Anda di halaman 1dari 10

INSTITUTE OF PHYSICS PUBLISHING JOURNAL OF PHYSICS D: APPLIED PHYSICS

J. Phys. D: Appl. Phys. 38 (2005) 213–222 doi:10.1088/0022-3727/38/2/005

Nanocomposite dielectrics—properties
and implications
J K Nelson and Y Hu
Department of Electrical, Computer & Systems Engineering, Rensselaer Polytechnic
Institute, Troy, NY 12180, USA

Received 29 June 2004, in final form 13 October 2004


Published 6 January 2005
Online at stacks.iop.org/JPhysD/38/213

Abstract
The incorporation of nanoparticles into thermosetting resins is seen to
impart desirable dielectric properties when compared with conventional
(micron-sized particulates) composites. Although the improvements are
accompanied by the mitigation of internal charge in the materials, the nature
of the interfacial region is shown to be pivotal in determining the dielectric
behaviour. In particular, it is shown that the conditions and enhanced area of
the interface changes the bonding that may give rise to an interaction zone,
which affects the interfacial polarization through the formation of local
conductivity.
(Some figures in this article are in colour only in the electronic version)

1. Introduction and background thermally stimulated current measurements, and by changes


in dielectric spectroscopy, particularly at elevated temperatures
The burgeoning application of nanotechnology in the [4, 5]. What is, perhaps, not so clear is the mechanism by which
semiconductor and biological fields was not mirrored in these changes take place. However, it is not only the magnitude
the field of electrical insulation until quite recently, despite and distribution of the internal charge that changes, but there is
a forward looking paper by Lewis [1] a decade ago. However, also a marked difference in the dynamics involved, particularly
recent activity in the formulation of nanodielectrics using, in with the charge decay process. As an example, table 1
particular, both thermosetting resins and polyolefins has shown documents the decay time constants associated with a TiO2 –
some significant promise both in terms of the mechanical [2] epoxy composite system in both micro and nano forms. The
properties and when used as electrical insulants [3]. Clays top line provides the decay time constant of charge taken
and inorganic oxides of nanometric dimensions are the most from a pulsed electroacoustic measurement [4], while the
common particulates used to provide the filler material, but second line represents a similar time constant extracted from
there is also, for good reason, interest in functionalizing the an electroluminescence experiment [5] undertaken on the
particle surface to bring about preferred coupling.
same materials. While the specimen geometry of these two
It has been shown [4] that one of the features of such
experiments is different and so comparison cannot be made in
nanodielectrics is the striking change brought about to the
a vertical direction in table 1, nevertheless, the differences
magnitude and distribution of internal charge associated
between the nanocomposite and microcomposite in either
with the electrical stressing of these materials. This has
experiment are dramatic. Comparison of the values in table 1
been determined through a pulsed electroacoustic study
in the horizontal direction indicates that both the decay of
of nanomaterials in comparison with the same resin with
the incorporation of the same loading of micrometre-
Table 1. A comparison of the charge decay time constants in a
sized particles. The expected Maxwell–Wagner interfacial TiO2 composite through internal charge and electroluminescence
polarization associated with such composite materials measurements.
is significantly altered when the particulate dimensions 38 nm 1.5 µm
are reduced to about that of the polymer chain length (and 20 ˚C TiO2 TiO2
the concomitant interfacial areas increase dramatically). The
Charge decay (s) 22 1800
role played by internal charge can be confirmed through the
Light decay (s) <60 1200
modification of a space–charge peak (the so-called ρ-peak) in

0022-3727/05/020213+10$30.00 © 2005 IOP Publishing Ltd Printed in the UK 213


J K Nelson and Y Hu

80

70
Nanoparticle counts

60

50

40

30

20

10

0
0 20 40 60 80 100 120 140
Nanoparticle diameter (nm)

Figure 1. Particle size distribution of TiO2 nanoparticles


determined by transmission electron microscopy.
Figure 2. Dispersion of 1.5 µm TiO2 particles in epoxy resin.
luminescence and of charge is significantly affected by the
size of the filler particles, and suggests that internal fields are 700

