Sei sulla pagina 1di 8

F E A T U R E A R T I C L E

Preparation of a Vegetable Oil-Based


Nanofluid and Investigation of Its
Breakdown and Dielectric Properties
Key words: vegetable insulating oil, nanofluid, breakdown voltage, dielectric properties

Introduction Jian Li, Zhaotao Zhang, and Ping Zou


Investigations during the last decade have shown that conduc-
tive nanoparticles can be dispersed in transformer oils to form
State Key Laboratory of Transmission Equip-
nanofluids. Well-dispersed nanoparticles are capable of increas- ment & System Security and New Technolo-
ing the breakdown voltage of the oil under power frequency gy, Chongqing University, Chongqing, China
and lightning impulses. They also increase the inception volt-
ages for partial discharge [1]. With increasing moisture content,
reduction of the breakdown voltage of the nanofluid at power
Stanislaw Grzybowski
frequency is significantly smaller than that in the corresponding Department of Electrical and Computer En-
transformer oil [1]. The electrical and thermal properties of four gineering, Mississippi State University, Mis-
types of nanofluid, prepared by dispersing Al2O3, Fe3O4, SiO2,
and SiC nanoparticles in transformer oils, were described in [2]. sissippi State, Mississippi
It has also been reported that the thermal conductivity of such
oil was enhanced by 8% when aluminum nitride nanoparticles Markus Zahn
were dispersed in it at a loading of 0.5% by weight, and its cool- Department of Electrical Engineering and
ing capability was improved by about 20% [3]. An electrody-
namic model has been developed describing streamer formation Computer Science, Massachusetts Institute of
in transformer oil-based nanofluids, which presents generation, Technology, Cambridge, Massachusetts
recombination, and transport equations for each charge carrier
type [4].
Vegetable insulation oils are based on natural ester oils,
which are environmentally friendly and fire resistant [4]–[9]. At An insulating nanofluid was pre-
the moment, little is known about the preparation of nanofluids
using natural ester oils and their dielectric, breakdown, and ag- pared by dispersing surface-modified
ing properties. Fe3O4 nanoparticles in a vegetable
Surface modification of nanoparticles is a very effective pro-
cedure to avoid nanoparticle agglomeration in insulating nano- insulation oil. The breakdown and
fluids [10]–[14]. However, the surface modification procedures dielectric properties of the nanofluid
used for mineral oils cannot be applied to vegetable oils because are compared with those of the oil.
of their very different molecular structures. We therefore inves-
tigated new approaches to the preparation of vegetable oil-based
nanofluids.
This paper presents some of the results of a study of the
breakdown voltages and dielectric properties of a vegetable Nanofluid Preparation
oil-based nanofluid. The nanofluid was prepared by dispersing The vegetable insulation oil (named RDB) was obtained from
Fe3O4 nanoparticles in a vegetable insulation oil obtained from raw rapeseed oil by using three procedures, namely, alkaline
a laboratory at Chongqing University. Oleic acid was used for refinement, vacuum distillation, and bleaching [15]–[18]. The
surface modification of the nanoparticles. basic physical, chemical, and electrical properties of RDB oil

