into the tower, and the voltage response gives an accurate measure of the lightning response. However, if the conditions on the right of way do not allow a straight run, or if the operator is inexperienced and does not pull all the cable off the reel, the results will be poor. The software in the Zed-Meter instrument can help the operator to recognize when there might be a problem with the routing of the current reaction lead. However, it is also helpful to read this appendix to understand where the overall requirements originated and to see quantitatively how disturbances might or might not influence the results.

Dipole Impedance Test Theory


When a step voltage wave is applied to a pair of wires (see Figure E-1) laid on the surface of the earth and running in different directions, a traveling electromagnetic wave is formed. The parallel capacitance of the leads to ground is charged up at a rate limited by the series inductance. The high-frequency surge impedance Z is a scalar value, given by Z=(L/C)1/2. If the step voltage is 200 V, a surge impedance Z=500 will result in a current of 400 mA. This situation, with surge voltage and current related by Z, will persist until there is an impedance discontinuity. Figure E-1 shows ground rods located 90 m away from the source in each direction. In soil with low resistivity, the ground rods will have impedance that is somewhat less than the surge impedance of the cable over ground.

E-1

Voltage wave with low soil resistivity


SIGg = 0.01 S/m; EPSRg = 10 A C 10 m D 30 m E 45 m

Voltage wave with high soil resistivity


SIGg = 0.001 S/m; EPSRg = 10 A C 10 m D 30 m E 45 m

0.6 0.5 0.4 Current [A] 0.3 0.2 0.1 0 -0.1

0.6 0.5 0.4 Current [A] 0.3 0.2 0.1 0 -0.1

0.5

1 1.5 Time [microseconds]

2.5

0.5

1 1.5 Time [microseconds]

2.5

Currents along 90-m propagation line (100 m)

Currents along 90-m propagation line (1000 m)

Figure E-1 Voltage and current traveling waves for dipole antenna near earth surface

In both cases, a current of about 400 mA initially flows in response to an applied voltage pulse of 200 V. The surge impedance of the wire over 100 m soil is about 450 . Calculations suggest that this surge impedance increases to about 530 over 1000 m soil, if everything else is the same. The time from just after the voltage spike to the transition at 0.6 s is the domain in which the Zed-Meter instrument takes its most important reading. After about 0.6 s, the currents in these two situations diverge. The low-resistance ground rod in 100-m soil returns a current wave that is about 20% higher than the initial value. This reflection coefficient of 0.2 means that the ground rod resistance is about 300 . In the case of the high, 1000-m soil resistivity, the reflection coefficient is nearly unity, meaning that the termination impedances of >3000 are much higher than the surge impedance of the propagation lines.

E-2

The reflection coefficient for current, -, is the negative of the voltage reflection coefficient, given by Equation E-1, where Z1 is the surge impedance of the lead over ground, and Z2 is the lumped value of the resistance of the ground rod at the remote end.

Z 2 Z1 Z 2 + Z1

Eq. E-1

In Figure E-1, the currents in the propagation line in low-resistivity soil stop abruptly at 1.4 s. This was the width of the applied voltage pulse in the numerical simulations. The currents in the high-resistivity soil are poorly behaved, with considerable oscillations above and below zero at the source end after the pulse voltage drops to zero. One improvement made in the current version of the Zed-Meter instrument was to extend the pulse width. It was practical to make this pulse very long compared to 1.4 s and still maintain electrical safety, by using a UL-approved electric fence pulse generator and tune-up circuit.

Theoretical Variation of Dipole Impedance with Height


The traditional theoretical value for the surge impedance of the dipole over a perfect ground plane is easy to compute, as shown in Equation E-2.
2h Z = 60 ln r
Eq. E-2

