Sunteți pe pagina 1din 6

X International Symposium on

Lightning Protection
9
th
-13
th
November, 2009 Curitiba, Brazil

COMPARISON OF LOW FREQUENCY RESISTANCE AND LIGHTNING
IMPULSE IMPEDANCE ON TRANSMISSION TOWERS
William A. Chisholm
1
, Emanuel Petrache
2
, Fabio Bologna
3

1
Kinectrics/UQAC, Canada W.A.Chisholm@ieee.org
2
Kinectrics, Canada Emanuel.Petrache@kinectrics.com
3
EPRI, USA fbologna@epri.com


Abstract A new field instrument, the Zed-Meter

, has been
developed to test the low-current lightning impulse
impedance Z of transmission tower footings and ground
electrodes. Surge impedance of test leads provided effective
reaction electrodes for the measurements, even when
ungrounded on frozen soil. Cross-calibration studies were
carried out using a variety of low-frequency measurements
of footing resistance R
f
to establish the ratio Z : R
f
for
compact and distributed electrodes.

1 INTRODUCTION

It has been traditional to measure low-frequency
resistance R
f
of transmission tower structures using low
currents and to use this resistance in estimates of
transmission lightning performance. The measured
values of R
f
can be adjusted for ionization and impulse
effects [1] to give:
Low-current transient impedance Z;
High-current transient impedance Z
I
.

Theoretically [2][3] and experimentally, a number of
factors affect the ratios Z : R
f
and Z
I
: R
f
, including:
The contribution of ground plane surge response
and series inductance of distributed electrodes,
which tend to increase the ratio Z : R
f
;
Reductions in the soil resistivity with increasing
frequency, and the effect of displacement current
in the soil dielectric, which both tend to reduce
the ratio Z : R
f
;
Ionization when local electric field gradients
exceed 150-300 kV/m in soil, increasing the
apparent size of small electrodes and reducing
the ratio Z
I
: R
f
as current increases.

Recent work [4] has suggested that, while ionization can
play an important role in the response of small electrodes
such as a single driven rod, for the typical 5 to 50 m
dimensions of transmission tower foundations and ground
electrodes the role of ionization is limited. A recent
analysis suggests that, for extensive electrodes, the
ionization effect reduces the contact resistance but not
the geometric resistance, and for transmission tower
footings the geometric resistance is dominant.

In order to understand the response of the transmission
tower to lightning surges, a new test method [5][6] was
proposed and developed. This test method used a
compact current impulse source and distributed leads as
reaction antennas to measure the transient impedance Z of
transmission tower ground electrodes in-situ.

This paper describes refinements of the test method [6] to
improve repeatability and then reports the results obtained
from Zed-Meter

tests on 70 different transmission


structures, ranging in low frequency resistance from
0.1 to 200 .

2 PREVIOUS STUDIES

Many researchers have carried out studies of the impulse
response of ground electrodes. Generally, experiments
have been carried out on isolated electrodes. Analysis has
considered either compact or distributed electrodes.

2.1 Impedance versus time for rods

Bellaschi [7] applied lightning impulses in the range of
0.6 to 9 kA to driven rods. A high-voltage impulse
generator, with a plan view in Fig. 1, was used.

The typical voltages and currents were measured with a
voltage divider (VD) and cathode ray oscilloscope
(CRO). Some typical results for driven rods are shown in
Fig. 2. For currents in the range 1.5 to 6.5 kA in
waveforms AF, AX and BJ, the voltage and current
waveforms were similar, suggesting that an instantaneous
ratio Z
I
(t)=v(t)/i(t) was constant for the 6x13 s or
12x50s test waves used in this work. The ground
around the test rod exploded in waveform BT, giving a
sharp reduction in the Z
I
(t) ratio after 15 s.
329


Fig. 1 Layout for laboratory tests of impulse ionization [7]



Fig. 2 Typical waveforms showing impulse ionization of
2.7 m vertical steel rods [7]

The ratio Z
I
to 60-Hz resistance R
f
of 2.7-m vertical steel
rods decreased from 0.85 (1 kA) to 0.55 (10 kA) as a
result of ionization effects. Similar tests were carried out
[8][9] at currents up to 50 kA, confirming the behaviour
and extending the understanding of how ionization
reduces the impedance Z
I
of compact electrodes.

2.2 Impedance versus time for foundations

Recently, Yasuda et al [10] reported test results as Z
I
(t) in
Fig. 3 for high-current impulses applied to the foundation
of a 500 kV tower in soil of high resistivity.