Tip electric field (kV/mm)


mitigated for nanoparticle formulations. 600

500 (b)
2. System studied 400

300
The formulation studied here is a TiO2 –epoxy composite which
200
has been engineered in two forms using micro and nano (a)

particulates. The conventional microparticles have an average 100


1 10 100 1000 10000
diameter of 1.5 µm and the nanomaterial has the particle size
Life (hr)
distribution depicted in figure 1, with a mean size of 23 nm
determined by transmission electron microscopy. Processing Figure 3. The voltage endurance of a 10% (by weight) TiO2 –epoxy
of these materials is critical in order to get proper dispersion of composite in both (a) micro (1.5 µm) and (b) nano (23 nm) forms.
the filler, cross linking, and avoidance of gaseous inclusions. In
practice, it is found that considerable shear forces are necessary one to measure capacitance in the range of 10−15 –10−10 F,
in order to prevent agglomeration, particularly in the case of conductivity up to 5 × 10−15 S, and tan δ down to 10−5 .
nanoparticles. Facilities for calibration were used.
The system studied is based on a Diglycidyl Ether- Conductivity measurements were performed at room
Bisphenol A (DGEBA) resin (Vantico CT1300) with an amine- temperature in a guarded dry air cell. A voltage of up to
based cross-linking agent (Vantico NY956EN). Nanoparticles 30 kV was supplied across the sample under test using a
from Nanophase having a purity better than 99.5% are made by Hipotronics HV power supply. The current was measured
the metal vapour method. They contain 80% anatase and 20% using a Keithley 485 Picoammeter, which was in series with
rutile. Microparticles from Alfa Aesar having 99.5% purity are the sample. Two opposing plane parallel brass electrodes were
also made with a metal based method. The morphology of the used. The lower electrode was split into a circular measuring
microparticles is rutile according to x-ray analysis. Particles electrode and an annular guard ring, which can effectively
are dried by heating to 195 ˚C in vacuum for 12 h prior to negate the alternate conductive paths. At each voltage level,
mechanically compounding with the resin at 40 ˚C using high the current was allowed to settle for 5 min before the reading
rates of shear. A sonic probe (Sonics ultrasonic processor was taken. A compressive load of 104 Pa was applied to the top
Model VC130) is additionally used to alleviate agglomeration. plate to ensure a good contact between the metallized sample
The composite resin and hardener are then vacuum degassed and the electrodes. A metal enclosure was also used to cover
for 2 h at 35 ˚C before being mixed and specimens cast the measuring instrument and components under test, which
in polished stainless steel moulds as has previously been are connected with shielded cables.
described [4]. A cure protocol (48 h at 25 ˚C followed by a
post-cure of 60 ˚C for 3 h) was followed, and the cross-linking 3. Practical measures
reaction checked with differential scanning calorimetry. This
is particularly important since some properties can be sensitive It has already been demonstrated [4] that the mitigation of
to the degree of cross-linking achieved. An example of the internal charge provides an enhancement of the short-term
dispersion obtained for a 10% (by weight) microcomposite is dielectric strength of a filled epoxy resin, and it has recently
shown in figure 2. also been shown that it is possible to realize strengths that
Dielectric spectroscopy measurements were performed on are greater than those of the base polymer. While the
a Novocontrol Alpha Analyser type K in combination with a dielectric withstanding voltage is important, from the industrial
Novocontrol active BDS-1200 sample cell. The sample cell perspective the voltage endurance of the material is perhaps a
has a parallel gold plated electrode design. This set-up allows more critical parameter. Figure 3 shows the voltage endurance

214
Nanocomposite dielectrics—properties and implications

Discharge Magnitude (pC)


250
200 (a)
150
(b)
100
50
0
5 6 7 8 9 10 11
Voltage (kV)