With an RG-58 coaxial cable, the radius of the sheath, r, is 1.8 mm, and it will typically be placed on the earth at a height, h, of perhaps 100 mm. This gives a surge impedance of 282 . If the wire is running through tall grass that keeps it at a constant height of h=1000 mm (1 m), the impedance from Equation E-2 will increase to 421 , and it should fall to 144 at a height of 10 mm. The measured values of the dipole surge impedance should settle to double the estimate from Equation E-2. However, the configuration in Figure E-1 reports the total voltage (across both legs) divided by the current in each leg, and the result should match Equation E-1 if the voltage divides equally between the two. Depending on the vegetation, the reaction and potential reference leads can be near the surface of the soil, or they can be suspended a meter or more off the ground. A lead close to the ground will have lower and more constant surge impedance as well as a slower propagation time. These factors will give a higher-quality measurement. To explain the practical extremes, Figure E-2 shows that a dipole of RG-58 coaxial cables at a height of 0.01 m over low-resistivity soil has lower impedance (576 ) when compared to 983 for 1 m above ground. Also, the wire close to the ground has a much wider time region, or sweet spot, from 400 to 1100 ns in which to establish this impedance value.

E-3

Sim# I2-C-L001-NGR-50 3000 2500 Impedance at feed point [Ohm] 2000 1500 1000 500 0 -500 -1000 -1500 -2000 0 0.5 Median 450-650 ns = 576 1 1.5 2 Time [microseconds] 2.5 3 Vsource / I Seg.No.201

Sim# I2-C-L1-NGR-50 3000 2500 Impedance at feed point [Ohm] 2000 1500 1000 500 0 -500 -1000 -1500 -2000 0 0.5 Median 450-650 ns = 983 1 1.5 2 Time [microseconds] 2.5 3 Vsource / I Seg.No.201

Wide region for impedance determination

Narrow region for impedance determination

Z(t) of dipole, 0.01 m over 50 m

Z(t) of dipole, 1 m over 50 m

Figure E-2 Effect of coaxial cable height above ground on impedance profile of dipole leads

This means that the coaxial cable should be placed as close to the ground as possible. If possible, it is best to walk back along the lead to force it close to the ground. If this is not feasible, the lead length should be increased. Two other main factors affect the surge impedancelead orientation and nonzero soil resistivity. There are no simple formulas for calculating the impedance over lossy ground that can also have high relative permittivity. Instead, numerical calculations are used to obtain estimates.

Theoretical Variation of Dipole Impedance with Soil Resistivity


Above lossy ground, the change in field compared to perfectly conducting ground can be visualized in several different ways. The concepts are generally based on the square root of (resistivity divided by frequency), (/f), and include the following: Skin depththe greater the resistivity and lower the frequency, the greater the depth of penetration. For lightning (100 kHz) and 100 m, this is about 16 m. Complex depth, where the conductor image has an imaginary term, also a function of (/f), with a value of 1/j (11 m) for 100 kHz and 100 m. Carsons equations, where an equivalent return depth of 21 m is used. Darveniza developed a semi-empirical model which states that the effective height of a distribution line can be modeled as heff=h+0.15 when calculating induced overvoltages from nearby lightning [11]. Figure E-3 shows that the Darveniza and NEC-4 numerical models are in close agreement.

E-4

Figure E-3 Predicted variation in Zed-Meter instrument test lead (3.6-mm diameter on ground surface) with underlying soil resistivity using complex depth, Darveniza and NEC-4 models

Calculation of impedance using complex depth at 1 MHz gives about the same results as the NEC-4 and Darveniza methods. None of these models are intended specifically for application in cases such as the Zed-Meter instrument application in which the leads are located close to the ground.

Theoretical Variation of Dipole Impedance with Lead Orientation


The angle between the two test leads can vary from 180 to 45 without making much change in the dipole impedance. The impedances were calculated using the NEC-4 computer program for the Zed-Meter instrument pulse shape in the time range from 450 to 650 ns. A relative permittivity value of 10 was assumed in all cases. Figure E-4 shows that, for cables laid close to the ground at a height of 10 mm, the effect of soil resistivity is considerable. Impedance for the reference configuration, 180 orientation, increases from 576 for highly conductive 50 m soil to 834 for 20,000 m rock. If the lead is at a height of 1 m, the impedance is generally higher, as expected, and the influence of soil resistivity is reduced. The graphs include calculations for the case in which one lead is straight and the other makes a zigzag or meander line, turning 90 every 25 m. This configuration is useful in tight areas. As long as the path is straight between the corners, the dipole impedance of this pattern falls between those of the 90 and 45 orientations.

