Anda di halaman 1dari 7

264

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 1, JANUARY 2005

Selection of Station Insulators With Respect to


Ice and SnowPart I: Technical Context and
Environmental Exposure
M. Farzaneh, Chairman and Principal Author (Contributors in Alphabetical Order), T. Baker, A. Bernstorf,
J. T. Burnham, T. Carreira, E. Cherney, W. A. Chisholm, R. Christman, R. Cole, J. Cortinas, C. de Tourreil,
J. F. Drapeau, J. Farzaneh-Dehkordi, S. Fikke, R. Gorur, T. Grisham, I. Gutman, J. Kuffel, A. Phillips, G. Powell,
L. Rolfseng, M. Roy, T. Rozek, D. L. Ruff, A. Schwalm, V. Sklenicka, G. Stewart, R. Sundararajan, M. Szeto,
R. Tay, and J. Zhang

AbstractIn selecting adequate insulators for substations at distribution and transmission voltage levels, special measures may be
needed in locations exposed to freezing conditions. This first part
of the paper describes the environmental and insulator parameters
that influence the risk of flashover on station insulators exposed to
freezing conditions.
Index TermsClimate, contamination, flashover, fog, high
voltage, ice, insulator coordination, snow, station insulators.

I. INTRODUCTION

N many cold climate regions of the world, substations


have demonstrated poor reliability with repeated electrical
flashovers on insulators covered with ice or snow. Singly or in
combinations, pollution, ice or snow accretion on insulators
may cause a drastic decrease in electrical insulation strength
that can lead to flashover and power outages at normal service
voltage. These outages involved both station and line insulators
[1][25]. In the USA, there are sixteen separate references to ice
problems causing major events [26]. The preliminary results of
a survey [27] conducted by CIGRE WG B2-03 (formerly WG
22-03) showed that ice and snow affected the electrical function
of line insulators about five times more than their mechanical
function. Mechanical failure, leading to insulator separation, is
a dominant concern for line insulators in the CIGRE results.
However, the electrical consequences of station flashovers,
including high fault currents and equipment stresses, are more
severe than for faults on line insulators.
An IEEE Task Force position paper [28] recently examined a
wide range of laboratory test methods and procedures for insulator icing. Common features from IEC and IEEE standards for
mechanical icing and high voltage testing were brought together
into a recommended test procedure.
The main purpose of these TF position papers (Parts I and
II) is to recommend criteria for selecting and/or improving the
electrical performance of station insulators in cold, hostile climates that experience ice and snow conditions. The focus on station insulators, including bus supports and apparatus housings,
Manuscript received March 1, 2004. Paper no. TPWRD-00110-2004.
The authors are with the IEEE Task Force on Insulator Icing Methods.
Digital Object Identifier 10.1109/TPWRD.2004.839731

is essentially based on two reasons. First, more than half of electrical icing flashover problems in North America are associated
with station insulators [2], [3], [26], [29]. Second, the flashover
risk is very sensitive to pollution exposure [28]. It is desirable
to develop experience with the modeling at specific locations,
where distances to nearby pollution sources are known, such as
substations, before attempting the more difficult problem of integrating all pollution sources along the length of an overhead
line.
These recommendations, dealing with all voltage levels, will
be helpful for selecting insulators in new installations, for insulator replacement programs, and for treatment options where
existing insulators must be upgraded. Parameters for insulator
selection include insulator type, profile and configuration, dry
arcing and leakage distances.
The first objective is to provide sufficient climate data and
environmental exposure to establish the risk factors at a fixed
location, such as a substation.
The second objective is to describe the relation between climate exposure and risk of flashover. This risk changes in discrete steps depending on the level of icicle bridging of shed
spacing, but it changes smoothly with increasing ice deposit
thickness and water conductivity.
The main focus of this first part of the paper is on the physics
and chemistry of important aspects of insulation flashover under
icing conditions.
II. SCIENTIFIC BASIS AND TECHNICAL CONTEXT
A. Contaminated Ice and Snow Deposits
Several ice parameters, including type and density, amount,
and distribution, and the eventual electrical conductivity of
water on the ice surface, have a significant influence on the
flashover performance of iced insulators [28].
1) Type and Density of Ice: Laboratory tests and field experience both show that the most dangerous type of ice is glaze
with icicles, having a density of about 0.87 g/cm [4], [30], [31].
In the case of glaze, the water film on the ice surface during the
accretion and melting periods creates the most critical flashover
condition.
In the case of low-density rime, water can penetrate the ice
during the melting period and the surface of the insulator may

0885-8977/$20.00 2005 IEEE

FARZANEH et al.: SELECTION OF STATION INSULATORS WITH RESPECT TO ICE AND SNOWPART I

Fig. 1.