Figure 4. Partial discharge characteristics of divergent field gaps in 10% (by weight) TiO2 –epoxy composite in both (a) micro (1.5 µm) and
(b) nano (23 nm) forms.

characteristic of both the micro- and nanocomposites subjected 1.04 Relative free volume in epoxy composites
to long-term endurance tests using a divergent field formed by
moulding an electrolytically etched tungsten point electrode,
1.02
having a tip radius of about 4 µm, in a sample exhibiting an
inter-electrode gap of 2 mm. The resulting point-plane gaps
were stressed with a 60 Hz alternating voltage, and figure 3 1.00
expressed in terms of the calculated tip stress to minimize hnano/hepoxy
differences between individual point electrodes. At a tip stress 0.98
hmicro/hepoxy
of, say, 200 kV mm−1 , it is evident that the endurance has been
hr(−)

enhanced by 3 21 orders of magnitude.


0.96
The partial discharge characteristics on the same system
are allied measurements and are depicted in figure 4. Two
features are clear. First, the discharge inception voltage for the 0.94
nanocomposite is enhanced, which is expected on the basis of
the previous electroluminescence characteristics [5] in which 0.92
light emission was increased because of the accumulation
of heterocharge in front of the cathode for the micrometre-
0.9
sized filler. Second, the magnitudes of discharges are reduced 360 380 400 420 440 460 480 500
throughout the voltage range. Both these attributes are clearly T (K)
also desirable in the industrial context.
Figure 5. The relative free volume for TiO2 –epoxy nano- and
microcomposites extracted from PVT measurements [9].
3.1. Free volume
While there are some indications that the enhancement in micro- and nanocomposites have a Tg of 73.9 ˚C and 52.4 ˚C,
the breakdown characteristics for nanocomposites may be respectively [4].
related to the known modification in the internal charge, it The finding of an increase in free volume for
has been postulated that the tethered entanglement associated nanocomposites shows that the Artbauer [7] theory cannot
with nanocomposites will reduce the polymer free volume [6]. be invoked to explain the properties and suggests that there
The free volume theory of breakdown [7, 8], which has been must be a more compressible layer or phase associated with
well authenticated for the intrinsic breakdown of polymers, the interfaces where the molecular mobility is higher than in
would then provide an explanation for the improvements the bulk. This is supported by photoluminescence data [5],
cited. The definition of free volume is somewhat varied. which indicate that the molecular environment changes in the
However, in the context of the Artbauer [7] theory of electric presence of sub-micrometre particles. Although these particles
breakdown, whether it includes nanovoids or artefacts that are had not been deliberately functionalized, it suggests that there
not attributable to the chemical structure, is not important. The may be other surface molecules contributing to the effects seen.
‘loosening’ of the material is the relevant issue. Furthermore, it would be expected that the incorporation of
Free volume measurement [9] by the PVT method was TiO2 particles having a permittivity ≈100 into such a resin
performed on both the nanocomposite and microcomposite would increase the real part of the relative permittivity through
for the temperature range 300–500 K. The results, as depicted the Lichtenecker–Rother rule for chaotic mixing. Although
in figure 5, indicate that, in fact, the opposite is true. The measurements [10] taken for microcomposites do, indeed,
nanoformulation exhibits an increase in free volume when show this effect, dielectric spectroscopy for the nanoparticles
compared with the base resin, whereas the micromaterial (at frequencies high enough that interfacial and quasi-dc
does, indeed, show a reduction in free volume. The conductivity effects do not mask the spectra) reproducibly
data in figure 5 have been shown to be reproducible and indicate a reduction in the real part of the permittivity below
statistical tests show them to be meaningful. The base epoxy that of the base resin when measured at 23 ˚C. This has also
system has a glass transition temperature, Tg , of 63.8 ˚C been seen in other systems (such as SiO2 –polyolefin) and is a
determined by differential scanning calorimetry and the clear indication that the large surface areas and higher surface