E-5

RG-58 coaxial cable, 10 mm above ground

RG-58 coaxial cable, 1 m above ground

Figure E-4 Effect of soil resistivity on input impedance of dipole for five lead orientations

The dipole impedance is a logarithmic function of height above ground as well as a complex function of soil resistivity. At present, there is no simple way to invert the dipole impedance to obtain an estimate of the average soil resistivity on its own. However, the important point is that the dipole impedance does not depend much on lead orientation. This means that, as long as leads running in parallel, 3 m apart, are avoided, the lead orientation will not have a major effect on the Zed-Meter instruments results. Generally, the longer the length of each leg of the dipole, the greater the duration of the surgeimpedance response. A recent calculation of the propagation velocity of cables directly buried and insulated from ground suggests that, from 100 kHz to 10 MHz, the propagation velocity is in the range of 0.2 to 0.55 times the speed of light [10]. (See Figure E-5.)

E-6

Figure E-5 Propagation velocity of electromagnetic waves in counterpoise and single-conductor cable for soil resistivity of 1000 and 4000 m (1 and 0.25 mS/m)

Calculations and numerical modeling both suggest that the propagation velocity is increased closer to the speed of light when the nearby soil resistivity is high. No matter what the propagation velocity, doubling the lead length will double the travel time if the underlying soil is uniform.

Conductors Running in Parallel with Test Leads


If a lead runs parallel to and close to another long horizontal conductor such as a fence, the surge impedance of this leg will be reduced. Physically, the fence is providing a parallel path for the surge potential on the test lead. The surge impedance of this path is given by the series sum of the mutual surge impedance from test lead to fence and the self-surge impedance of the lead in isolation. (See Figure E-6.)

E-7

Test Lead

Dab

Parallel Conductor hb

ha

Dab

ha

hb

Image of Test Lead

Figure E-6 Mutual coupling from test lead to parallel conductor

The mutual surge impedance between the test lead and the parallel conductor over perfectly conducting earth is given by

D ab Z ab = 60 ln D ab
where Dab and Dab are the distances from the test lead to the parallel conductor and its image. The voltages and currents on the test lead and parallel conductor will be related by

va = ia Z aa + ib Z ab vb = ia Z ab + ib Z bb
where Zaa is the surge impedance of the test lead, and Zbb is the surge impedance of the parallel conductor, estimated using Equation E-2. If the parallel conductor is insulated and floating (for example, an electric fence wire), the current ib is zero, and there will be no effect on the relation between va and ia. There will be an induced potential, vb = Zab /Zaava, on the fence. This effect will also be occurring on each of the insulated phase conductors of the transmission line on the OHGWs (which are not effectively grounded for the time that the Zed-Meter instruments test pulse is traveling along the test lead). If the parallel conductor is grounded at several points, it is at zero potential (vb=0). The current induced in the fence ib will be -iaZab/Zbb, and the measured surge impedance of the test lead will reduce to

Z2 va = Z aa ab ia Z bb

E-8

If the test lead is close to the ground (ha=0.1 m) and the parallel conductor (hb=1 m, rb=0.003 m) is 1 m away horizontally, the values work out to Dab=1.345 m, Dab=1.487 m, Zab = 6 and Zbb = 390 . This will reduce the impedance of the test lead by a negligible 0.1 . If, instead, the test lead is at the same height as the conductor, 1 m above ground, Dab=1 m, Dab = 2.236 m, Zab=48 , and the impedance of the test lead will fall by 6 . It might be possible to measure this difference if the noise level in the dipole test is low enough to yield results with low standard deviation. Overall, this calculation mainly illustrates that there is a secondary benefit if the reaction lead heights are kept close to the ground and a meter or more away from parallel conductors.

2003 Field Studies at American Electric Power


This field study was carried out on 69-kV, 138-kV, and 765-kV American Electric Power transmission lines in the Appalachian Mountains of Virginia. A pulse generator with 50- internal impedance and 440-ns pulse width was used. One measure of the overall performance of the instrumentation at that time is the record of the system response to a short circuit, which was carried out at Tower 29_19 (see Figure E-7).