September/October — Vol. 28, No. 5 0883-7554/12/$31/©2012/IEEE 43


are shown in Table 1. The moisture contents of the measured was subjected to ultrasonic dispersion and mechanical
oil samples were below 60 mg/kg. The corresponding properties agitation for 2 h.
of two commercial vegetable oils, FR3 [6] and BIOTEMP [7], 2. The nanoparticles were separated from the ethanol–
manufactured by Cooper Power Systems and ABB, respectively, oleic acid solution by centrifugation, washed three
are also presented in Table 1. times in ethanol, and then dried under vacuum at 60°C
Table 1 also shows the testing standards appropriate to the for 10 h. The dried nanoparticles were ground into a
various measurements. The properties of the RDB oil were mea- powder (of oleic acid-coated Fe3O4 nanoparticles).
sured by an authorized testing institute in China, using Chinese 3. The RDB oil samples were dried under a pressure of 50
standards identical to the corresponding International Electro- Pa for 72 h at 85°C. The oleic acid-coated nanoparticles
technical Commission (IEC) and ISO (International Organiza- were then added and the mixture was ultrasonically ag-
tion for Standardization) standards. In accordance with the IEC itated for 20 minutes at 30°C to disperse the nanopar-
60156 standard [19], a pair of spherical electrodes with 2.5-mm ticles through the oil. The weight ratio of nanoparticles
spacing was used for measurements of breakdown voltages at 50 to oil was 0.004%.
Hz on the RDB samples. (As reported in [6] and [7], the break-
down voltages of FR3 and BIOTEMP at 60 Hz were measured
using a pair of spherically capped electrodes with 2-mm spac- Characterization
ing, in accordance with [20]). In this work the dissipation factor,
relative permittivity, and volume resistivity of the RDB samples Nanoparticle Sizes
were measured at 90°C, as required by IEC 60247. The ASTM A laser particle-size analyzer was used to measure the size
(American Society for Testing and Materials) standards adopted distributions of the nanoparticles before and after surface modi-
(by other authors) for FR3 and BIOTEMP specify 25 and 100°C. fication (Figure 1). Agglomerated nanoparticles with sizes in the
Analytical grade oleic acid and ethanol, and Fe3O4 nanoparti- range 0.2 to 2 μm were removed by the surface modification
cles with an average diameter of 30 nm were used to prepare the procedure [21], which also increased the nanoparticle sizes (be-
RDB-based nanofluid. The nanoparticles were purchased from cause their surfaces were coated with oleic acid molecules).
Nanjing Emperor Nano Material Co. Ltd. (Nanjing, China). The
nanofluid was prepared in three steps as follows: Transmission Electron Microscopy
The morphologies of the nanoparticles before and after sur-
1. Ultrasonic dispersion and mechanical agitation were face modification were observed by transmission electron mi-
used to disperse 5 g of nanoparticles in 100 mL of etha- croscopy, as shown in Figure 2. Before being placed in the mi-
nol at 60°C. After 5 minutes, the oleic acid was added croscope, the nanoparticles were washed in alcohol, carefully
to the nanoparticle–ethanol solution at a ratio of 0.25 g spread over fine copper grids, and dried at room temperature.
of oleic acid/100 mL of ethanol. The resulting mixture Agglomeration of the unmodified nanoparticles is clearly shown

Table 1. Basic Physical, Chemical, and Electrical Properties of Three Types of Vegetable Oil.1

RDB FR3 BIOTEMP

Property Value Test method Value Test method Value Test method

Appearance Light yellow IEC 61099 Light green ASTM D1524 Clear & bright ASTM D1524

Density (kg·m−3) 0.90/20°C ISO 3675 0.92/25°C ASTM D1298 0.91/25°C ASTM D1298

Kinematic viscosity (mm2·s−1) 43/40°C ISO 3104 34/40°C ASTM D445 45/40°C ASTM D445

Pour point (°C) −18 ISO 3016 −21 ASTM D97 −15 to −25 ASTM D97

Flash point (°C) 325 ISO 2592 316 ASTM D92 330 ASTM D92

Acid value (mgKOH·g−1) 0.03 ISO 660 0.04 ASTM D974 0.075 ASTM D974

Interfacial tension (mN·m )−1


30 ISO 6295 24 ASTM D971 — ASTM D971

Breakdown voltage (kV) 73 IEC 60156 56 ASTM D1816 65 ASTM D1816

Dissipation factor (%) 2/90°C IEC 60247 3/100°C ASTM D924 2/100°C ASTM D924

Relative permittivity 2.9/90°C IEC 60247 3.2/25°C ASTM D1169 3.2/25°C ASTM D1169

Volume resistivity (Ω·m) 1 × 1010/90°C IEC 60247 2 × 1011/25°C ASTM D924 1 × 1011/25°C ASTM D924

IEC = International Electrotechnical Commission; ASTM = American Society for Testing and Materials; ISO = International Organization for Standardization.
1

44 IEEE Electrical Insulation Magazine


Figure 1. Nanoparticle size distribution (a) before and (b) after
surface modification.