215
J K Nelson and Y Hu

1.2
Uncured epoxy resin
1.1

1.0
0.9

0.8
0.7
Absorbance

0.6
0.5

0.4

0.3
0.2

0.1
-0.0
-0.1
4000 3500 3000 2500 2000 1500 1000 500
Wavenumber (cm-1)
2.2 Cured epoxy resin

2.0

1.8

1.6

1.4
Absorbance

1.2

1.0

0.8

0.6

0.4

4000 3500 3000 2500 2000 1500 1000 500


Wavenumber (cm-1)

Figure 6. FTIR spectrum of uncured and cured DGEBA resin.

energy density associated with nanoparticles affect the local NH2-CH2-CH2-NH-CH2-CH2-NH-CH2-CH2-NH2


environment.

Primary Secondary Primary


4. The nature of the interfacial region Amine Amine Amine

4.1. Fourier transformed infra-red measurements Figure 7. Structure of the triethylene tetra amine (TETA).

To trace the chemical changes introduced by particles, intensity of –OH groups is difficult to locate. The reason is
Fourier transformed infra-red (FTIR) spectroscopy was used. related to the structure of the triethylene tetra amine (TETA)
Absorption spectra were obtained on a Magna-IR 560 curing agent shown in figure 7 [13]. Among the four reactive
spectrometer in transmission mode. Potassium bromide amine groups, the two on the extremities will act as the primary
(KBr) pressed pellets were prepared as a carrier for liquid amines. The two secondary amine groups in the middle will
state resin and hardener. The spectra of composite and be less active than the primary. After three steps of the
cured resin samples were obtained by using cast thin films polymerization reaction as shown in figure 8, only one –OH
directly. group stays on the cross-linked network.
The base resin (DGEBA), has been well studied and Nanocomposite and microcomposite FTIR spectra are
documented by FTIR [11, 12]. By comparing the spectrum shown in figure 9. The spectra show two major
of the base resin and composites in figure 6, one can differences between the two composites at 600–700 cm−1 and
recognize the curing process as the decrease of the intensity 1050–1150 cm−1 . Around 1070 cm−1 , the nanocomposite
of the characteristic epoxide band [11] at 915 cm−1 and the shows two low split peaks in contrast to the single high peak in
increase of the intensity of the characteristic band of C–N the microcomposite. These two split peaks are also observed
bonds [11] at 1071 cm−1 . The expected increase in the in the uncured resin in the same band. This shows that in

216
Nanocomposite dielectrics—properties and implications

O OH
R1NH2 + CH2-CH-R2 I R1NH-CH2-CH-R2

Primary amine Secondary amine

R1 O R1 OH
II
NH + CH2-CH-R3 N-CH2-CH-R3
R2-CH-CH2 R2-CH-CH2
OH OH

Secondary amine Tertiary amine

III
R1-NH-CH2-CH-R2
O-CH2-CH-R3
OH
Ether group
R 1, R 2, R 3 Carbonyl group

Figure 8. Cross-linking reactions [13].

