Вы находитесь на странице: 1из 10

Scientific papers

Effects of Long Term Corona


and Humidity Exposure
of Silicone Rubber Based
Housing Materials
Maoqiang Bi1, 4, Stanislaw M. Gubanski1, Henrik Hillborg2, Jens M. Seifert3, and Bin Ma1

High Voltage Engineering, Chalmers University of Technology


SE-412 96, Gteborg, Sweden
2
Power Technologies, ABB Corporate Research
SE-721 78, Vsters, Sweden
3
LAPP Insulators GmbH
D-95632, Wunsiedel, Germany
4
on leave from State Key Laboratory of Power Transmission Equipment & System Security and New
Technology Chongqing University, 400044, Chongqing, China
1

Introduction
In recent decades market share of composite insulators used for outdoor high voltage
insulation is increasing steadily. This can be
attributed to their advantageous properties
over the traditional ceramic and glass insulators, among which low weight, higher mechanical strength to weight ratio, resistance
to vandalism and better performance in the
presence of heavy pollution belong to the
most important ones [1].
In a search for appropriate test methods and
minimum material property requirements,
CIGRE
working
group
WGD1.14
concentrated its past work on defining the
physical parameters important for the use
of polymeric materials in outdoor insulation
and on checking if relevant test methods
are available today. Twelve properties have
been identified [2], whereas standardized
test methods and minimum requirements
have been available for eight of them. Since
2004, the work of WG D1.14 had therefore
focused on developing new test methods
and on setting minimum requirements
for the remaining four properties, among
which the resistance to corona and ozone
was listed as being of great importance.
In the continuation of this work a test
procedure had been developed to treat

4 No. 267 - April 2013

ELECTRA

and compare behaviour of various housing


materials exposed to corona discharges
under dry conditions[3-5]. It was found that
a 100-hours long corona treatment yielded
significant changes of nearly all the materials
tested, especially their surface resistivity. The
treatment also influenced the mechanical
properties, including breakdown strength
and elongation at break. The observed
changes could mainly be attributed to
surface oxidation reactions.
In service there are basically two possible
sources of corona discharges in the vicinity of
the insulator surface: improperly designed or
degraded insulator hardware (including corona rings) and the presence of water droplets.
The first source acts locally and promotes
ageing of insulator housing even at dry conditions. This can be avoided using proper design of the field grading devices [6]. However,
the unavoidable presence of water droplets
on insulator surface acts as a discrete and distributed source of corona that yields ageing at
different insulator parts [7, 8].
The deterioration mechanism of housing
surface under corona discharge is complicated.
Material surface is simultaneously subjected
to a mixture of energetic and reactive species
as well as radiation. For example, the most
important ones in case of silicone-based

Experimental procedure
The treatment was performed at ambient
temperature and pressure in a sealed glass
chamber containing the corona electrode
system, previously used by CIGRE WG D1.14

Scientific paperS

polymers include an increase of the oxygen


content at the surface by formation of silanol
and hydroxyl groups [9], oxidative crosslinking
[10], and degradation of the polymer
network structure resulting in the formation
of low molecular mass compounds [6]. Also
degradation by nitric acid has been claimed
as one of the possible mechanisms [11, 12],
though there is only little evidence that this
mechanism really operates. Therefore there
exists a need to elucidate this question
further. The studies described in this paper
are therefore concentrated on two aspects,
the first one referred to defining the degree of
material damage after a long term exposure
to corona discharges in dry condition and
storage in a moist environment whereas the
other one attempts to detect the formation
of acidic species during these conditions.
Test sheets of two types of silicone rubber
housing materials were tested: a HTV
rubber filled with alumina trihydrate (ATH,
Al(OH)3) and reinforcing silica, and a LSR
containing reinforcing silica only. During the
whole treatment cycle the samples were
mechanically pre-stressed (by ~3%). Ozone
concentration and power released from
the corona electrode during exposure were
monitored. The effects of the treatment were
evaluated by means of different standard
surface characterization techniques, such
as contact angle measurements, X-ray
photoelectron spectroscopy (XPS), reflection
infrared spectroscopy (ATR-FTIR), Scanning
electron microscopy (SEM) and atomic force
microscopy (AFM). Measurements of surface
and volume resistivity as well as mechanical
properties were also performed.