Figure 3. Attenuated total reflection-Fourier transform infrared


in Figure 2a. The nanoparticles maintained their nearly spheri-
spectra of (a) unmodified and (b) surface-modified nanopar-
cal shapes after surface modification (Figure 2b), and their sizes
ticles.
were consistent with those shown in Figure 1.

Attenuated Total Reflection-Fourier Transform


Infrared Spectroscopy bonding between the oleic acid molecules and the nanoparticle
Separate samples of the unmodified and surface-modified surfaces.
nanoparticle powders were thoroughly milled with iodine bro- Thermogravimetric and Differential
mide. The resulting mixtures were pressed into thin films, on
which attenuated total reflection-Fourier transform infrared
Thermal Analyses
Thermogravimetric analysis (TGA) and differential thermal
spectroscopy measurements were made. The absorption peaks
analysis (DTA) measurements were made on the nanoparticles
of the unmodified nanoparticles (3420 and 1634 cm−1; Figure
before and after surface modification. Before measurement, the
3a) originate in the hydroxyl (−OH) groups on the surfaces of
nanoparticles were washed in ethanol and dried in vacuum at
the unmodified nanoparticles. The two strong absorption peaks
60°C for 10 h. During measurement, they were heated from 20
at 630 and 584 cm−1 originate within the nanoparticles [22], [23].
to 900°C at a rate of 20°C/min in a nitrogen atmosphere.
The surface-modified nanoparticles show other absorption
Figure 4 shows the TGA and DTA curves for the unmodified
peaks at 2922 and 2852 cm−1 (Figure 3b), originating in –CH2
nanoparticles. The peaks in the DTA curve correspond to the
groups of the oleic acid used for surface modification [24]. The
mass losses in the TGA curve. Peaks 1 and 2 in the DTA curve
peaks at 1469 and 1404 cm−1 originate in COO– groups derived
result from the loss of water and hydroxyl groups, respectively,
from carboxyl (–COOH) groups [25] and reveal the chemical
from the nanoparticle surfaces; they represent about 2 and 3.5%,
respectively, of the total nanoparticle mass. Peak 3 near 850°C
corresponds to the phase transition from Fe3O4 to FeO within the
nanoparticles.
The DTA curve for the nanoparticles after surface modifica-
tion (Figure 5) shows four peaks. Peaks 1 and 4 are due to the
removal of water and the phase transition from Fe3O4 to FeO,
respectively. Peak 2 corresponds to a mass loss of about 3%;
the fact that it appears around the boiling point of oleic acid
suggests that some of the oleic acid molecules are physically ad-
sorbed on the nanoparticle surfaces and escape (volatilize) when
the temperature approaches the boiling point. Peak 3 also corre-
sponds to a 3% mass loss; the fact that it appears around the de-
composition temperature of oleic acid suggests that some of the
oleic acid molecules are chemically bonded to the nanoparticle
surfaces and escape when the temperature approaches the de-
composition point. The fact that the peak 2 and 3 mass losses are
approximately equal suggests that approximately equal masses
Figure 2. Transmission electron microscopy of nanoparticles of oleic acid are physically adsorbed on, and chemically bonded
(a) before and (b) after surface modification. to, the nanoparticle surfaces. Because chemical bonding forces

September/October — Vol. 28, No. 5 45


Figure 4. Thermogravimetric analysis (TGA) and differential
thermal analysis (DTA) curves for the unmodified nanopar-
ticles.

are stronger than van der Waals forces associated with physical
adsorption, chemical bonding probably contributes more to the
stability of the nanoparticle dispersion in the oil than does physi-
cal adsorption.