265

Density function of melted-ice conductivity [15].

be wettable. This also may present a critical flashover condition


in cases of heavily polluted snow or pre-contaminated insulator
surfaces.
2) Amount and Distribution of Ice: The amount of ice has
a significant influence on flashover performance [4], [28], [30],
[31]. Normally, a thickness of 530 mm, measured on a standard
rod [28], is recommended for ice testing. The thickness should
be decided by utilities depending on their service experience in
exposure [55] and desired reliability [28].
The distribution of ice along the insulators is another important parameter. For typical precipitation intensities and duration,
ice and icicles cover the windward side of the insulator and the
opposite side is free of ice. The level of ice bridging between the
sheds determines the size and number of air gaps where initial
discharges and partial arcs are triggered. This level of bridging is
mainly influenced by the electric field distribution along the insulator, which in turn is governed by the level of applied voltage,
the presence of grading rings and grading devices on apparatus
insulators (e.g., capacitors), bus bars, support structures and insulator orientation.
3) Electrical Conductivity of Water: Ions in cloud droplets,
rain, and wet snow may stem from sea salt, nitrates, sulphates
and various industrial pollutants [15].
As concerns insulator icing tests, three types of conductivity
can be identified: (i) applied water conductivity (also called
freezing water conductivity), (ii) dripping water conductivity,
and (iii) melted ice conductivity. The conductivity of applied
water is always measured and usually reported. However, it is the
conductivity of the water film on the ice surface, best correlated
with dripping water conductivity that mainly determines the
flashover voltage. This is similar to the insulator surface conductivity in the case of flashover mechanisms of polluted insulators.
The Norwegian Power Grid Company (STATNETT) provided information about the conductivity of 122 ice samples
taken from service, between 1987 to 1995 [16]. The probability
density function of the conductivity of these melted-ice samples
is shown in Fig. 1.
These samples, mostly of rime ice, were taken from racks and
towers close to the ground, which thus corresponds to applied
water conductivity. The samples are assumed to be more contaminated than snow on the ground, because of the higher content of rime. The results were statistically evaluated and it was
found that a log-normal distribution, with mean value of 3.2 (av-

Fig. 2. Relationship between ESDD and distance from seacoast.


TABLE I
POLLUTION LEVELS AND RECOMMENDED SPECIFIC LEAKAGE DISTANCE
ACCORDING TO IEC 60 815 AND ESTIMATED ESDD IN-SERVICE LEVELS

erage conductivity 25.4 S/cm) and standard deviation of 0.8


(6.4 S/cm) gave the best fit. The log-normal distribution for
conductivity can be approximated by (1), as follows:
(1)

B. Clean Ice Deposit on Contaminated Surfaces


Fig. 2 shows monthly rates of increase of contamination
levels as a function of distance from seacoast in Japan [61];
superimposed are three increase rates in ESDD observed for
urban, suburban, and rural areas in Ontario.
IEC 60 815 [56], currently under revision, has defined four
pre-existing pollution levels and specific leakage distance requirements for insulators, as shown in Table I. As well, the
IEEE TF has added a recommended range of pollution levels,
expressed as ESDD values, to Table I.
The pre-existing ESDD will dominate the electrical conductivity of the ice for any (generally thin) layer of ice, freezing
fog or drizzle. Only the surfaces in direct contact with ice accretion, typically top surfaces and vertical portions of insulators,
will contribute to the conductivity of ice. The contribution of the
ESDD to the ice conductivity can be calculated using [57]
ESDD Area
Volume

(2)

Area is in cm , ESDD in g/cm , volume in ml, conductivity


in S/cm.

266

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 1, JANUARY 2005

C. Combined Pollution: Ice and Snow on Contaminated


Surface
One of the most difficult winter conditions for insulators is
associated with build-up of contamination over a period of a
month, as shown in Fig. 2, combined with a moderate amount
of ice or wet-snow accretion. Accretion onto a pre-contaminated surface will draw in the surface ions, enhancing or dominating the natural conductivity of the ice. Through the process
of freeze-thaw partitioning, these ions will migrate to the ice
surface, lower the surface impedance and reduce the flashover
voltage further as described in Part II of the paper.
An estimate of the increase in ice conductivity is given by
(2), using the ESDD on the top surface of the sheds, the surface
area and volume of the ice cap. For complex situations, where
different surfaces have different ESDD levels, it is suggested
that the ice conductivity and weight be measured directly or
inferred from the dripping water properties.