Short circuit current: 4.5 A

Voltage across loop

Impedance, standard deviation

Figure E-7 Short circuit performance of Zed-Meter instrument, 2003

The impedance of the short-circuit wiring loop was measured to be <1 near the end of the current pulse. The negative swing of the current and the corresponding response in voltage have lower rise times, but they converge to nearly the same value. A number of options for the current reaction lead were then tested (see Figure E-8).

E-9

Figure E-8 Arrangement of current monitors and pulse output cable in Zed-Meter instrument, 2003

An Avtech AVL-S pulse generator with 50- impedance and 200-V open circuit voltage was used along with an internal trigger circuit, running at 1 kHz. The green current transformer used to measure current into the tower in Figure E-7 was a Pearson model 2110, with a sensitivity of 0.5 V/A into 50- termination impedance. An Ion Physics current transformer, yellow in Figure E-7, was used to measure the current into the remote current lead, with a sensitivity of 1 V/A into 50 . Both current transformers had rise times of >10 ns and fall time constants of >10 s. Figure E-9 shows the applied voltage across the ground leads, and Figure E-10 shows the measured currents for the most typical case, injection to the sheath of a single coaxial cable.

E-10

Figure E-9 Typical applied pulse voltage between tower leg and remote current lead

Figure E-10 Measured currents into tower leg (blue) and remote current lead (violet)

The currents are being injected into two arms of an antenna system. The tower leg has a larger surface area compared to the sheath of the RG-58 cable, so its impedance is initially lower, leading to a higher current. However, as the surge travels further up the tower and along the wire, the input impedance increases to a steady value. Configurations with lower surge impedance have faster settling times, giving more time for valid measurements with short pulse widths. This would be especially important in measurements on lines with short span lengths. Table E-1 shows the effects of four different configurations on the injection impedance at the source.

E-11

Table E-1 Surge impedances of four possible reaction wires using RG-58 coaxial cables Input impedance between 765-kV lattice tower and core of single RG-58 coaxial cable 200 ns357 300 ns366 400 ns361 Pulse voltage180.6 V

Input impedance between 765-kV lattice tower and sheath of single RG-58 coaxial cable 200 ns308 300 ns312 400 ns311 Pulse voltage175.0 V

Input impedance between 765-kV tower 29-22 and cores of two parallel RG-58 cables 200 ns189 300 ns202 400 ns198 Pulse voltage163.5 V

Input impedance between 765-kV tower and sheaths of two parallel RG-58 cables 200 ns162 300 ns173 400 ns167 Pulse voltage158.1 V

E-12

There was no point in using the core of the RG-58, as all this did was to place 50 impedance in series with the surge impedance of the sheath to ground. A single coaxial lead, run away from the tower at right angles to the line, settled to 361 when the wire core was energized and to (361 - 50 ) or 311 when the sheath was energized. A pair of coaxial cables placed about 1 m apart led to 200- input impedance for the core injection and (200 - 50/2) = 175 when the sheaths were both energized. For the large, 765-kV lattice tower, the two-sheath electrode gave the best results, but all the results were considered acceptable, with constant impedance after about 200 ns. The most important advantage of the double-cable reaction electrode was that the settling time was fasteron the order of 150 ns, compared to more than 300 ns for the single cable. In areas in which rights of way are restricted and there are no complications of buried counterpoise, it can be more effective to extend the sweet spot of the impedance measurement by using two cables in parallel. Practical considerations of minimizing the number of wires to be run and maintaining good electrical contact between the leads and the grounding clamps led to the selection of the singlesheath configuration. A straight lead of 90 m was selected so that any reflection from the remote ground rod would arrive in time to contaminate the measurement. The two-way transit time of the 90-m wire would be more than 600 ns. Because it is not always possible to run the current lead straight off the right of way for a distance of 100 m, a series of zigzag configurations were tested, using the full extent of the wire. Generally, these arrangements led to more ringing on the currents, and the currents did not settle to the same value as quickly. A figure of merit of the agreement between the two measured currents was adopted in the evaluation of these results. Early in this test series, the value of extending the pulse width was recognized to take advantage of the slow propagation velocity along the reaction lead.