[3, 13], on sheets of the tested materials in


size of 120 100 2.15 mm3 (HTV) and 120
100 1.75 mm3 (LSR). Corona discharges
were generated from needle electrodes
energized at 20 kVrms in a dry air (< 2% RH)
and separated from the treated surfaces by
40 mm. The diameter of the treated surface
region under the electrode was ~90 mm.
The air flow rate was maintained constant
at 5 L/min for securing a stable level of
ozone concentration in the chamber, which
was monitored by means of Kitagawa gas
detector tubes [14]. The detection limits of
these tubes were as follows: ozone 1 ppm
(tube No. 182 SB), NO 1 ppm (tube No.
174 A), NO2 - 0.01 ppm (tube No. 117 SD)
and HNO3 - 0.5 ppm (tube No. 233 S). A
Schematic view of the test arrangement is
illustrated in Figure 1.

Figure 1: Schematic view of corona treatment


arrangement

After a continuous corona exposure of


100 hours, 400 ml of deionized water was
injected into the chamber. Consequently, the
relative humidity in the chamber increased to
above 90% and was maintained at this level
for the next 65 hours. Thereafter the chamber was flushed with dry air for 2 hours and
the corona treatment started again. Such
cycles were repeated ten times, cumulatively
resulting in 1000 hours of corona exposure
and 650 hours of storage in the humid air.
The whole procedure is in detail illustrated in
Figure 2.

Figure 2: One cycle of long term corona and humidity exposure of housing materials

No. 267 - April 2013

ELECTRA

Scientific paperS

results and discussion


3.1 Corona discharges
A typical corona discharge activity from the
multi-needle electrode arrangement is illustrated in Figure 3.The power release from the
electrode, estimated by integrating the measured discharge current [15], was constant at
1.4 W during the whole experiment.

Figure 4: Ozone concentration during the corona


treatment of the tested materials. Each test round
consists of 100 h continuous corona exposure.

3.2 Dielectric properties

Figure 3: Corona exposure of a sheet of silicone rubber,


where the corona light emission can be clearly seen.

The concentration of ozone produced by


the discharges showed a decreasing tendency as presented in Figure 4. The detected
amount of ozone was significantly lower for
the HTV rubber, compared with LSR. The
reason for this effect still needs to be elucidated. At the same time no evidence for the
formation of NO, NO2 and HNO3 could be
detected.

The measurements of surface and volume


charging resistivity were performed on the
same samples before and after the 5th round
of the treatment by corona and humidity. The
results are summarized in Table 1, where the
arrows in the columns effect of treatment
indicate the varying trends in logarithmic
scale as compared to the initial reference values. For the surface resistivity the decrease
was exceeding four orders of magnitude,
whereas no significant change in volume resistivity was observed.
As regards the dielectric response the
HTV rubber, both the real and imaginary
parts of permittivity became affected as

Table 1: Values of surface and volume resistivity of the tested materials before and after 5 rounds of the corona and
humidity treatment (calculated for the time of measurement equal to 4 104 s).

Figure 5: Real (a) and imaginary (b) parts of relative permittivity of HTV rubber

6 No. 267 - April 2013

ELECTRA

Scientific paperS

Figure 6: Real (a) and imaginary (b) parts of relative permittivity of LSR rubber.

compared with the initial values (see Figure


5). The real part of permittivity increased significantly at low frequency (10-4 1 Hz), while
it remained unchanged at higher frequencies
(above 10 Hz). The imaginary part increased
mainly at frequencies above 10-3 Hz. In case
of LSR (Figure 6), the real and imaginary parts
of permittivity decreased at the low frequencies, but at the same time, a slight increase
could be noticed in the imaginary part at
higher frequencies.

3.3 Mechanical properties


One of the main requirements set
for housing materials is retaining their
mechanical integrity, which after exposure

to corona and humidity treatment can be


assessed by the change of tensile strength
and elongation at break. Figure 7 presents
the results of such investigations on the
tested materials after completion of
the ten cycles. The tensile strength and
elongation of break of the stiffer HTV
rubber was not significantly influenced,
even though a trend towards lower values
can be appreciated. The tensile strength
and elongation of break of the more
flexible LSR material was significantly
reduced after the test, indicating a certain
degradation of the material. It should
however be noted that the elongation at
break after the test is still > 250%.

Figure 7: Tensile strength (a) and elongation at break (b) of the investigated
materials before and after corona and humidity treatment

8 No. 267 - April 2013

ELECTRA

Scientific paperS

Figure 8: Water contact angles at various stages of corona and humidity treatment for HTV (a) and LSR (b) (note: the
test round 0 indicates reference measurements before the treatment).