Fig. 3.

Annual number of hours with freezing rain.

Fig. 4.

Annual number of hours of freezing drizzle.

Fig. 5.
[33].

Number of hours per year with fog or freezing fog with Ta

D. Methods for Evaluating the Electrical Performance of


Insulators and Ranking Under Icing Conditions
Laboratory test methods should simulate a wide range of ice
accretion types and regimes. Two common accretion methods
are (i) by freezing precipitation, where generally icicles are
formed by run-off of supercooled water; and (ii) through consolidation of snow or rime in melting conditions. In both cases,
a water film may appear during melting caused by a rise in air
temperature above 0 C or by solar input.
It is generally agreed that the simultaneous presence of a
highly conductive water film and initial air gaps are needed
for flashover to occur [32]. Thus, for the evaluation tests, the
experimental procedure should reproduce these conditions
as closely as possible. The procedure for icing test methods
including preparation, exposure, quantification, and evaluation
features have been reported in detail in the previous position
paper [28]. The two following methods for evaluating and
comparing the electrical performance of ice-covered insulators
have already been recommended in that work [28].
i) Compare the icing performance of the insulators for a
given test severity.
ii) Compare the icing performance of different insulators at
the maximum possible ice severity (e.g., maximum possible level of bridging), giving the worst case icing test
performance for each insulator.
The approaches for the evaluation and ranking of insulator
performance following this outline have already been proposed
and discussed in [28].
III. CLIMATE AND ENVIRONMENTAL CONTEXT
A. Climate
The risks of flashover on station insulators are correlated with
the occurrence, in general, of freezing rain, drizzle or wet snow.
Hence, meteorological statistics may provide relevant information for the estimation of risk levels. However, it is also important to note that the pollution level of such precipitation may
vary significantly, as a result of both long-range [15] and short-

<0

range transport of contamination. It is also important to be aware


of enhanced salt content in wet snow, stemming from rough sea
and heavy production of sea spray in the upwind area, prior
to the deposition. Two seemingly equal wet snow or freezing
rain events may therefore have different impacts on electrical
equipment.
Figs. 35 show periodic observations of the annual number of
hours with freezing rain [33], freezing drizzle [58], and freezing
fog, respectively. This data can be used further for the estimation
of a number of icing events.

FARZANEH et al.: SELECTION OF STATION INSULATORS WITH RESPECT TO ICE AND SNOWPART I

B. Environment
Insulator pollution standards, such as IEC 60 815 [56],
currently under revision, provide some guidance in Table I regarding the leakage distance needed for reliable long-term use
as a function of insulator pollution level. The present IEEE TF
recommends the use of four pollution levels in Table I as ranges
of equivalent salt deposit density (ESDD). A revised standard
IEC 60 815 will also take into account another important
pollution parameter, i.e., nonsoluble deposit density (NSDD).
Equivalent advice is given below for icing and cold-fog conditions. However, there is little guidance for establishing the
anticipated ESDD level, except for test values near the sea [13].
Winter exposure is somewhat similar to desert exposure in
the sense that the top surfaces are not exposed to rain for extended periods of weeks or months, depending on the local ambient temperature. Short-term exposure of several weeks tends
to have a linear relationship between ESDD and duration, while
the ESDD levels off for long-term exposure of 6 to 18 months.
Under winter conditions, top-surface ESDD values increase to
levels higher than bottom surfaces. The rate of increase of ESDD
with time must be established from the local pollution environment, and used in conjunction with climate records for the median or extreme duration of dry periods.
A model for pollution exposure at substations can be derived from the mass flux of nearby point, line, or area pollution
sources. An example of a point source would be a chimney located several kilometers upwind of the station. Road salting on
metric tons of NaCl per lane km is
expressways at a level of
a common line source, and generating stations located near the
sea are exposed to a half-plane source whenever the wind blows
in from the ocean. The source strength of the half-plane formed
by the sea is strongly influenced by wind speeds and wave action [13].
The wind rosette, a polar plot of wind speed and direction
over the exposure period, is used in common surface deposition models. The probability of being directly downwind from
a point source is fairly low, so the exposures from line and area
sources usually dominate the calculation. When such models are
applied on a sea surface area, it is important to consider the differences in surface winds and the wind direction at 5001000 m
levels, where the main transportation of the pollution, as well
as the release of precipitation, takes place. Due to the veering
of wind with height, the local surface wind direction may be as
much as 90 (anticlockwise on the northern hemisphere) off the
transport direction of sea salt aloft, depending on the roughness
of the terrain.