2005 Current Reaction Lead Studies at the CN Tower in Toronto


A series of preliminary tests of current injection methods were conducted at the Toronto CN tower. This well-known structure attracts as many as 70 lightning flashes each year due to its height above ground of 553 m. Access was granted to the CN tower foundation during a construction project. An RG-58 coaxial cable was run, starting from the grounding connection of the lightning protection system at the center of the tower and extending to an exterior point approximately 160 m away (see Figure E-11).

E-13

Current injection at lightning protection system ground in CN tower basement level Figure E-11 Tests of current injection into base of CN tower

Routing of wire through piping access

Some compromises in obtaining a straight wire path were necessary. Figure E-12 shows that the lead ran in metal conduit at several locations in order to traverse walls. However, for the most part and wherever possible, the wire was laid on the surface of soil or concrete, heading radially out from the tower. Figure E-11 shows the path of the lead on a grass area east of the tower.

Figure E-12 Exterior routing of current reaction lead for tests at Toronto CN tower

The RG-58 cable was first terminated in an open circuit. This gave a relatively constant pulse current, as shown in Figure E-13, for 1100 ns. When the cable was extended for an additional 90 m, the pulse current was essentially the same initially, but it then stayed constant right and tracked the applied pulse voltage at and beyond 1400 ns.

E-14

Applied voltage Figure E-13 Exterior routing of current reaction lead for tests

Current in reaction lead90 m or 180 m

The various narrow points and turns in the current lead gave an imperfect reaction lead impedance that makes analysis of the results more difficult. However, the following effects were noted for this configuration:

The travel time along the cable was slow. The reflection from the open point at 90 m came at 1000 ns, corresponding to 0.6 c. There was considerable loss of high-frequency energy because the fall time corresponding to the open circuit was much slower than the measured rise time of the injected current. The rise time of current in the reaction lead, measured at five intervals along the 180-m cable in Figure E-12 showed the greatest loss of high frequency energy in the first 65 m out from the CN tower base (see Figure E-14). These currents were measured with a Pearson 110 (35 MHz) current monitor and a Fluke portable 20-MHz oscilloscope.

Figure E-14 Change in current along 180-m reaction lead

E-15

Dipole Test Results, 90-m Versus 130-m Lead Length


The bench tests in Section 3, Zed-Meter Instrument Dipole Tests of Reaction Leads, have shown some important aspects about Zed-Meter instrument testing. Those procedures should be reviewed before trying out a dipole test. When measuring low values of resistance, the effects of series test lead inductance take some time to damp out before the sweet spot, with constant voltage and current, enables a valid calculation. Also, measuring resistances of 500 or more can also lead to high-frequency oscillations. The advantage of a 130-m cable over a 90-m cable was evaluated in a test series in freezing conditions, which would be the worst case with extremely high surface resistivity in the frozen soil and snow layer. Because it is pointless to drive ground rods into frozen soil, the leads were left floating. This meant that the traveling wave reflections from the ends of the cables were marked by clear reductions or reversals of current. In Figure E-15, the sweet spot of constant and equal current in each leg starts at about 500 ns in all cases. The suitable evaluation time continues to 800 ns for the 90-m leads, corresponding to a propagation velocity of 0.75 c. The sweet spot increases to 1100 ns for the 130-m leads, also corresponding to 0.75 c. This is a worthwhile gain for an extra minute or two of walking time in each direction.

Pulse G enerator Measures current to the left wire Insulated 90 m Wire Me asure s curr ent to the right wire Insul ated 90-m Wire

Snow and/o r Frozen S oil La yer

Pulse G enerator Measures current to the l eft wire Insulate d 130 m Wire Me asure s cur rent to the right wire Insulated 1 30-m Wire

Snow and/or Frozen Soil L ayer

Figure E-15 Comparison of 90-m and 130-m RG-58 coaxial cable laid on surface of frozen soil

Dipole Test Results, 125-m Coaxial Cable Test Leads


Field measurements are now used to illustrate typical results from dipole tests. In both cases, the propagation lines used the sheaths of RG-58 coaxial cables, with a length of 125 m and a twoway propagation time of 833 ns at the speed of light. Figure E-16 compares the results on the frozen soil in Manitoba with previous results of the same test carried out in New Jersey [3].