Dispersion Stability of Nanoparticles in Oil


The dispersion stability of the nanoparticles in the nanofluid
was determined using the natural sedimentation method. The
unmodified and surface-modified nanoparticles were dispersed
in the oil samples, at nanoparticle-to-oil weight ratios of 1%,
to form nanofluid samples 1 and 2, respectively. To disperse the
nanoparticles in the oil, both samples were ultrasonically agitat-
Figure 6. Unmodified (1) and surface-modified (2) nanopar-
ed for 20 minutes at 30°C, i.e., the same procedure as described
ticles dispersed in oil.
in the Nanofluid Preparation section above, except for a much
larger nanoparticle/oil weight ratio (1%), chosen to expedite the
natural sedimentation observations. Sample 1 was allowed to
stand for 1 night and sample 2 was allowed to stand for 30 days,
both at 30°C. In Figure 6, sample 1 shows two clearly separated
volumes; the darker volume at the bottom of the container con-
sisted mainly of agglomerated nanoparticles. In contrast, sample
2 shows no visible nanoparticle–oil separation.

Breakdown Voltages
The AC breakdown voltages of the oil and the corresponding
nanofluid (prepared in the manner described in the Nanofluid
Preparation section) were measured in accordance with standard
IEC 60156 [19]. The absolute moisture content of the oil was
288 mg/kg. Measurements were made on five samples of the oil
and on five samples of the nanofluid. The results are shown in
Table 2. The average breakdown voltage of the nanofluid sam-
ples is approximately 20% greater than that of the oil samples
[21]. This increase may be due to the trapping of free electrons
by the polarized nanoparticles, and thus the prevention of fur-
ther streamer development [4]. More detailed explanations are
Figure 5. Thermogravimetric analysis (TGA) and differen- presented below.
tial thermal analysis (DTA) curves for the surface-modified The lightning impulse breakdown voltages of the oil and the
nanoparticles. nanofluid were measured using the container and the electrode

46 IEEE Electrical Insulation Magazine


Table 2. Breakdown Voltage of Oil and Nanofluid Samples at 50 Hz. Table 3. Positive Lightning Impulse Breakdown Voltages.

Breakdown voltage (kV) Breakdown voltage (kV)

Sample 1 2 3 4 5 Average Sample 1 2 3 4 5 Average

Oil 48.6 51.5 45.3 56.2 48.0 49.9 Oil 77.8 73.4 74.5 70.2 73.4 73.9

Nanofluid 58.3 54.7 63.1 62.5 60.4 59.8 Nanofluid 104.8 99.4 99.4 102.6 101.5 101.5

configuration shown in Figure 7. The high-voltage electrode was oil samples. The results in Tables 3 and 4 show a previously
a steel needle, the grounding electrode was a 13-mm-diameter unreported phenomenon, namely, that the average positive light-
steel ball, and the distance between the tip of the needle and the ning impulse breakdown voltage of the nanofluids is greater than
ball was 15 mm. These dimensions comply with IEC 60897 [26] the corresponding value for negative impulses.
for liquid dielectrics. Standard lightning impulse voltages with The measured times to lightning impulse breakdown of the
both negative and positive polarities were applied to each oil and oil and nanofluid samples are shown in Tables 5 and 6. The aver-
each nanofluid sample. age times to positive and negative lightning impulse breakdown
Table 3 shows the measured positive lightning impulse break- in the nanofluids are, respectively, 21 and 14% greater than those
down voltages. The average breakdown voltage of the nanofluid in the oil alone. The average streamer velocity can be calculated
samples is approximately 37% greater than that of the oil sam- using the average time to breakdown and the electrode gap. The
ples. values for the nanofluids are 1.25 and 1.18 km/s for positive and
Table 4 shows the measured negative lightning impulse negative lightning impulse voltages, respectively, and 1.51 and
breakdown voltages. The average breakdown voltage of the 1.35 km/s for the oil alone.
nanofluid samples is approximately 12% greater than that of the The charge relaxation time constant of nanoparticles in a
nanofluid has a major impact on electrodynamic processes oc-
curring in the nanofluid, i.e., if the relaxation time constant is
short relative to the time scales of streamer growth, the nanopar-
ticles will significantly modify the electrodynamics [4]. The
charge relaxation time constant τ for the oil–nanoparticle system
is given by τ = (2ε1 + ε2)/(2σ1 + σ2) [4], where ε1 and ε2 are the
permittivities of the oil and the nanoparticles, respectively, and
σ1 and σ2 are their conductivities. Substituting ε1 = 2.68 × 10−11
F/m, σ1 = 6.29 × 10−11 S/m, ε2 = 7.08 × 10−10 F/m, and σ2 = 1 × 104
S/m [4], we obtain τ = 7.62 × 10−14 s, which is extremely short
compared with the microsecond time scale involved in streamer
propagation. It follows that Fe3O4 nanoparticles will dramati-
cally affect the electrodynamics of streamer development.
Negatively charged nanoparticles are formed in an external
field when the polarized nanoparticles trap free electrons [4].
Free electrons travel quickly toward the anode, but the negative-
ly charged nanoparticles tend to remain in the ionization zone
[4] because of their much smaller mobilities. Negatively charged
nanoparticles generate a spatial electrical field En, which is su-
perimposed on the external field E and reduces the field strength
near the cathode. Consequently, the rate of injection of electrons
into the nanofluid at the cathode will fall. The low mobility of
negatively charged nanoparticles hinders the development of
space charge at the tip of a streamer [4], which in turn alters