267

ditions that lead to slow melting, (3) and (4) have approximated
the insulation strength [60]:
kV m dry arc
kV m dry arc

for ice
for snow.

(3)
(4)

ISP is the product of the ice layer weight (g/cm of dry arcing
distance) and the ice layer conductivity ( S/cm). Reference [8]
also notes a nonlinear relationship for EHV insulators.
B. Diameter
North American experience during heavy ice storms suggests
that station insulators of greater diameter are more prone to
ice flashover than neighboring post insulators of smaller diameter [34]. The relationship between ice accretion diameter and
flashover strength for a fixed shed profile was recently studied
[35] and can be expressed as:
V kV m dry arc

(5)

where W is the ice width (525 cm).


A rough relation, as recently established during the experiment at the CIGELE laboratory (UQAC), can be made between
the ice thickness on the reference cylinder and the weight of accreted ice per meter of 300-mm uniform shed profile station post
insulators, as follows:
Weight

Thickness

(6)

During this experiment the applied-water conductivity was


80 S/cm and the maximum ice thickness was 20 mm on a
rotating monitoring cylinder.
C. Shed Spacing
Typically, station post insulators differ from line insulators of
the same voltage class in their shed spacing. Cap-and-pin line
insulators with 146-mm construction height have a spacing of
about 100 mm from the bottom of one skirt to the top surface
of the next insulator. Sheds on typical ceramic post station insulators are about 50 mm apart. It is obvious that, with the same
ice exposure and similar diameter, ice bridging will occur much
more rapidly on station insulators. Once the insulators are fully
bridged, the accumulation rates on station and line insulators are
almost the same, and their electrical performance tends to converge. This means stations will have more problems than lines,
especially in areas where there are several moderate ice storms
per year.
D. Shed Profile

IV. EFFECTS OF ATMOSPHERIC ICE AND POLLUTION WITH


RESPECT TO INSULATOR PARAMETERS
A. Exposed Surface to Icing
The surface exposed to icing is normally a product of insulator dry arcing distance and diameter, with some adjustment for
different shed profiles. Whenever the leakage distance is fully
bridged with ice or snow, the insulation strength is established
mainly by the electrical properties of the ice or snow, and the exposed surface area. Under the most severe meteorological con-

Additional leakage distance is provided on insulators by, for


example, forming bottom parts of the sheds with deep underribs. The leakage distance thus obtained is 2 or 3 times longer
than the dry arcing distance. This parameter improves pollution
performance for a wide range of pollution conditions when no
icing occurs.
Many different shed profiles can be obtained for the same
dry-arcing distance. Under conditions of heavy icing, deeper
sheds tend to pack more ice and snow, before bridging, leading
to a greater ice weight (g/cm of dry arcing distance) for the same

268

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 1, JANUARY 2005

Fig. 6. Comparison of Clean-Fog and Cold-Fog flashover strength as a


function of ESDD.

icing exposure, but this effect is less important than insulator diameter. Alternating diameters of sheds and multi-cone profiles
can improve ice and snow flashover performance in some conditions by delaying the onset of ice bridging, compared to designs with uniform shed profile [1], [18], [21], [24], [36], [42].
Booster sheds of various designs have also been evaluated for
this function.
Tests on a wide range of shed profiles are reported for 24-kV
class insulators [10]. These results suggest that the influence of
shed profile on flashover voltage is relatively limited compared
to the parameters of the ice deposit.
E. Leakage Distance
Leakage distance and the so-called protected leakage distance play an important role in pollution flashover strength for
very light levels of freezing rain, drizzle, and fog accumulation,
but are less important when ice or snow bridging of insulator
shed spacing occurs. In the worst case, the effective leakage distance is reduced to the dry arcing distance.
Generally, there is a reasonable agreement between the results
of cold-fog tests versus clean-fog tests, according to IEC 60507,
with cold-fog giving strengths that are 1020% lower than clean
fog results [40]. Empirically the electrical flashover strength of
leakage distance in cold-fog conditions, as a function of ESDD,
is given by (7).
V kV/m leakage

ESDD

(7)

ESDD in g/cm .
Fig. 6 shows the strength predicted by (7), superimposed on
a summary of conventional clean-fog strength, replotted from
data in [40] using a rough conversion factor of 2.5x for the relation between leakage and arcing distance.
F. Orientation
Testing experience for iced station insulators has generally
been for vertical orientations, since these are most common.
Results for V-strings of suspension insulators [37] suggest an
improvement in strength for oblique angles, compared to vertical. Operational and testing experience with wet flashover of

Fig. 7. Icing test on a T-shaped circuit-breaker at (a) open position and


(b) closed position.