3.4 Surface hydrophobicity


The performed investigations included
measurements of both the advancing (a) and
the receding (r) water contact angles after
each treatment cycle, (Figure 8). The initial
hydrophobicity of the HTV rubber was higher,
compared to the LSR, exhibiting higher
advancing and receding angles. The hydrophobicity was measured within one hour
after each completed test round of 100 h
corona and 65 hours humidity exposure. The
hydrophobicity of the HTV rubber gradually
decreases during the three first test rounds.
After the subsequent seven test rounds the
hydrophobicity gradually recovered. This behavior was even stronger for the RTV rubber,
which showed a minimum in hydrophobicity
after the first test round, followed later by a
hydrophobicity similar to or even higher as
compared to the initial values. Especially the
advancing water contact angle increased,
while the receding angle remained similar to
the initial level. High values in contact angle
hysteresis (i.e. difference between advancing

and receding angles) are typical for surfaces


which exhibit heterogeneous domains with
different surface energy and/or high surface
roughness. While the contact hysteresis of
the HTV rubber was relatively similar after
each test round, the LSR exhibited a gradual
increase after the first test round. The latter
indicates a large change in the surface structure or the chemical composition of LSR.

3.5 X-ray photoelectron spectroscopy


(XPS)
The atomic compositions of the top 20 nm
surface of the rubbers before and after the
completed test are summarized in Table 2.
It should be noted that the samples were
extracted in hexane prior to XPS analysis
in order to remove migrating silicone oils,
which otherwise will contaminate the surface
during the analysis. Theoretically, the composition of a pure polydimethylsiloxane (PDMS),
the base polymer in silicone rubber, is 50%
of carbon (C), 25% of oxygen (O) and 25% of
silicon (Si). As can be seen, the atomic

Table 2: Atomic composition of the surface layers according to XPS

No. 267 - April 2013

ELECTRA

Scientific paperS

Table 3: Analysis of peak area of Si 2p (%)

compositions of the two materials in their initial state exhibit a similar composition. The
presence of aluminum was not observed. The
fillers, which have a higher surface energy
compared with the polymer, are covered by
a thin layer of silicone rubber which is thicker
than the probing depth of the instrument.
.However, clear differences between the
reference state and the state after the treatment could be observed. The carbon content
of the HTV rubber and LSR were reduced
by 35.2% and 54.2%, respectively. Instead
the oxygen content increased by 57.0% and
86.8%. At the same time the content of silicon increased by 12.1% and 19.2%, respectively. These results show that the degree
of surface oxidation was higher for the LSR,
compared to the HTV rubber. This can further be linked to the higher ozone concentration measured during the corona treatment
of the LSR.
Furthermore, high resolution XPS scans
of the Si2p orbital peak were performed to
investigate the changes in the chemical state
of the surfaces. Figure 9 shows the behavior
of Si2p orbital peak for the HTV rubber. A
broadening and strengthening of the peak

with binding energy at ~104.0 eV, associated with formation of an inorganic silica-like
phase (SiOx), is observed; while the peak at
102.6 eV, associated with organic silicone
phase, gets weaker. An even higher conversion into a silica-like phase was found for the
LSR (Table 3). It should be noticed that the
higher conversion into a silica-like surface
layer for the LSR corresponds well to the
higher oxygen content, shown in Table 2.

3.6 Infrared analyses (FTIR)


The infrared spectra of both the tested
materials before and after the treatment are
shown in Figure 10. The infrared analysis is
less surface sensitive compared to the XPS,
probing a depth up to 4 m into the rubber.
The following main chemical bonds could
be identified in all of them: CH bond in CH3
groups at 2960 cm1, SiCH3 bond (side chain)
at 1260 cm-1 and SiOSi bond (silicone backbones) at 1054 cm-1. For the HTV rubber,
OH in ATH was identified at 3200-3700 cm1
region as well. For the LSR, a broad absorption peak at 3200-3700 cm-1 was found after
the exposure, which indicates formation of
hydroxyl (OH) groups [16]. A similar peak is

Figure 9: XPS high resolution Si2p spectra of HTV rubber: initial state (a) and after treatment (b). The Si2p peak is
resolved into the organic silicone phase at 102.6 eV peak and an inorganic silica-like phase at ~104.0 eV.