E-16

Manitoba Hydro, Zed-Meter instrument, frozen soil, tower 360

Public Service Electric and Gas, oscilloscope, unfrozen soil, Levittown-Cinnaminson tower

Figure E-16 Dipole impedance test results from Manitoba and New Jersey

Figure E-17 shows impedances, Z(t), that were calculated from the sample-by-sample ratio of voltage divided by currents I1 and I2 in each branch of the dipole. The standard deviation of the average impedance was also taken over 30 samples that followed this value. The standard deviation fell continuously to about 800 ns, leveled off until 1000 ns, and then started to rise as the effects of the open termination of the current reaction lead began to appear. It also took about 800 ns for the Zed-Meter instruments result on frozen soil to settle down to a constant value of about 675 . The standard deviation reached a minimum of 12 in the sweet spot from 800 to 1000 ns. This is an expected value and is due to the high resistivity of the snow and ice. In comparison, the Public Service Electric and Gas dipole result settled to a constant value of 380390 in about 300 ns. The low value of surge impedance corresponds to the low (50 m) soil resistivity at this location, as does the long propagation delay of 1400 ns to the ends of the 90-m leads. The latest version of the Zed-Meter instrument hardware has an extended pulse width. This means that two separate time domains can be evaluated. The first, relying on the surge impedance of the reaction and potential leads, has just been described. When the reaction leads are both terminated in the soil with ground rods, a second period of stability will occur after two or three travel times up and down the lead. Dipole tests at National Grid (see Figure E-16) show two distinct regions of stability in the measured pulse voltages, currents, and Z(t) profiles.

E-17

Dipole test

4VW tower 234

ZN205 tower 205

Applied voltage from pulse generator across pair of reaction leads

Resulting currents into each leg: reaction lead, ground reference potential lead.

Impedance of dipole Total termination rod impedance Reflection coefficient

354 5 , 500700 ns 693 -0.43 (Current Wave)

439 8 , 500700 ns 944 -0.51 (Current Wave)

Figure E-17 Applied voltage, dipole leg currents and derived surge impedance for 90-m leads tested in the United Kingdom

With the experimental results in Figure E-17, the time for the return of reflections was 10001300 ns, and it was best seen in the current waves. This time corresponds to a propagation speed along the cables of 0.46 to 0.60 times the speed of light. The test with the lower dipole impedance value had the slower propagation speed. This is predicted by theory, as well.

E-18

In both of these cases, the current dropped when the reflections from the remote ends arrived back at the source. This means that the termination resistances were larger than the surge impedances of each reaction wire. With the new Zed-Meter instrument hardware, it is possible to look at the impedance at 5000 ns or more in order to establish a value for the sum of the ground rod resistances in series. In this case, the resistances are low enough that there is plenty of current200 to 300 mAto obtain a low-noise measurement.

Dipole Test Results, 300-m, 14-Gauge Solid Copper Wires


The Manitoba Hydro grounding test crew was equipped with hydraulic winders and spools of 14-gauge solid wire. The crew were interested to see whether this option would be a suitable substitute for the Zed-Meter instruments test leads. When they saw that extra lead length increased the sweet spot of the measurements, they proposed to increase the test lead length using the power winder equipment, while maintaining the same overall test time. A dipole impedance test was conducted, using a pair of 14-gauge wires laid on the snow at a 90 angle. The lead arrangement and the automatic readout of impedance on the laptop computer screen are shown in Figure E-18.

Lead arrangement for dipole test with 300-m 14-gauge solid copper wires

Zed-Meter instrument automatic processing of dipole impedance result

Figure E-18 Zed-Meter instrument test arrangement and results for dipole test on 14-gauge wire

E-19

Figure E-19 plots the post-processed data from this experiment, including denoising. The graph shows that the standard deviation stayed low from 1000 ns to 2000 ns and turned upward with the reflections from the open ends at 2500 ns. A two-way travel time of 2000 ns would be expected at the speed of light (1000 ns per 300 m).