wall bushings [41] suggest that horizontal orientation may perform in unexpected ways compared to vertical post insulators.
Ice or snow accretion on the top of the horizontal insulator will
bridge the dry arcing distance at a lower precipitation amount.
However, overall ice thickness will not build up as fast because
icicles will develop vertically downward, not across the shed
spacings, see example in Fig. 7.
The placement of two station insulators in close proximity
will tend to degrade the performance, compared to the single
insulator. This is true for both heavy snow and ice accretion and
was previously noted for line insulators [38].
A special question for orientation concerns the ice performance of T-shaped circuit-breakers, or disconnecting circuitbreakers. To simulate the service case, it was decided to perform
ice accretion on both chambers (horizontal part) and support insulator (vertical part). The Ice Progressive Stress Method (IPS)
was used for testing [20], [21]. The test results of the breaker
shown in Fig. 7 demonstrated a drastic difference in ice distribution at the breaker in open and closed operating positions (see
examples). This also resulted in different flashover voltages in
open and closed positions.
G. Surface Material and Finish
The ice accretion onset and rate tend to be relatively unaffected by insulator surface materials, whether ceramic or nonceramic. However, the electrical performance of nonceramic
or silicone-coated ceramic insulators is somewhat better, especially under conditions of partial ice bridging [2], [39]. Another surface that improves performance under icing conditions
is a semi-conductive glaze with a surface resistivity of about
1030 M [63]. This surface suppresses the onset of arcing
[39], [42][46], [59] and is more effective under heavy ice conditions than silicone.

FARZANEH et al.: SELECTION OF STATION INSULATORS WITH RESPECT TO ICE AND SNOWPART I

The surface finish of the insulator affects performance in


an indirect way. Test results show that ice slides off more
easily from smooth, new ceramic surfaces than from aged
porcelain, silicone-coated porcelain or nonceramic sheds of the
same shape. This shortens the time that melting ice is on the
insulators and proportionally reduces risk of flashover.
V. CONCLUSION
This first part of the paper has provided the physical and
chemical background needed to describe the electrical flashover
on iced insulators. Several ice and snow parameters, including
type and density, amount and distribution, water conductivity,
as well as certain insulator parameters, such as diameter, shed
spacing and profile, insulator orientation, and exposed surface
to ice, influence flashover mechanisms and performance.
The most important aspect of icing performance is specifically related to the partial or complete bridging of insulator
leakage distance by icicles extending from shed to shed. Under
fully bridged conditions, the insulator leakage distance is reduced nearly to the dry arcing distance. This drastic reduction
in leakage distance (by a factor of about three) can lead to severe but sporadic ice-induced flashovers at a number of affected
utilities.
Flashover can also occur along the leakage distance of the station insulator under conditions of light or moderate icing combined with pre-deposited pollution. For many cold regions, the
maximum levels of insulator contamination (ESDD) are reached
in the winter. The exposure can be aggravated from the effects
of high winds and road salting. An additional margin of 1020%
over the existing design recommendations for specific leakage
distance for adequate clean-fog performance is needed for sufficient cold-fog performance on vertical ceramic post insulators.
Part II of the paper [62] applies the environmental parameters,
as described in Part I, to the selection process and mitigation
options.
ACKNOWLEDGMENT
The Chairman would like to thank Dr. W. A. Chisholm especially for his enthusiasm, dedication, and contributions, to Tom
Grisham for helping with IEEE administration and to Dr. Igor
Gutman for links to the IEC 60 815 revision process.
REFERENCES
[1] E. A. Cherney, Flashover performance of artificially contaminated and
iced long-rod transmission line insulators, IEEE Trans. Power App.
Syst., vol. PAS-99, no. 1, pp. 4652, Jan./Feb. 1980.
[2] W. A. Chisholm et al., The cold-fog test, IEEE Trans. Power Del., vol.
11, no. 4, pp. 18741880, Oct. 1996.
[3] M. Farzaneh and O. T. Melo, Flashover performance of insulators in
the presence of short icicles, Int. J. Offshore Polar Eng., vol. 4, no. 2,
pp. 112118, 1994.
[4] M. Farzaneh and J. Kiernicki, Flashover problems caused by ice
build-up on insulators, IEEE Electr. Insul. Mag., vol. 11, no. 2, pp.
517, Mar./Apr. 1995.
[5] J. F. Drapeau and M. Farzaneh, Ice accumulation characteristics on
Hydro-Quebec HV insulators, in Proc. 6th Int. Workshop on Atmospheric Icing of Structures, Budapest, Hungary, 1993, pp. 225230.
[6] L. Gu et al., AC flashover characteristics of EHV line insulators for
high altitude contamination regions, in Proc. ICPAPM, 1988.