Table 4. Negative Lightning Impulse Breakdown Voltages.

Breakdown voltage (kV)

Sample 1 2 3 4 5 Average

Figure 7. Electrode configuration and oil container for light- Oil 85.3 84.2 83.2 82.1 84.2 83.8
ning impulse breakdown voltage measurements on oil and
Nanofluid 96.1 94.0 92.9 96.1 89.6 93.7
nanofluid samples.

September/October — Vol. 28, No. 5 47


Table 5. Time to Breakdown of Oil and Nanofluid Samples at Positive Light-
ning Impulse Voltages.

Time to breakdown (μs)

Sample 1 2 3 4 5 Average

Oil 10.0 12.1 8.8 9.0 9.7 9.9

Nanofluid 10.0 11.3 13.2 10.3 15.2 12.0

the electrodynamics involved in the development of the electric


field wave. The latter is the dominant mechanism in streamer
propagation, leading to electrical breakdown [4] in the nanoflu-
id. Consequently, electrical breakdown in oil-based nanofluids
occurs at higher voltages than that in oil without nanoparticles.
The results presented above show that the negative lightning
Figure 8. Volume resistivity versus frequency for oil and nano-
impulse breakdown voltage of the nanofluid exceeds that of the
fluid.
host oil. This result is contrary to that observed in mineral oil-
based nanofluids [1]. The explanation may be that the viscosity
of vegetable oil is greater than that of mineral oil, and that the
viscosity of a liquid dielectric may influence the development of Figure 10 shows the dissipation factor versus frequency for
streamers in that liquid. the oil and the nanofluid. The difference between the two is very
small at frequencies greater than approximately 1 Hz. Below 1
Hz, the values begin to diverge, with that for the nanofluid be-
Dielectric Properties ing smaller and the difference exceeding 1% at 10−2 Hz. There
The volume resistivities of the oil and the nanofluid were
are two main loss mechanisms in oil under an AC electric field,
measured using a Concept 80 broadband dielectric spectrometer
namely, conductance loss and polarization loss. Because RDB
(Novocontrol Technologies GmbH & Co. KG, Hundsan-
oil is a weakly polar liquid dielectric, the conductance loss is
gen, Germany). Figure 8 shows volume resistivity as a function
dominant at low frequency [27]. The volume resistivity of the
of frequency. Electrically neutral nanoparticles dispersed in an
oil-based nanofluid is greater than that of the oil at low frequen-
insulating fluid are polarized when the fluid is subjected to an
cy, so the dissipation factor of the nanofluid is smaller than that
externally applied electric field, and in that state tend to trap free
of the oil at low frequency. At frequencies above 1 Hz, the vol-
electrons [4]. Thus the concentration of highly mobile electrons
ume resistivities of the oil and nanofluid are very similar (Figure
in the nanofluid declines and the concentration of low-mobil-
9), so the differences in dissipation factors also become very
ity negatively charged nanoparticles increases. As a result, the
small.
electrical conductivity of the nanofluid is smaller than that of
the RDB oil. With increasing frequency, the duration of each
half cycle of the electrical field decreases and the probability of
electron capture by the nanoparticles decreases. This may be the
reason for the smaller difference in volume resistivity between
the RDB oil and the nanofluid at frequencies greater than ap-
proximately 0.3 Hz.
Figure 9 shows relative permittivity versus frequency for oil
and nanofluid samples. The relative permittivity of the nanopar-
ticles is 80 [4], much greater than that of the oil. Thus polariza-
tion of the nanoparticles causes the relative permittivity of the
nanofluid to exceed that of the vegetable oil.