Figure E-19 Dipole impedance test on 300-m wires oriented at 90, tower 762, February 8, 2008

In this case, the automatic built-in processing of the waveform indicated a reading of 636 from 800 to 900 ns. With the long settling time in the waveform, the extra length of the sweet spot with low standard deviation, from 1000 to 2000 ns, would certainly improve the measurement quality. Also notable is the long duration of the oscillations resulting from the use of banana plug adapters in the time from 0 ns to 400 ns.

E-20

F
MODELING ZED-METER INSTRUMENT LEADS WITH NEC-4 SOFTWARE
NEC-4 Software for Detailed Calculations Using Inverse Fast Fourier Transform
The NEC-4 computer program follows the same input card sequences as the NEC-2 program, with an important extension. The antenna segments are allowed to penetrate into lossy earth. At low frequencies, the response of a line segment in soil matches the value calculated from an expression for the resistance of a ground rod. As the frequency increases, the impedance changes, with considerable effects found for frequencies that have quarter-wave resonance and their multiples. Figure F-1 shows the overall methodology for evaluating the response of a tower and grounding system to the Zed-Meter instruments input. A frequency spectrum is defined from the source input. This frequency spectrumalong with a three-dimensional description of the tower, the overhead groundwires, the grounding system, the Zed-Meter instruments source location and lead layoutis used as input in the NEC-4 model. For each frequency step, the NEC-4 program computes the voltages and currents on each line segment. The output contains more information than is strictly required, so an automatic method that extracts the relevant segment currents and voltages must be developed, whether cut-and-paste for a single simulation or an automatic program or script. The computed currents for each frequency of interest are convolved with the inverse fast Fourier transform of the input source, and the result is converted back from the frequency to the time domain with the standard fast Fourier transform. Figures F-2 through F-5 illustrate the steps in the process.

F-1

Figure F-1 Methodology for establishing Zed-Meter instrument response for specific tower and lead layout

Figure F-2 Details of step 1 in NEC-4 modeling for Zed-Meter instrument results

F-2

Figure F-3 Details in step 2 of NEC-4 modeling for Zed-Meter instrument results

Figure F-4 Details in step 3 of NEC-4 modeling for Zed-Meter instrument results

F-3

Figure F-5 Details in step 4 of NEC-4 modeling for Zed-Meter instrument results

Other Modeling Software


The frequency-domain calculations with the method of moments in NEC-2 and NEC-4 are one way, but not the only way, to evaluate the response of a tower and grounding system to Zed-Meter instrument tests in a computer simulation. Finite-difference time-domain (FDTD) software offers another option.

Finding an Experienced Practitioner


There are several options for finding an antenna expert with experience running the NEC-4 computer program. Funders of the Zed-Meter instrument development program have access to internal experts through collaboration opportunities. With a well-established method (see Figures F-1 through Figure F-4), programming a novel lead arrangement or a new tower requires extra work primarily to capture the geometry of the tower segments. Ham radio operators often have practical experience with modeling the input impedance of wire antennas using NEC-2. With some coaching, they can also be helpful in evaluating coupling and in developing efficient methods to deliver high-frequency energy into metal structures. A literature search in IEEE Xplore or other databases for NEC-2, NEC-4, or FDTD, filtered for the universities nearest to your utility, can help to locate a helpful academic collaborator willing to develop student experience with a practical and verifiable electromagnetic compatibility problem.
F-4

Export Control Restrictions

The Electric Power Research Institute (EPRI)

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), with major locations in Palo Alto, California; Charlotte, North Carolina; and Knoxville, Tennessee, was established in 1973 as an independent, nonprofit center for public interest energy and environmental research. EPRI brings together members, participants, the Institute's 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. EPRI's members represent over 90% of the electricity generated in the United States. International participation represents nearly 15% of EPRI's total research, development, and demonstration program. TogetherShaping the Future of Electricity

2008 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power Research Institute, EPRI, and TOGETHERSHAPING THE FUTURE OF ELECTRICITY are registered service marks of the Electric Power Research Institute, Inc. Printed on recycled paper in the United States of America 1015904

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

Potrebbero piacerti anche