Fig. 3 Z
I
(t) for 500-kV transmission tower footing at two
levels of impulse current, adapted from [10]

At the early time (3-5 s) of interest in calculating
flashover performance, the measured values of Z
I
(t) are
in close agreement. Later at 10-25 s, there is a broad
minimum of Z
I
(t)=27 s for 60 kA and 22 at 120 kA.
Simplified [1] and detailed [8] modelling also suggest that
surge currents of more than 100 kA are needed to cause
ionization effects in large foundations.

2.3 Impedance versus time for buried horizontal wires

Modeling of the response of buried horizontal wires
(counterpoise) to surges started with the work of Bewley
[11]. Other tests of the transition from an initial surge
impedance to low-frequency resistance of 30.5-m buried
wire can be found in Fig. 4 [3].



Fig. 4 Transient impedance of 30.5-m counterpoise in six soil
types for high-voltage impulse tests (adapted from [3])

The modeling in Fig. 4 assumed an initial impedance of
70-90 that fell or rose to levels close to the low-
frequency resistance with time constants of 0.5-3 s.

Fig. 5 shows a recent test method [12] that uses a local
impulse source of approximately 1 kV to provide impulse
current of 2-6 A into driven-rod and buried-wire
electrodes. A multiple-rod auxiliary electrode is located
30 to 60 m away as a return path for the injected current
from the floating generator.


Fig. 5 Experimental setup of Visacro and Rosado [12] for
testing impulse response of ground electrodes

330

The electrode under test provides the ground reference for
the oscilloscope, and potential to remote earth is
measured on a lead oriented at 90. A typical result for
an 18-m buried wire is shown in Fig. 6.

Fig. 6 Transient impedance v(t)/i(t) of 18-m horizontal
electrode in 300 m soil from setup of Fig. 5 [12]

The location of the impulse source and the measurement
equipment together at the electrode, rather than separated
as shown in [13], is an important advantage in testing Z(t)
because the rise time of the test signal degrades with
distance along a horizontal line near ground [6].

3 ZED-METER METHOD

The Zed-Meter is a test instrument optimized for
measuring the transient impedance Z(t) of ground
electrodes. The original test method [5] used the sheath
of a 60-m coaxial cable to achieve a surge impedance of
400 with a two-way propagation time longer than
440 ns as a current reaction electrode. With a 200-V
pulse, a tower base current of about 0.5 A was measured
using a wideband current transformer. Transient voltage
was measured from tower base to a remote potential lead,
which was also 60 m long and oriented at 90 to the
current reaction lead.

The R&D program leading to the present configuration
included:
Adapting a commercial impulse source with
appropriate safety certifications for human and
animal contacts;
Extension of pulse width from 440 ns to cover a
longer time of interest of 1.5 s or more;
Validation of the constant surge impedance of
test leads, in 180 and 90 dipole configuration,
fed from the center and grounded or open at
remote ends;
Development of automated averaging and
analysis critera for Z(t) waveforms to report Z;
Crosss-checks of Z with measurements of R
f
;
Packaging of instrument and test leads.

The Zed-Meter test methods were co-developed with a
series of numerical simulations [6] using the NEC-4
computer program [14] for frequency-domain analysis of
antennas in soil of finite conductivity. Numerical
simulations were carried out of a steel lattice transmission
tower with four footings, equipped with a single overhead
groundwire. A Fourier transform provided the frequency
spectrum of the time-domain pulse input. The NEC-4
computer program was used to compute the potential rise
from tower base to remote potential lead at each
frequency, considering the influence of the currents in the
reaction lead, tower and OHGW. An inverse Fourier
transform then gave the time-domain voltage and current
waveforms.

The NEC-4 program confirmed a finding from initial field
tests of the Zed-Meter, which was that the Z(t) results
were relatively insensitive to the lead orientation. For
example, Fig. 7 shows that Z(t) decays slowly from 7 to
5 in the time from 100 to 700 ns, with a medain value
of 6 from 450 to 650 ns.


Fig. 7 Z(t) profile for test leads at 90 orientation

The rise in impedance after 750 ns in Fig. 7 corresponds
to the return of a reflection from the end of the current
reaction lead, which terminates the period for which the
reaction wire can be treated as constant surge impedance.
The drop in impedance at 1.5 s in Fig. 6 corresponds to
the falling edge of the applied voltage pulse in the model.

It was anticipated that leads oriented at right angles would
have a minimum of coupling. In fact, the orientation of
leads at 180 led to a faster settling time to constant
impedance, as shown in Fig. 8. This lead orientation
along the right-of-way is also more practical for most
measurements, as the rights of way are maintained for
easy physical access and the adjacent property may be
private. This was adopted as the reference configuration.

331


Fig. 8 Z(t) profile for test leads running along transmission
right-of-way (ROW) at 180 orientation

In both of these cases, a median impedance taken over the
200-ns range from 450 to 650 ns gave satisfactory results,
measured by a low standard deviation over time and using
surge currents into both the reaction lead and tower.