269

[7] L. Shu, C. Sun, J. Zhang, and L. Gu, AC flashover performance on


iced and polluted insulators for high altitude regions, in Proc. 7th Int.
Symp. High Voltage Engineering, vol. 4, Dresden, Germany, 1991, pp.
303306. Paper 43.13.
[8] V. Sklenicka and J. Vokalek, Insulators in icing conditions: Selection
and measures for reliability increasing, in Proc. 7th Int. Workshop
Atmospheric Icing of Structures, Chicoutimi, QC, Canada, 1996, pp.
7276.
[9] J. S. Forest, The performance of high voltage insulators in polluted
atmospheres, in Proc. Conf. Paper IEEE Winter Meeting, New York,
1969.
[10] K. Kannus, K. Verkonnen, and V. Lavkervi, Effects of ice coating on the
AC performance of medium-voltage insulators, in Proc. Nordic Symp.
Elect. Insulation, NORD-IS 86, Esbo, Finland, 1986, pp. 111.
[11] H. Matsuda, H. Komuro, and K. Takasu, Withstand voltage characteristics of insulator strings covered with snow or ice, IEEE Trans. Power
Del., vol. 6, no. 3, pp. 12431250, Jul. 1991.
[12] T. Fujimura, K. Naito, Y. Hasegawa, and K. Kawaguchi, Performance
of insulators covered with snow or ice, IEEE Trans. Power App. Syst.,
vol. PAS-98, no. 5, pp. 16211631, Sep./Oct. 1979.
[13] R. Matsuoka, S. Ito, K. Sakanishi, and K. Naito, Flashover on contaminated insulators with different diameters, IEEE Trans. Dielectr. Electr.
Insul., vol. 26, no. 6, pp. 11401146, Dec. 1991.
[14] M. Yasui, K. Naito, and Y. Hasegawa, AC withstand voltage characteristics of insulator string covered with snow, IEEE Trans. Power Del.,
vol. 3, no. 2, pp. 828838, Apr. 1988.
[15] S. M. Fikke, J. E. Hanssen, and L. Rolfseng, Long range transported
pollution and conductivity on atmospheric ice on insulators, IEEE
Trans. Power Del., vol. 8, no. 3, pp. 13111321, Jul. 1993.
[16] S. M. Fikke, T. M. Ohnstad, T. Telstad, H. Frster, and L. Rolfseng,
Effect of Long Range Airborne Pollution on Outdoor Insulation, 1994.
NORD-IS 94, Paper 1.6.
[17] A. Meier and W. M. Niggli, The influence of snow and ice deposits on
supertension transmission line insulator strings with special reference to
high altitude operations, in Proc. IEEE Conf. Publ. 44, London, U.K.,
1968, pp. 386395.
[18] D. Wu, K. A. Halsan, and S. M. Fikke, Artificial ice tests for long insulator strings, in Proce. 7th Int. Workshop Atmospheric Icing of Structures, Chicoutimi, QC, Canada, 1996, pp. 6771.
[19] D. Wu and R. Hartings, Correlation between the AC withstand voltage
and insulator lengths under icing tests, in Proc. 9th Int. Workshop Atmospheric Icing of Structures, Chester, U.K., 2000, p. 6.
[20] I. Gutman, S. Fikke, and K. Halsan, Development of the ice progressive test method applicable for the full-scale testing of the 420 kV class
overhead line insulators, in IWAIS, Brno, Czech Republic, Jun. 2002.
61.
[21] I. Gutman, K. Halsan, and D. Hbinette, Application of ice progressive
stress method for selection of different insulation options, in Proc. 13th
Int. Symp. High-Voltage Engineering (ISH), Delft, Holland, Aug. 2003.
[22] M. Kawai, AC flashover test at project UHV on ice-coated insulators,
IEEE Trans. Power App. Syst., vol. PAS-89, no. 8, pp. 18001804,
Nov./Dec. 1970.
[23] M. D. Charneski, G. L. Gaibrois, and B. F. Whitney, Flashover tests on
artificially iced insulators, IEEE Trans. Power App. Syst., vol. PAS-101,
no. 8, pp. 24292433, Aug. 1982.
[24] H. M. Schneider, Artificial ice tests on transmission line insulatorsA
progress report, in Proc. IEEE/Power Eng. Soc. Summer Meeting, San
Francisco, CA, 1975, pp. 347353. Paper A75-491-1.
[25] Z. Vuckovic and Z. Zdravkivic, Effect of polluted snow and ice accretion on high voltage transmission line insulators, in Proc. 5th Int. Workshop Atmospheric Icing Structures, Tokyo, Japan, 1990. Paper B4-3.
[26] Repercussions of ice and snow on the flashover performance of outdoor
insulators, CIGRE Working Group 33 Colloq. Group 4, Jul. 1997.
[27] Reply to the Questionnaire on the Mechanical & Electrical Failures of
Insulators Caused by Ice and Snow, 2000. CIGRE WG B2-03, document
B2-02 (WG03) IWD-190.
[28] M. Farzaneh et al., Insulator icing test methods and procedures, IEEE
Trans. Power Del., vol. 18, no. 4, pp. 15031515, Oct. 2003.
[29] J. F. Drapeau, M. Farzaneh, M. Roy, R. Chaarani, and J. Zhang, An
experimental study of flashover performance of various post insulators
under icing conditions, in Proc. Conf. Electrical Insulation Dielectric
Phenomena (CEIDP), vol. 1, Victoria, BC, Canada, 2000, pp. 359364.
[30] M. Farzaneh and J. F. Drapeau, AC flashover performance of insulators
covered with artificial ice, IEEE Trans. Power Del., vol. 10, no. 2, pp.
10381051, Apr. 1995.