10 No. 267 - April 2013

ELECTRA

expected to occur also for the HTV rubber,


but it is most likely obscured by the hydroxylgroups bound within the ATH filler. This can be
seen indirectly by increased peak heights of
the ATH filler after the treatment. For both the
material types, the absorption peak of SiOSi
became broader and slightly higher after the
exposure, as can be seen in the partial enlarged insert in Figure 11. This broadening is a
typical signature of oxidative crosslinking of
silicone rubber polymer matrix [13]. It is also
important to note that no evidence for the
presence of formed nitric acid, characterized
by an absorption peak at 1380 cm-1 [12], on the
surface of the treated materials was found.

3.7 Surface inspection by scanning


electron microscopy (SEM)
Changes of color appearance were found
on the surfaces of both types of the mate-

Figure 11: SEM images of surfaces of HTV rubber (a, c)


and LSR (b, d) at different magnification after long-term
corona and humidity exposure.

rials after the treatment, especially on the


LSR. Here a network of small cracks was seen
already after the first treatment round. No
such cracks were seen by the naked eye on
the HTV rubber, even after 10 rounds of the
treatment. To further investigate the surface
change, SEM investigation were conducted
after the test was completed, and examples
of results are shown in Figure 11. A weak
crack pattern was present on the treated
surfaces of the HTV rubber, as seen Figure
11(a). Under the same magnification, extensive cracks were found on LSR (Figure 11(b)).
The width of the cracks were typically ~1 m
(marked with arrows) and 5-8 m for the HTV
rubber and LSR, respectively (Figures 11(cd)). Thus a brittle surface layer was formed
on both the materials. The question was then
how thick these layers became after a total
exposure of 1000 h corona and 650 h

Scientific paperS

Figure 10: IR spectra of HTV (a) and LSR (b) rubbers.

Figure 12: Cross sections of freeze-fractured specimens of HTV


rubber (a, c) and LSR (b, d).

No. 267 - April 2013

ELECTRA

11

Scientific paperS

Figure 13: Atomic force micrographs of 5 5 m2 scans before and after


corona and humidity treatment: (a) initial and (b) treated surfaces of HTV
rubber, (c) initial and (d) treated surfaces of LSR, (e) crack in the silica-like
layer on a treated HTV rubber surface. Note the difference scales in height.

storage in moist conditions.


The thickness of the formed silica-like
layers can be determined using cross-sections of freeze-fractured specimens of the
treated materials (Figure 12). The thickness
was generally less than one micrometer, in
spite of the extensive corona treatment. The
thickness was lower for the HTV rubber, in
the order of 0.3 0.4 m; whereas the thickness was around 1-1.2 m for the LSR. The
silica-layer on top of the LSR appeared more
brittle, partially delaminating during the
freeze-fracture (Figure 12(d)). In addition the
ATH filler particles could be readily observed
in the bulk of the HTV rubber, whereas only
smaller aggregates of the silica-filler with
lower particle size was seen in the bulk of the
LSR.
Fracturing of the silica-like layer may either
be caused by internal tension or by external
mechanical stresses and it accelerates the
hydrophobicity recovery due to easier migration of low molar mass species through the
surface cracks. The appearance of extensive
surface cracking of the LSR after the first
test round correlate well with the enhanced
hydrophobicity after each of the subsequent

12 No. 267 - April 2013

ELECTRA

test rounds. It is therefore believed that the


fracture properties of silica-like layers largely
control the rate of hydrophobic recovery,
which may explain the observed differences
in the behavior of both the studied materials
[16] (comp. Figure 8).

3.8 Atomic force microcopy (AFM)


The structure of the silica-like layer, in
between the cracks, was further characterized on a sub-micrometer scale using atomic
force microscopy. The typical appearance of
the surface topography, before and after the
test, is shown in Figure 13. The virgin materials exhibit a relatively rough surface structure, reflecting both imprints from the molding tool during manufacturing, as well as
the presence of filler particles (covered with a
thin layer silicone rubber), (Figure 13 (a & c)).
The surface roughness of the silica-like layer
formed after exposure is lower, and similar
for both materials. It can also be seen that
the layers are heterogenous on the nanometer scale, where sub-micrometer sized silicarich domains are surrounded by less oxidized
silicone [17, 18]. Also larger filler particles
(0.5 1 m) can be found, as seen. The

Scientific paperS

formation of silica-like layers on silicones can


be utilized for electronic, analytical as well as
medical applications [19]. Figure 13(e) shows
a typical crack in the silica like layer on the
HTV silicone rubber. The values of the surface roughness, expressed both in root mean
square and arithmetic mean values, are summarized in Table 3. It can be found that the
treatment yielded, in general, a reduction of
the surface roughness, which was also observed earlier [20]. Although, the difference in
the roughness of the surfaces of both materials before the treatment was obvious, it
became afterwards similar to each other.