Table 6. Time to Breakdown of Oil and Nanofluid Samples at Negative Light-


ning Impulse Voltages.

Time to breakdown (μs)

Sample 1 2 3 4 5 Average

Oil 11.6 12.4 9.6 10.8 11.2 11.1


Figure 9. Relative permittivity versus frequency for oil and
Nanofluid 12.6 11.7 14.7 13.2 11.1 12.7
nanofluid.

48 IEEE Electrical Insulation Magazine


Colloids Surfaces A: Physicochem. Eng. Aspects, vol. 335, no. 1–3, pp.
88–97, 2009.
[3] C. Choi, H. S. Yoo, and J. M. Oh, “Preparation and heat transfer proper-
ties of nanoparticle-in-transformer oil dispersions as advanced energy-ef-
ficient coolants,” Current Appl. Physics, vol. 8, no. 6, pp. 710–712, 2008.
[4] J. G. Hwang, M. Zahn, F. M. Sullivan, A. A. Pettersson, O. Hjortstam,
and R. Liu, “Effects of nanoparticle charging on streamer development
in transformer oil-based nanofluids,” J. Appl. Phys., vol. 107, no. 1, pp.
1–17, 2010.
[5] T. V. Oommen, C. C. Claiborne, and C. T. Mullen, “Biodegradable elec-
trical insulation fluids,” in Proc. IEEE Electrical Insulation Conference,
Chicago, IL, pp. 465–468, 1997.
[6] Cooper Power Systems, “ENVIROTEMP® FR3TM fluid,” Bulletin 00092,
Cooper Power Systems, Waukesha, WI, Jun. 2001.
[7] T. V. Oommen, C. C. Claiborne, E. J. Walsh, “Introduction of a new fully
biodegradable dielectric fluid,” in 1998 Annual Report of the Textile,
Fiber and Film Industry Technical Conference, Charlotte, NC, pp. 1–4,
May 1998.
[8] T. V. Oommen, “Vegetable oils for liquid-filled transformers,” IEEE
Electr. Insul. Mag., vol. 18, no. 1, pp. 6–11, 2002.
[9] L. Yang, R. Liao, C. Sun, and M. Zhu, “Influence of vegetable oil on the
thermal aging of transformer paper and its mechanism,” IEEE Trans.
Figure 10. Dissipation factor versus frequency for oil and nano- Dielectr. Electr. Insul., vol. 18, no. 3, pp. 692–700, 2011.
fluid. [10] K. D. Kim, S. S. Kim, Y. H. Choa, and H. T. Kim, “Formation and surface
modification of Fe3O4 nanoparticles by co-precipitation and sol-gel
method,” J. Ind. Eng. Chem., vol. 13, no. 7, pp. 1137–1141, 2007.
[11] R. Hong, T. Pan, J. Qian, and H. Li, “Synthesis and surface modification
Conclusions of ZnO nanoparticles,” Chem. Eng. J., vol. 119, pp. 71–81, 2006.
The following conclusions may be drawn from this work: [12] K. M. Kamruzzaman Selim, Y.S. Ha, S.J. Kim, Y. Chang, T. J. Kim, G. H.
Lee, and I. K. Kang, “Surface modification of magnetite nanoparticles us-
ing lactobionic acid and their interaction with hepatocytes,” Biomaterials,
1. The attenuated total reflection-Fourier transform in- vol. 28, no. 4, pp. 710–716, 2007.
frared spectroscopy, TGA, and DTA data indicate that [13] A. B. Bourlinos, A. Bakandritsos, V. Georgakilas, and D. Petridis,
physical adsorption and chemical bonding occur at the “Surface Modification of ultrafine magnetic iron oxide particles,” Chem.
oleic acid/Fe3O4 nanoparticle interface. Mater., vol. 14, no. 8, pp. 3226–3228, 2002.
[14] T. Tsai, L. Kuo, P. Chen, D. Lee, and C. Yang, “Applications of ferro-
2. The power frequency breakdown voltages of the oil- nanofluid on a micro-transformer,” Sensors, vol. 10, no. 9, pp. 8161–
based nanofluids are 20% greater than those of the oil 8172, 2010.
itself. The positive lightning impulse breakdown volt- [15] X. Li, J. Li, and C. Sun, “Properties of transgenic rapeseed oil based di-