270

[31] M. Farzaneh and J. Kiernicki, Flashover performance of IEEE standard


insulators under ice conditions, IEEE Trans. Power Del., vol. 12, no. 4,
pp. 16021613, Oct. 1997.
[32] M. Farzaneh, Ice accretion on high-voltage conductors and insulators
and related phenomena, Philos. Trans. Roy. Soc., vol. 358, no. 1776,
pp. 29713005, Nov. 2000.
[33] J. V. Cortinas Jr., C. C. Robbins, B. C. Bernstein, and J. W. Strapp, A
climatology of freezing rain, freezing drizzle and ice pellets across North
America, in Proc. 9th Conf. Aviation, Range, Aerospace Meterology,
Orlando, FL, 2000, AMS, pp. 292297.
[34] J. F. Drapeau, R. Beauchemin, J. Laflamme, R. Martin, and M. J.
Roy, Tenue sous verglas de lisolation externeRapport #1: Risques
de contournements en conditions de verglasd et impacts des pertes
dquipements associs, in hhTenue sous verglasii (Confidential):
Hydro-Qubec Working Group, Nov. 1996.
[35] R. Chaarani, tude de linfluence des caractristiques des isolateurs sur
leurs performances lectriques dans des conditions de givrage, Ph.D.
dissertation, Univ. Quebec Chicoutimi, Apr. 2003.
[36] C. C. Erven, 500 kV insulator flashovers at normal operating voltage,
in Proc. Presentation to the Can. Elect. Assoc. (CEA) Spring Meeting,
Montreal, QC, Canada, 1988.
[37] K. Kannus, K. Verkonnen, and E. Lakervi, Effect of ice coating on the
dielectric strength of high voltage insulators, in Proc. 4th Int. Workshop
Atmospheric Icing of Structures, Paris, France, 1988, pp. 296300.
[38] M. M. Khalifa and R. M. Morris, Performance of line insulators
under rime ice, IEEE Trans. Power App. Syst., vol. PAS-86, no. 6, pp.
692698, Jun. 1968.
[39] J. S. Barrett and M. A. Green, A statistical method for evaluating electrical failures, IEEE Trans. Power Del., vol. 9, no. 3, pp. 15241530,
Jul. 1994.
[40] R. S. Gorur, E. A. Cherney, and J. T. Burnham, Outdoor insulators,
Ravi S. Gorur Inc., 1999.
[41] P. J. Lambeth, Y. Beausjour, and S. I. Kamel, Behavior of HVDC wall
bushings under nonuniform rain, Rep., Report for the Canadian Electrical Association CEA 203 T 787, Jul. 1996.
[42] M. Farzaneh, J. F. Drapeau, S. Brettschneider, and M. Roy, tude comparative de performance lectrique des isolateurs externes dans des conditions de glace atmosphrique en vue du choix adquat disolateurs de
postes 735 kV, in Colloquium on Icing, ACFAS Conf. Sherbrooke, QC,
2001, http://icevolt.uqac.uquebec.ca/cigele.
[43] V. Jaiswal, M. Farzaneh, and D. A. Lowther, Impulse flashover performance of semi-conducting glazed insulators equipped with booster
sheds under icing conditions, in Proc. Canadian Conf. Electrical Computer Eng., Montreal, QC, Canada, May 2003.
[44] V. Jaiswal and M. Farzaneh, Effects of semi-conducting glaze coating
on the electrical performance of a H.V. station post insulator covered
with ice, in Proc. 13th Int. Symp. High-Voltage Engineering, Delft, Holland, Aug. 2003.
[45] J.-F. Drapeau, M. Farzaneh, and M. Roy, An exploratory study of various solutions for improving ice flashover performance of station post
insulators, in Proc. Int. Workshop Atmospheric Icing Structures, Brno,
Czech Republic, Jun. 2002.
[46] W. A. Chisholm, Effects of Flue-Gas Contamination of Ceramic Insulator Performance in Freezing Conditions,, EPRI Rep. TR-110 296,
1998.