4 ZED-METER RESULTS

The Zed-Meter is normally used as a three-terminal
device, with connections to the current reaction lead, the
tower under test and the remote potential lead. It can be
used as a two-terminal device by connecting two leads
together. The two-terminal results show principles of
operation, while the three-terminal results give the desired
values of Z for the transmission towers.

4.1 Two-terminal measurements with the Zed-Meter

In two-terminal mode, the Zed-Meter is well suited for
measuring the input impedance of dipole leads laid on the
ground [6], with a typical result in Fig. 9.


Fig. 9 Zed-Meter test: 90-m terminated diople

Normally, if the leads are terminated with a temporary
driven rods (totaling 944 in Fig. 9), the resistance will
be higher than the surge impedance of the reaction leads
(444 ). The initial surge impedance is measured with a
2% relative standard deviation. After the slow reflections
from the terminations, the instrument eventually reaches a
second stable reading.

Two separate field studies [15][16], carried out in frozen
conditions at temperatures as low as -20C, established
that it was completely unnecessary to terminate the test
leads in driven rods or other auxiliary electrodes,
provided that Z(t) was analyzed prior to the arrival of
reflections from the open ends.


Fig. 10 Zed-Meter test: 300-m diople, open ends, frozen soil
[15]

In Fig. 10, the 300-m test leads over frozen soil provided
constant 660- surge impedance from 300 ns to 2.5 s.

When used to test the input impedance of buried wires
against the combined impedance of all other tower
foundation and guy electrodes in parallel [16], the Zed-
Meter gave an impedance profile Z(t) that decayed with
time as shown in Fig. 11.


Fig. 11 Zed-Meter test: Four, radial 38-m wires with R
f
=26-

332

In this case, an L/R
f
time constant of 0.3 s corresponds
to an equivalent series inductance of 8 H for the four
buried wires in parallel.

4.2 Three-terminal measurements with the Zed-Meter

Generally, when the Zed-Meter was used in three-
terminal mode to measure the impedance of transmission
towers [16], Z(t) was relatively constant after 300 ns, as
shown in Fig. 12.


Fig. 12 Zed-Meter measurement of isolated 500-kV lattice
tower with R
f
= 23

With adequate test lead length and a low-impedance
connection to the tower, the relative standard deviation of
the measurements decreased below 10% at times greater
than 300 ns, and stayed at this level until 1100-1400 ns.
In general, the profiles of Z(t) were constant over this
time as in Fig. 12 or increased slightly from 500 ns to
1000 ns as in Fig. 13.


Fig. 13 Zed-Meter measurement of 161-kV steel pole with
0.22-m radius and 175 m soil resistivity, R
f
= 40

Some measurements, in particular on towers with
multiple guy wires, did not settle quickly to constant
values of Z. The worst examples were H-frame 230-kV
towers with sixteen guy wires [16]. The test results in Fig.
14 show significant oscillation and high relative standard
deviation up to about 700-750 ns.

Fig. 14 Zed-Meter measurement of 230-kV H-frame tower
with sixteen guy wires

In some cases, R
f
was measured after isolating the
overhead groundwires (OHGW) from the structure. In all
of these field tests, the measured values of Z did not
change appreciably (within observation error), because
the parallel impedance of the OHGW was significantly
greater than the measured values of Z in each case

Three 500-kV towers with high values of R
f
(163, 102
and 208 ) also had isolated overhead groundwires. The
towers behaved as parallel capacitance [16], given by
tower travel time divided by tower surge impedance.

4.3 Comparing Zed-Meter and low-frequency results

Low-frequency resistance was measured on many of the
towers at the same time that Zed-Meter tests were
performed. The test methods included traditional fall-of-
potential measurements; tests of individual leg resistance
using a 30-cm clamp-on current transformer; tests of
isolated tower resistance using an oblique probe method,
taking a voltage profile on a traverse at 60 to the current
lead and calculating both R
f
and local soil resistivity.
Fig. 15 and Fig. 16 show the relations between R
f
and Z.


Fig. 15 Comparison of Zed-Meter results at 500 and 1000 ns
with R
f
for compact electrodes
333

Fig. 15 shows that impedances read out at 500 ns and
1000 ns were highly correlated, but significantly lower,
than the R
f
measurements taken at the same time in test
campaigns. Towers with compact (local foundation)
electrodes in Fig. 15 had Z<R
f
with good correlation for
all but the lowest values of R
f
<3 and Z<7.

In contrast, Fig. 16 shows that all ten distributed
electrodes with buried wire length greater than 40 m had
Z>R
f
at 500 ns, and only two of these ten had Z<R
f
at
800 ns. The four towers treated with four, 38-m radial
wires tended to behave more like compact electrodes,
with Z<R
f
at 800 ns.