Table 3: Surface roughness of (uncracked) silica like layer


determined by atomic force microscopy
Typical standard deviation = 20%; Rq - root mean square
roughness, Ra - arithmetic mean roughness

At the same time the mechanical properties


were not affected significantly, less for the
HTV material. At the same time, the hydrophobic properties, after an initial decrease,
recovered with the duration of the treatment.
These changes could be coupled to the formation and fracturing of oxidized and brittle
silica-like surface layers. The latter was more
intensive on the LSR material. The penetration depth of this process remained however
limited to a depth of ~1 m, indicating that a
long term exposure of silicone rubber based
materials to subsequent corona and humidity
cycles does not create strong risks as regards
materials mechanical and electrical integrity.
The investigations aiming at detecting
gases formed during the corona treatment as
well as the later performed infrared analyses
of the treated surfaces did not reveal any
detectable evidences of acidic contaminations, both in the discharge chamber as well
on the material surfaces. This may indicate
that acidic attack on the housing materials is
not likely to operate at conditions similar to
those described in this paper.

conclusions

acknowledgment

A long term treatment by repetitive corona


discharge treatment combined with storage
in high humidity (> 90% RH) at room temperature was performed in well controlled laboratory conditions on two types of silicone
rubber based housing materials, typically
used for the manufacturing of composite insulators: a HTV rubber filled with alumina trihydrate (ATH, Al(OH)3,) and reinforcing silica,
and a LSR containing reinforcing silica only.
The treatment yielded a strong reduction
of surface resistivity of both the materials,
though still remaining within the limits set
for safe functioning in outdoor applications.

Financial support from Lapp Insulators


GmbH is gratefully acknowledged. Besides,
Xiangdong Xu, a Ph.D. student at the Division of High Voltage Engineering at Chalmers
University of Technology is acknowledged
for support in performing dielectric measurements. Also China Scholarship Council (CSC) is
acknowledged for providing financial support
to one of the authors (Maoqiang Bi) for performing the experimental work at the Division of
High Voltage Engineering of Chalmers University of Technology. G. Mattsson, ABB Corporate Research, Sweden, is acknowledged for
performing the SEM analyses.

Reference
[1] R. Hackam, Outdoor HV composite polymeric insulators,
Dielectrics and Electrical Insulation, IEEE Transactions on,
vol. 6, pp. 557-585, 1999.
[2] CIGRE WG D1.14, Mterial properties for non-ceramic
outdoor insulation: State of the art, CIGRE Techinical
Brochure No. 255, 2004.
[3] M. Bin, J. Andersson, and S. M. Gubanski, Evaluating
resistance of polymeric materials for outdoor applications
to corona and ozone, Dielectrics and Electrical Insulation,
IEEE Transactions on, vol. 17, pp. 555-565, 2010.
[4] J. Aderson, B. Sonerud, Y. Serdyuk, S. M. Gubanski, and

14 No. 267 - April 2013

ELECTRA

H. Hillborg, Evaluation of polymeric materials resistance


to corona and ozone, presented at the 15th International
Sympoium on High Voltage Engineering (ISH), Ljubljana,
Slovenia, 2007.
[5] S. M. Gubanski, M. Bin, J. Aderson, and C. Leuci,
Evaluating resistance of materials for outdoor insulator
housings to corona and ozone, presented at the INMR,
2009.
[6] T. G. Gustavsson, S. M. Gubanski, H. Hillborg, S. Karlsson,
and U. W. Gedde, Aging of silicone rubber under ac or dc
voltages in a coastal environment, Ieee Transactions on