IEEE TRANSACTIONS ON POWER DELIVERY, VOL. 20, NO. 1, JANUARY 2005

[47] M. Farzaneh, C. Volat, and A. Gakwaya, Electric field calculation


around ice-covered insulator using boundary element method, in Proc.
IEEE Int. Symp. Electrical Insulation, Anaheim, CA, Apr. 2000, pp.
349355.
[48] V. Sklenicka et al., Influence of conductive ice on electric strength of
HV insulators, in Proc. Int. Symp. Pollution Performance of Insulators
and Surge Diverters, vol. 1, Madras, Tamilnadu, India, 1983. paper 1.02.
[49] M. Farzaneh, J. Zhang, and X. Chen, Modeling of the AC arc discharge
on ice surfaces, IEEE Trans. Power Del., vol. 12, no. 1, pp. 325338,
Jan. 1997.
[50] M. Farzaneh, I. Fofana, C. Tavakoli, and X. Chen, Dynamic modeling
of DC arc discharge on ice surfaces, IEEE Trans. Dielectr. Electr. Insul.,
vol. 10, no. 3, pp. 463474, Jun. 2003.
[51] J. Zhang and M. Farzaneh, Computation of AC critical flashover
voltage of insulators covered with ice, in Proc. Int. Conf. Power System
Technology, vol. 1, Beijing, China, Aug. 1998, pp. 524528.
[52] M. Farzaneh, Y. Li, J. Zhang, L. Shu, X. Jiang, W. Sima, and C. Sun,
Electrical performance of ice-covered insulators at high altitudes,
IEEE Trans. Dielect. EIectr. Insul., vol. 11, no. 5, pp. 870880, Oct.
2004.
[53] Influence of Ice and Snow on the Flashover Performance of Outdoor
InsulatorsPart I: Effects of Ice, 1999. CIGRE TF 33 04 09, Electra no.
187.
[54] Influence of Ice and Snow on the Flashover Performance of Outdoor
InsulatorsPart II: Effects of Snow, 2000. CIGRE TF 33 04 09, Electra
no. 188.
[55] IEEE/ANSI Std. C2.
[56] Guide for Selection of Insulators in Polluted Areas. IEC 60 815.
[57] W. A. Chisholm, P. G. Buchan, and T. Jarv, Accurate measurement of
low insulator contamination levels, IEEE Trans. Power Del., vol. 9, no.
3, pp. 15521557, Jul. 1994.
[58] C. C. Robbins and J. V. Cortinas Jr., Local and synoptic environments
associated with freezing rain in the contiguous united states, AMS
Weather Forecasting, pp. 4765, Feb. 2003.
[59] M. Farzaneh, J.-F. Dapeau, J. Zang, M. J. Roy, and J. Farzaneh,
Flashover performance of transmission class insulators under icing
conditions, in Proc. Insulator News Market Rep. Conf., Marbela,
Spain, Aug. 2003, pp. 315326.
[60] W. A. Chisholm, J. Kuffel, and M. Farzaneh, The icing stress product:
A measure for testing and design of outdoor insulators in freezing conditions, in Proc. 2nd Int. Workshop on High Voltage Engineering, Tottori,
Japan, 2000. session 8.
[61] Technical Guide, NGK Nagoya, Tech. Guide, Cat. No. 91-R, 1991.
Second Edition, p. 107.
[62] IEEE Task Force On Icing Performance of Station Insulators et al.,
Selection of station insulators with respect to ice and snowPart II:
Methods of selection and options for mitigation, IEEE Trans. Power
Del., vol. 20, no. 1, pp. 271277, Jan. 2005.
[63] Y. Suzuki, S. Seike, and O. Imai, A practical study of semiconducting
glaze for insulators, Elect. Eng. Jpn., vol. 131, no. 1, pp. 1018, 2000.

Anda mungkin juga menyukai