ages of the nanofluids are greater than the correspond- electric liquid,” in IEEE Southeast Conference, Memphis, TN, pp. 81–84,
2006.
ing negative voltages. [16] C. Sun., J. Li, and X. Li, “Electric properties of vegetable oil-based
3. The volume resistivities of the nanofluids and the oil dielectric liquid and lifetime estimation of the oil-paper insulation,” in
are almost equal at frequencies exceeding 100 Hz, and 2006 Annual Report Conference on Electrical Insulation and Dielectric
their dissipation factors are very similar at frequencies Phenomena, Kansas City, MO, pp. 680–683, Oct. 2006.
[17] X. Li, “Research on physical-chemical and electric properties of vegeta-
exceeding 10 Hz. However, the relative permittivity of ble oil based dielectric liquid,” Ph.D. dissertation, Chongqing University,
the nanofluids is greater than that of the oil at all fre- Chongqing, China, July 2006.
quencies between 10−2 and 107 Hz, probably because [18] J. Li, S. Grzybowski, Y. Sun, and X. Chen, “Dielectric properties of rape-
of the much higher relative permittivity of the Fe3O4 seed oil paper insulation,” in 2007 Annual Report Conference on Electri-
cal Insulation and Dielectric Phenomena, Vancouver, British Columbia,
nanoparticles. Canada, pp. 500–503, Oct. 2007.
[19] Insulating Liquids—Determination of the Breakdown Voltage at Power
Acknowledgments Frequency—Test Method, IEC 60156, 1995.
[20] Standard Test Method for Dielectric Breakdown Voltage of Insulating Oils
The authors acknowledge support of this work from the Na- of Petroleum Origin Using VDE Electrodes, ASTM D1816, 2004.
tional Science Foundation of China (Project No. 50877080 and [21] Z. Zhang, J. Li, P. Zou, and S. Grzybowski, “Electrical properties of
No. 51021005). Funding from the Natural Science Foundation nano-modified insulating vegetable oil,” in 2010 Annual Report Confer-
of Chongqing, China (CSTC 2009BA4048) and the 111 Project ence on Electrical Insulation and Dielectric Phenomena, West Lafayette,
IN, pp. 1–4, Oct, 2010.
of the Ministry of Education, China (B08036), are also acknowl- [22] G. Qiu, Q.Wang, C. Wang, W. Lau, and Y. Guo, “Polystyrene/Fe3O4 mag-
edged. netic emulsion and nanocomposite prepared by ultrasonically initiated
miniemulsion polymerization,” Ultrason. Sonochem., vol. 14, pp. 55–61,
2007.
References [23] M. Yamaura, R. L. Camilo, L. C. Sampaio, M. A. Macedo, M. Nakamura,
[1] V. Segal, A. Hjortsberg, A. Rabinovich, D. Nattrass, and K. Raj, “AC
and H. E. Toma, “Preparation and characterization of (3-aminopropyl)
(60 Hz) and impulse breakdown strength of a colloidal fluid based on
triethoxysilane-coated magnetite nanoparticles,” J. Magn. Magn. Mater.,
transformer oil and magnetite nanoparticles,” in Conf. Record of the 1998
vol. 279, no. 2–3, pp. 210–217, 2004.
IEEE International Symposium on Electrical Insulation, Arlington, VA,
[24] J. Amici, P. Allia, P. Tiberto, and M. Sangermano, “Poly(ethylene glycol)-
pp. 619–622, Jun. 1998.
coated Fe3O4 nanoparticles by UV-thiol-ene addition of PEG dithiol on
[2] M. Chiesaa and K. Sarit, “Experimental investigation of the dielectric
and cooling performance of colloidal suspensions in insulating media,”