[14] KITAGAWA Gas Detector Tube System Handbook. Tokyo


Janpan: Komyo Rikagaku Kogyo K. K., 2008, p.^pp. Pages.
[15] M. Bin, Effects of Corona and Ozone Exposure on
Properties of Polymeric Materials for High Voltage
Outdoor Applications, Ph. D, Department of Materials
and Manufacturing Technology, High Voltage Engineering,
Chalmers University of Technology, Gothenburg, 2011.
[16] H. Hillborg, M. Sandelin, and U. W. Gedde, Hydrophobic
recovery of polydimethylsiloxane after exposure to partial
discharges as a function of crosslink density, Polymer, vol.
42, pp. 7349-7362, 2001.
[17] J. Song, J. F. L. Duval, M. A. Cohen Stuart, H. Hillborg,
U. Gunst, H. F. Arlinghaus, and G. J. Vancso, Surface
Ionization State and Nanoscale Chemical Composition of
UV-Irradiated Poly(dimethylsiloxane) Probed by Chemical
Force Microscopy, Force Titration, and Electrokinetic
Measurements, Langmuir, vol. 23, pp. 5430-5438,
2007/05/01 2007.
[18] H. Hillborg, N. Tomczak, H. Schnherr, and G. J. Vancso,
Converting Poly (dimethylsiloxane) Surfaces to Silica-like
layers: A Surface Study of UV/Ozone Treated Sylgard
184, Langmuir, vol. 20, pp. 785-794, 2004.
[19] M. Owen and P. Dvornic, Eds., Silicone Surface Science.
Verlag: Springer, 2012, p.^pp. Pages.
[20]
H. Hillborg, J. F. Ankner, U. W. Gedde, G. D.
Smith, H. K. Yasuda, and K. Wikstrm, Crosslinked
polydimethylsiloxane exposed to oxygen plasma studied
by neutron reflectometry and other surface specific
techniques, Polymer, vol. 41, pp. 6851-6863, 2000.

When Moisture is a Problem,


Knowledge is Power.
Literally.
Protect Your High-Voltage Assets.
How to revive an ageing eet of high-voltage assets?
Can you extend network lifetime and ensure performance?
Just add knowledge online, real-time moisture level information from
transformers, switchgear, turbines, generators and diesel engines.
Its easy, affordable and fast.
Plug in our on-line transmitters and in just minutes you are able to
prevent rapid deterioration caused by moisture, boost performance,
optimize maintenance and get extra years of safe peak performance.

Did you know? Multiparameter transmitter DPT145 monitors online


SF6 gas pressure, density, temperature and moisture.
Providing solutions and services to customers in more than 120 countries. Thousands of installations in the
power industry. Service centers in Finland, China, Japan and USA. Over 75 years of measurement expertise.

www.vaisala.com/power
sales@vaisala.com

Meet us at CIRED, 10-13 June, Stockholm.


Stand E 09

Scientific papers

Dielectrics and Electrical Insulation, vol. 8, pp. 1029-1039,


Dec 2001.
[7] A. J. Phillips, D. J. Childs, and H. M. Schneider, Aging of
nonceramic insulators due to corona from water drops,
IEEE Transactions on Powe r Delivery, vol. 14, pp. 1081-9,
1999.
[8] J. P. Reynders, I. R. Jandrell, and S. M. Reynders, Surface
ageing mechanisms and their relationship to service
performance of silicone rubber insulation, London, UK,
1999, pp. 54-8.
[9] J. R. Hollahan and G. L. Carlson, Appl Pol Sci, vol. 14, p.
2499, 1970.
[10] C. H. Hillborg, Loss and recovery of hydrophobicity
of polydimethyl-siloxane after exposure to electrical
discharges, PhD degree PhD thesis, Dept. of Polymer
Technology, Royal Institute of Technology, Stockholm
Sweden, 2001.
[11] Y. Koshino, I. Umeda, and M. Ishiwari, Deterioration of
silicone rubber for polymer insulators by corona discharge
and effect of fillers, in Electrical Insulation and Dielectric
Phenomena, 1998. Annual Report. Conference on, 1998,
pp. 72-79 vol. 1.
[12] I. Umeda, K. Tanaka, T. Kondo, K. Kondo, and Y. Suzuki,
Acid aging of silicone rubber housing for polymer
insulators, in Electrical Insulating Materials, 2008. (ISEIM
2008). International Symposium on, 2008, pp. 518-521.
[13] M. Bin, S. M. Gubanski, and H. Hillborg, AC and DC
zone-induced ageing of HTV silicone rubber, Dielectrics
and Electrical Insulation, IEEE Transactions on, vol. 18, pp.
1984-1994, 2011.

Вам также может понравиться