the nanocomposite, fewer C–N bonds were produced in the By reducing the curing agent, the cross-link density and C–N
curing process, which implies that the cross-link density in bonding will be decreased leaving some unopened epoxide
nanocomposites is less than in the equivalent micromaterial. rings. Figure 10 gives the FTIR spectrum of the under-cured
Two possible mechanisms can cause the cross-link density resin. Two split peaks were observed at the 1050–1150 cm−1
decrease. One is the etherification mechanism [13]. Normally, bands, which is almost the same as in the nanocomposite
if the epoxide ring reacts with the amine group, a three- case. The similarity between these two cases suggests that
dimensional network will be formed with an increase in C–N common properties exist in both of them—the decreased cross-
bonding. The FTIR spectrum shows that both microparticles link density and unopened epoxide rings.
and nanoparticles have –OH groups on the surface [14]. By Based on the above analysis, the spectrum around
adding particles to the polymer, the number of surface hydroxyl 1050–1150 cm−1 appears to be the cross-link characteristic
groups in the material is increased. If the concentration region. In all the FTIR spectra, the base resin gives the highest
of hydroxyl groups is high enough, the epoxide rings will single peak, which indicates the most thorough cross-linking
open up and react with the hydroxyl groups instead of with reaction. The microcomposite’s spectrum shows a lower single
the amine groups. Due to the much larger surface area peak, and the nanocomposite’s peak is the weakest and is split.
of nanometric particles, there will be more –OH groups These findings suggest that both micro- or nano-TiO2 will
at the surface of the particles in the nanocomposite than affect the cross-link density through the surface –OH groups.
in the microcomposite. Another contributing factor to the However, due to the huge surface area of nanoparticles, the
–OH group in nanocomposites has a much more significant
surface –OH group is the drying process. As has been
influence on the cross-link density. As a result, the network
shown by previous thermogravimetric analysis, the dried
will be left with unopened epoxide rings.
nanoparticles may have almost twice the surface –OH group
density than the as-received nanoparticles [14]. As a result,
in the nanocomposite, some epoxide rings will react with 4.2. Dielectric spectroscopy
–OH groups at the interface and the cross-link density in the As a secondary tool in examining the polymer’s behaviour,
nanocomposite will be decreased. Besides the etherification, if dielectric spectroscopy has been used to provide some insight
the curing agent preferably tends to attach onto the surfaces, a into the effects of particles on the network environment.
very thin layer of curing agent will surround the particles. This To complement the FTIR results, dielectric spectroscopy
may keep the curing agent around the particles from reacting was applied to nano- and microcomposites, under-cured
with the resin. This ‘curing agent concentration’ mechanism resin and normal base resin samples. Measurements have
will also decrease the cross-link density. been conducted at temperatures of 298, 313, 328, 343,
To elucidate these two mechanisms, an under-cured resin 358, 373 and 388 K. Dielectric responses were recorded
sample with only 80% of nominal curing agent was prepared. in the frequency range 10−4 –106 Hz. Each sample goes

217
J K Nelson and Y Hu

2.8 Microfilled
2.6
2.4
2.2
2.0
1.8
1.6
Absorbance

1.4
1.2
1.0
0.8
0.6
0.4
0.2
4000 3500 3000 2500 2000 1500 1000 500
Wavenumber (cm-1)

2.0 Nanofilled

1.8

1.6

1.4
Absorbance

1.2

1.0

0.8

0.6

0.4

0.2

4000 3500 3000 2500 2000 1500 1000 500


Wavenumber (cm-1)

Figure 9. FTIR spectra of micro- and nanocomposites.

2.2
Under-cured resin
2.0

1.8
Absorbance

1.6

1.4

1.2

1.0

0.8

0.6

0.4

4000 3500 3000 2500 2000 1500 1000 500

Wavenumber (cm-1)

Figure 10. FTIR spectra of under-cured resin.

218
Nanocomposite dielectrics—properties and implications

Figure 11. Loss tangent data at different temperatures: (from bottom to top: 25 ˚C, 40 ˚C, 55 ˚C, 70 ˚C, 85 ˚C, 100 ˚C, 115 ˚C, respectively)
(a) 10% nanocomposite, (b) 10% microcomposite.