Fig. 16 Comparison of Zed-Meter results at 500 and 1000 ns
with R
f
for distributed electrodes [16]

The results in Fig. 16 suggest that distributed electrodes
have high initial impulse impedance, as has been reported
in many previous studies.

5 CONCLUSION

With an optimum placement of surge source, it is possible
to inject fast-rising surge currents into the base of
transmission tower foundations and ground electrodes
using the surge impedance of a wire on the surface of the
earth for the return path, rather than the resistance of an
auxiliary remote electrode. In some cases, tests with no
auxiliary electrodes at all over frozen soil gave
satisfactory results as long as reaction wire length was
sufficiently long.

For many tested transmission tower foundation
electrodes, the settling time to a constant ratio of voltage
to current was on the order of 300 ns. Settling time was
increased for guyed structures and for lines with insulated
overhead groundwires. Close agreement between current
into reaction lead and current into the tower was a good
indicator that oscillations had decayed to acceptable
levels.

The ratio of measured transient impedance Z to low-
frequency resistance R
f
varied, depending on soil
resistivity and type of electrode. Generally, Z<R
f
for the
tested towers with compact electrodes and high soil
resistivity, and Z>R
f
for distributed electrodes in
agreement with previous findings and modelling.

6 REFERENCES

[1] W. A. Chisholm and W. Janischewskyj, "Lightning surge
response of ground electrodes", IEEE Transactions on
Power Delivery, vol. 4, n. 2, pp. 1329-1337, Jan. 1989.
[2] S. Visacro. A comprehensive approach to the grounding
response to lightning currents. IEEE Transactions on
Power Delivery, vol. 22, n. 1, pp. 381-386, Jan. 2007.
[3] S.S. Devgan and E.R. Whitehead, Analytical Models for
Distributed Grounding Systems, IEEE Transactions on
Power Apparatus and Systems vol. PAS-92, n. 5, pp. 1763-
1770, Sept. 1973.
[4] CIGRE Working Group C4.406, Performance of trans-
mission line grounding electrodes for lightning currents.
Electra, in press
[5] W. A. Chisholm, J. Williamson, S. Burnett and G. Hodges,
"Recent Progress in Design and Test Methods for
Transmission Line Ground Electrodes", VII International
Symposium on Lightning Protection, Curitiba, Nov. 2003.
[6] E. Petrache, W.A. Chisholm and A. Phillips, "Evaluating the
transient impedance of transmission line towers", IX
International Symposium on Lightning Protection, Foz do
Iguau, Brazil, Nov. 2007.
[7] P.L. Bellaschi, Impulse and 60-Cycle Characteristics of
Driven Grounds, Transactions of the American Institute of
Electrical Engineers, vol. 60, n. 3, pp. 123-128, Mar. 1941.
[8] A.C. Liew and M. Darveniza, Dynamic model of impulse
characteristics of concentrated earths, Proceedings of the
IEE, vol. 121, n. 2, pp. 123135, 1974.
[9] A. Phillips and J.G. Anderson, High Current Impulse
Testing of Full-Scale Ground Electrodes, Technical Report
1006866, EPRI, Palo Alto, California. 2002.
[10] Y. Yasuda, S. Kondo, T. Hara, J. Ikeda, Y. Sonoi and
Y. Furoka, Measurement of Soil-Ionization Characteristics
of Grounding and its Analysis using Dynamic Grounding
Model, IEEJ Transactions on Power and Energy, vol.
123B, n. 6, pp. 718-724, Jun 2003.
[11] L.V. Bewley, Theory and Tests of the Counterpoise,
Transactions of the American Institute of Electrical
Engineers, vol. 53, n. 8, pp. 1163-1172, Aug. 1934.
[12] S. Visacro and G. Rosado, Response of Grounding
Electrodes to Impulsive Currents: An Experimental
Evaluation, IEEE Transactions on Electromagnetic
Compatibility, vol. 51, n. 1, pp. 161-164, Feb. 2009.
[13] CIGRE Working Group C4.2.02, Methods for measuring the
earth resistance of transmission towers equipped with earth
wires, CIGRE Brochure 275, Jun. 2005.
[14] G.J. Burke, Numerical Electromagnetics Code NEC-4
Method of Moments, Lawrence Livermore National
Laboratory UCRL-MA-109338, 1992.
[15] W.A. Chisholm, Zed-Meter Trials at Manitoba Hydro,
Proving the Concept and Equipment in Frozen Soil, EPRI,
Palo Alto, CA and Manitoba Hydro: 2008.
[16] W. A. Chisholm, Summary of Zed-Meter Field Tests:
Transient Imedance of Transmission Line Grounds, EPRI,
Palo Alto, CA: 2006. Product ID 1012314.
334

S-ar putea să vă placă și