September/October — Vol. 28, No. 5 49


vinyl-functionalized magnetite surface,” Macro. Chem. Phys., vol. 212, Stanislaw Grzybowski (SM’70, F’99)
pp. 1629–1635, 2011. received the M.Sc. and Ph.D. degrees in
[25] F. Söderlind, H. Pedersen, R. M. Petoral, Jr., P.-O. Käll, and K. Uvdal,
“Synthesis and characterisation of Gd2O3 nanocrystals functionalised by electrical engineering in 1956 and 1964
organic acids,” J. Colloid Interface Sci., vol. 288, no. 15, pp. 140–148. from the Technical University of Warsaw,
2005. Poland, and the Dr. Hab. degree from the
[26] Methods for the determination of the lightning impulse breakdown volt- same university in 1984. In 1956, he joined
age of insulating liquids, IEC 60897, 1987.
[27] R. Richert, A. Agapov, and A. P. Sokolov, “Appearance of a Debye the Faculty of Electrical Engineering at the
process at the conductivity relaxation frequency of a viscous liquid,” J. Technical University of Poznan, Poland,
Chem. Phys., vol. 134, no. 104508, 2011. before moving to Mississippi State Univer-
sity in 1987, where he is now professor and director of the High-
Voltage Laboratory in the Electrical and Computer Engineering
Jian Li (M’05) received the M.S. and Department. His main research interests are in high-voltage en-
Ph.D. degrees in electrical engineering in gineering, aging, partial discharges, and lightning protection. He
1997 and 2001 from Chongqing Univer- is author or coauthor of three books in high-voltage engineering
sity, China, where he is now a professor and more than 200 publications.
and head of the High Voltage and Insula-
tion Technology Department. His major Markus Zahn (F’93) received the
research interests include online detection B.S.E.E., M.S.E.E., and Sc.D. degrees
of insulation conditions in electrical devic- from the Department of Electrical Engi-
es, partial discharges, and insulation fault neering at Massachusetts Institute of Tech-
diagnosis in high-voltage equipment. He is author or coauthor nology (MIT), Cambridge, Massachusetts,
of more than 20 journal papers and 25 papers in international in 1968, 1969, and 1970, respectively. His
conference proceedings. primary research areas are ferrohydrody-
namics and electrohydrodynamics for mi-
Zhaotao Zhang received the B.S. degree crofluidic and biomedical applications, and
in physics from Shandong Normal Univer- nanoparticle technology for improved high-voltage performance
sity, Jinan, China, and the M.S. degree in of electric power apparatus. Zahn is co-inventor on 19 patents.
material physics from Chongqing Normal He has contributed to about 10 book and encyclopedia chapters,
University, China, where he is now a Ph.D. about 115 journal publications, and about 175 conference pa-
student in the High-Voltage and Insulation pers.
Technology Department. His major re-
search interests include online detection of
insulation conditions in electrical devices,
partial discharges, and insulation fault diagnosis for high-volt-
age equipment.

Ping Zou was born in Chongqing, China,


in 1981. He received his B.S. and M.S. de-
grees in physics from Chongqing Univer-
sity in 2003 and 2007, and is now a Ph.D.
student in the High-Voltage and Insulation
Technology Department. His major re-
search interests include online detection of
insulation conditions in electrical devices,
partial discharges, and insulation fault di-
agnosis for high-voltage equipment.

50 IEEE Electrical Insulation Magazine

Potrebbero piacerti anche