through the measurement at seven different temperatures. Table 2. Calculated activation energies.
Multiple samples for each category were made and measured. TiO2 nanocomposite TiO2 microcomposite
The results show good reproducibility at temperatures
below 100 ˚C. Above 100 ˚C, the variance between samples Below Tg 1.3 ± 0.1 eV 1.1 ± 0.1 eV
Above Tg 1.8 ± 0.1 eV 2.5 ± 0.1 eV
becomes larger but is still of the same order. Besides the
different dispersion of the particles, the variance at elevated
temperatures could also be associated with the post-cure
process (despite the precautions taken), which is associated peaks, whereas above 85 ˚C (i.e. above the glass transition
with the –OH group from the particle surfaces as discussed in temperature, Tg ), the loss tangent of the nanocomposite
section 4.1. It should be noted that the formulation and post- displays broad peaks, which shift to higher frequencies with
cure protocols were not the same as those in [10]. In the increase in temperature. This indicates that the relaxation
low-frequency, high-temperature domain, permittivities rise to processes involved in nanocomposites above Tg are thermally
anomalously high values which is consistent with the work of activated.
Griseri [13]. To facilitate analysis, the loss tangent data were grouped
The loss tangent plots in figure 11 show that the into two categories: below Tg (25–70 ˚C) and above Tg
nanocomposite and microcomposite materials have quite (85–115 ˚C). The activation energy was calculated by the
different dielectric behaviour. The loss tangent of the micro- normalization method [15] by shifting the frequency spectra
composite exhibits an increase with frequency without any laterally and determining the frequency shift required to bring
noticeable strong dispersion in the mid-frequency range. the curves into coincidence. Table 2 shows the calculated
The nanocomposite, on the other hand, shows two different activation energy results.
patterns. In the 25–70 ˚C temperature range, the loss tangent Below Tg , both materials have similar activation energies,
data show a monotonically increasing trend without any while above Tg the microcomposite shows a much higher

219
J K Nelson and Y Hu

1.E+07
microcomposite
1.E+06 nanocomposite

Real relative permittivity


base resin
1.E+05
less-cured resin
1.E+04

1.E+03

1.E+02

1.E+01

1.E+00
1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07
Frequency/Hz

Figure 12. Comparison of the real part of relative permittivity at 115 ˚C.

activation energy. This shows that when the temperature Recently Tschope [17] and Lewis [18] have extended
is above Tg , the nanocomposite network has more mobility the Gouy–Chapman layer theory to solids. In composites,
than the equivalent microcomposite. With the free volume an interaction zone can be expected between particles and
measurement results at the above Tg temperature (see the epoxy network. When the particles’ size is reduced to
figure 5), the mobile network activated by the temperature in less than 100 nm, the interface between the particles and their
nanocomposites can be expected to contribute to the observed surrounding becomes very important. For an evenly dispersed
increase of free volume. While the interface interactions nanocomposite system with uniform sized spherical particles
have been approached in section 4.1 from the cross-linking the interparticle distance, l, can be expressed as
perspective, there is also increasing speculation that this zone   
may also be described in terms of a Gouy–Chapman layer as 4π 1/3
discussed in section 5. l=r −2 , (1)
3v
As shown in figure 12, the under-cured resin gives lower
real permittivity than base resin. This implies a connection
where r is the radius of the particle and v is the volume
between the cross-link density and the real part of the
fraction [19]. The distance between the 23 nm particles for
permittivity. In the under-cured sample, due to the insufficient
the 10% nanocomposite would be about 37 nm. This result
curing agent, some epoxide rings will be kept. Compared
suggests that the system is dilute, and there would not be much
to the linear C–O–C linkages formed by the opening-up of
overlap between particles or the interfacial regions. However,
epoxide rings, the unopened epoxide rings may be expected
as shown in the distribution of particle sizes in figure 1, there
to give a different dielectric response. The decrease in the
number of C–N bonds could also contribute to the lower are some large particles with diameters up to 100 nm. In
permittivity. Similar results have been seen in other epoxy addition, despite the precautions taken, it is known that there
resin composites [16]. will be some agglomeration in the bulk resulting in local
In the nanocomposite, due to the curing agent overlapping interfacial regions. Indeed, the insensitivity of the
concentration around the particles and the large surface area, bulk conductivity (see later) to the size of the particles further
there will be a surfeit of unreacted epoxy resin, which would confirms that the percolation limit has not been reached. One
give rise to the low permittivity observed at room temperatures, may thus speculate that, because these interfacial zones exist
cited in section 3.1. At elevated temperatures the attracted locally, there may be local modifications of the conductivity
curing agent around the nanoparticles may be stimulated to without significant changes in the bulk properties, as confirmed
react with unopened epoxide rings. At the same time, the by experimental results [3, 5]. Such local conductivity can
remaining –OH groups on particle surfaces may also take part provide paths for charges to move relatively freely along
in bonding. limited distances, and may provide an explanation for the faster
time response observed in nanocomposites (table 1). This is a
5. Discussion and appraisal feature of the double layer model proposed by Lewis [18].
Figure 13 shows the comparison of the relative
In the partial discharge and voltage endurance tests, the permittivity between nano- and microcomposites at 115 ˚C. In
composites based on nanoparticles clearly showed properties the low frequency region, the real and imaginary permittivities
superior to traditional microcomposites. Combined with the of the nanomaterial are parallel to each other with a slope
earlier pulsed electroacoustic (PEA) and electroluminescence of −1 on a log–log plot. This is strong evidence of
results [10], the improvement shown in the nanocomposite case a ‘low frequency dispersion’ [15]. This low frequency
appears to be closely related to the control of internal charges dispersion may be regarded as being primarily associated
in the bulk. In turn, the mitigation of internal charge normally with charge carriers present in the material with limited
associated with interfacial polarization in composite structures movement. The microcomposite, on the other hand, shows
is believed to be related to the interfacial area introduced by a real permittivity which has only half the slope of the
nanoparticles. imaginary part. One may speculate that this phenomenon

220
Nanocomposite dielectrics—properties and implications

1.E+07
nanocomposite-real
1.E+06
nanocomposite-
1.E+05 imaginary
Relative permittivity
microcomposite-real
1.E+04

1.E+03
microcomposite-
imaginary
1.E+02

1.E+01

1.E+00

1.E-01

1.E-02
1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07

Fequency/Hz

Figure 13. Comparison of relative permittivity at 115 ˚C.

3.50E-13 1.E+01

Relative permittivity (imaginary)


Volume conductivity (S/cm)

Microcomposite
3.00E-13 Nanocomposite
2.50E-13 1.E+00

2.00E-13

1.50E-13 1.E-01
1.00E-13 nanocomposite
microcomposite
5.00E-14
1.E-02
0% 10% 20% 30% 40% 50%
1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 1.E+06 1.E+08
Loading (% by weight) Frequency

Figure 14. Comparison of volume conductivity as a function of Figure 15. Permittivity comparison at 25 ˚C.
loading.

may also result from the limited conduction associated with between the nanoparticles and the epoxy resin matrix. These
the interaction zones and explain the favourable mitigation bonds between reactive radicals on the particle surfaces and
and relaxation of the internal charge. Lewis [18] has also the epoxy resin are probably the origin of the Gouy–Chapman
pointed out that the observed unusual high permittivity of layer. This would be consistent with the observed differences
nanocomposites at low frequencies would result from an out- in the cured structure for the nanocomposite indicated by the
of-phase dipole moment developed by the charges accumulated FTIR results. However, it has become clear that these interface
at the particles’ poles. Such augmentation of the permittivity regions are pivotal in determining the dielectric properties of
would create values substantially greater than those derived nanodielectrics, and dominate by virtue of the large interfacial
from a Maxwell–Wagner theory alone. areas involved.
A bulk dc-conductivity measurement was also undertaken
to study the effects of the local conductivity on the bulk Acknowledgments
material. The experimental method is introduced in section 2.
The results are shown in figure 14. At all the tested The authors are indebted to Drs L Schadler, R MacCrone,
loading levels, the nanocomposite has slightly lower bulk C W Reed and L Utracki who have contributed to this work in
conductivity than the microcomposite. This result is consistent
numerous different ways.
with the dielectric measurement result at 25 ˚C, in which
the nanocomposite has a lower imaginary permittivity than the
microcomposite as shown in figure 15. The results indicate References
that the local conductivity does not necessarily (for modest
loadings) imply a degraded bulk conductivity. [1] Lewis T J 1994 Nanometric dielectrics IEEE Trans. Diel.
The nature of the Gouy–Chapman layer is still Electr. Insul. 1 812–25
[2] Irwin P C, Cao Y, Bansal L and Schadler L S 2003 Thermal
under investigation, since, unlike the liquid phase, the and mechanical properties of polyimide nanocomposites
electrochemistry here is poorly understood. Recent electron Ann. Rep. Conf. on Electrical Insulation and Dielectric
paramagnetic resonance (EPR) measurements [9] on this Phenomena IEEE (Albuquerque, NM) pp 120–3
polymer system reveal more about the interfacial area. The [3] Henk P O, Kortsen T W and Kvarts T 1999 Increasing the
result shows that the TiO2 nanoparticle surface has many electrical discharge endurance of acid anhydride cured
DGEBA epoxy resin by dispersion of nanoparticle silica
reactive species. After the particles were incorporated into High Perf. Polym. 11 281–96
epoxy resin, the EPR result shows a significant decrease in [4] Nelson J K, Fothergill J C, Dissado L A and Peasgood W 2002
the oxygen radical species, which indicates active bonding Towards an understanding of nanometric dielectrics Ann.

221
J K Nelson and Y Hu

Rep. Conf. on Electrical Insulation and Dielectric [12] Scherzer T 1996 Characterization of the molecular
Phenomena IEEE (Cancun, Mexico) pp 295–8 deformation behavior of glassy epoxy resins by rheo-optical
[5] Nelson J K, Hu Y and Thiticharoenpong J 2003 Electrical FTIR spectroscopy J. Polym. Sci. 34 459–70
properties of TiO2 nanocomposites Ann. Rep. Conf. on [13] Griseri V 2000 The effects of high electric fields on an epoxy
Electrical Insulation and Dielectric Phenomena IEEE resin PhD Thesis University of Leicester, UK
(Albuquerque, NM) pp 719–22 [14] Ma D 2003 Investigation into the dielectric behavior of
[6] Sternstein S S and Zhu A-J 2002 Reinforcement mechanism titanium dioxide/polyethylene nanocomposites PhD Thesis
of nanofilled polymer melts as elucidated by nonlinear Rensselaer Polytechnic Institute, USA
viscoelastic behavior Macromolecules 35 7262–73 [15] Jonscher A K 1983 Dielectric Relaxation in Solids (London:
[7] Artbauer J 1996 The electric strength of polymers J. Phys. D: Chelsea Dielectrics)
Appl. Phys. 29 446–56 [16] Imai T, Hirano Y, Hirai H, Kojima S and Shimizu T 2002
[8] Dissado L A and Fothergill J C 1992 Electrical Degradation Preparation and properties of epoxy-organically modified
and Breakdown in Polymers (Peter Peregrinus) layered silicate nanocomposites IEEE Int. Symp. on
[9] Nelson J K, Utracki L A, MacCrone R K and Reed C W 2004 Electrical Insulation (Boston, MA) pp 379–83
The role of the interface in determining the electrical [17] Tschope A 2001 Grain size-dependent electrical conductivity
properties of nanocomposites Ann. Rep. Conf. on Electrical of polycrystalline cerium oxide space charge model Solid
Insulation and Dielectric Phenomena IEEE (Boulder, CO) State Ion. 139 267–80
[10] Nelson J K and Fothergill J C 2004 Internal charge [18] Lewis T J 2004 Interfaces and nanodielectrics are synonymous
behaviour of nanocomposites Nanotechnology 15 586–95 Int. Conf. on Solid Dielectrics (Toulouse) pp 792–5
[11] Cherdoud-Chihani A, Mouzali M and Aradie M J M 2003 [19] Hong J I, Schadler L S and Siegel R W 2003 Rescaled
Study of cross-linking acid copolymer/DGEBA systems electrical properties of ZnO/low density polyethylene Appl.
by FTIR J. Appl. Polym. Sci. 87 2033–51 Phys. Lett. 82 1956–8

222

Anda mungkin juga menyukai