You are on page 1of 6

J. Am. Ceram. Soc.

, 91 [11] 3593–3598 (2008)


DOI: 10.1111/j.1551-2916.2008.02704.x
r 2008 The American Ceramic Society

Journal
An AFM/EFM Study of the Grain Boundary in ZnO-Based
Varistor Materials
Simoni M. Gheno,w,z Ruth H. G. A. Kiminami,z Márcio M. Morelli,z Jusmar V. Bellini,y
and Pedro I. Paulin Filhoz
z
Department of Materials Engineering—DEMa, Federal University of São Carlos—UFSCar Rod. Washington Luiz,
São Carlos 13565-905, SP, Brazil
y
Department of Physics—DF, State University of Maringa, UEM, Maringa 87020-900, PR, Brazil

Zinc oxide (ZnO)-based varistors are metal oxide varistors Zinc oxide (ZnO), one of the various types of ceramic va-
whose nonlinear properties are characterized by electrical resis- ristors, is a semiconducting electroceramic material commonly
tance that decreases as the applied field increases. The multi- used in commercial varistors to protect electrical/electronic com-
phasic nature of varistors leads to the formation of Schottky ponents from transient overvoltages. Furthermore, ZnO ceram-
barriers, which are responsible for the materials’ nonlinear be- ics have long been used as varistors because of their highly
havior. The objective of this work was to image the potential nonlinear current–voltage (I–V) characteristics.3–12
barriers in ZnO doped with 0.5 mol% Cu and x wt% G (G is a ZnO varistor ceramics have a highly complex structure, which
frit and x 5 0, 1, and 5 wt%). The frit served to form a glassy makes it difficult to control their morphology and electrical
insulating layer around the grain boundaries. Samples were sin- characteristics. The main factor for the success of the technology
tered at 10501C and the microstructures were analyzed using a is the discovery of various dopants that impart the desired elec-
Nanoscope IIIa atomic force microscope. The results of the trical properties to ZnO, while also allowing for the necessary
electric force microscopy experiments map the electric field dis- microstructural control.3,9–18
tribution on the surface of CuO–ZnO-based ceramics. The complexities of the microstructure, electrical response,
processing, and synergistic effects have made defining the role(s)
of specific additives elusive. The energy states in the microstruc-
ture of ZnO-doped ceramics are created by dislocations and
I. Introduction dangling bonds arising from the crystallographic mismatch be-
tween adjacent grains. Free carriers, usually electrons present in
T ODAY’S demand for electrical/electronic components re-
quires the development of materials with controllable prop-
erties, which include varistor ceramics.1–6 Varistor ceramics are
the slightly n-type ZnO material, are trapped, leading to the
creation of electrostatic barriers at the interface. The potential
gaining increasing technological importance because of their barrier prevents current conduction, yielding large impedance at
highly nonlinear electrical characteristics. They are used to pro- relatively low voltages and nonlinear characteristics.11,13,18 The
tect electrical circuits from voltage surges and exhibit highly doped grain boundaries trap electrons and create an electrostatic
nonlinear I–V characteristics.3,5–11 Over the last few decades, barrier against charge transport. The release of electrons at the
varistor behavior has been improved through advances in pro- breakdown voltage gives the material its nonlinear characteris-
cessing, including the type, quality, and quantity of reagents tics, which are essential for the varistor function and have been
attributed to phenomena that occur in grain boundaries.21,22
used.8,12–17
An important factor that must be kept in mind is that va- Varistors act as insulators until they reach the breakdown
ristors are polycrystalline ceramics with high concentrations of voltage, whereupon they act as conductors. This voltage is pro-
structural, surface, and electronic defects. The type and quantity portional to the number of barriers formed between parallel
of defects are related directly to the processing stages used in the sides of the varistor. The conduction mechanism through
production of the ceramic. Therefore, the main feature of these Schottky barriers may be due to thermionic emission from the
systems is the presence of grains connected at their interfaces. ZnO-doped grain conduction band to the isolating film conduc-
This is considered a determining factor of the systems’ electrical tion band. Eda,23 who studied doped ZnO, reported that
properties, i.e., their stability after a transient voltage surge is Schottky-type barrier voltages are separated by an insulating
directly related to the stability of the electrical barriers in the film of finite width. He proposed an energy band model at the
grain boundaries.3,4,9,17 This stability is affected by the choice of grain boundary and a conduction mechanism mainly governed
by field emission and thermionic emission at the reverse-biased
dopants, and it is also conditioned to the uniformity of the do-
pants’ distribution at the grain boundaries and to the chemistry Schottky barrier.
at the boundaries. If the dopants are distributed uniformly at The barrier voltage of ZnO-doped varistors can be electri-
numerous grain boundaries, more borders will be active and the cally, chemically, and thermally degraded during use, leading to
current will be conducted uniformly through the device. Greater the reduction of barrier voltage height and, consequently, to the
homogeneity leads to improved reproducibility of the process, increase of leakage current, which could be catastrophic for
and enhances and equalizes the material’s properties.5,17–20 surge arresters. The degradation of this barrier has been studied
extensively.22–27 However, little effort has focused on its imag-
ing, and its real dimensions are not well known, although this
L. M. Levinson—contributing editor
information is crucial in ZnO varistor technology.
The durability of varistors is related directly to the state of ion
polarization at the grain boundaries, which create the potential
Manuscript No. 24643. Received May 7, 2008; approved August 8, 2008. barrier. This polarization state is produced by ion diffusion.
This work was financially supported by the Brazilian agencies FAPESP and CAPES the This study used electrical force microscopy (EFM) to image and
LNLS (Brazilian National Laboratory of Synchrotron Light) in Campinas, SP, Brazil for
the AFM/EFM analyses.
dimension the potential barrier in ZnO–CuO–frit varistors. The
w
Author to whom correspondence should be addressed. e-mail: gheno@dema.ufscar.br investigation and mapping of the potential barrier distribution
3593
3594 Journal of the American Ceramic Society—Gheno et al. Vol. 91, No. 11

across the ZnO–CuO–frit surface was carried out using atomic frequency shifts between the oscillations of the bias, the canti-
force microscopy (AFM/EFM), varying the potential applied on lever, and the piezoelectric driver as a function of bias voltages
the surface. applied to the cantilever. The initial EFM imaging conditions
The purpose of the use of the varistor system based on CuO- were: interleaved frequency drive, 25 Hz; integral gain, 0.35; and
doped ZnO was to confirm the results obtained previously by proportional gain, 2.5. The images of surface potential and bar-
Chiou and Chung28 and Kutty and Raghu.29 These authors rier layer were obtained by applying 0, 4, 8, and, 12 V in situ to
showed that the presence of CuO in ZnO varistors conferred to the sample. Before the EFM imaging, metal electrodes (conduc-
the ZnO–CuO varistors the same properties as those that have tive Ag ink) were applied to the bottom and the sides of the
several dopings such as those proposed by Matsuoka and col- samples and kept in contact with the AFM scanner throughout
leagues12,20 However, Lee and Tseng30–33 showed that due to the the measurements. Imaging of these samples as fired was carried
poor sinterability of this system and in an attempt to densify the out at room temperature.
material (CuO doped ZnO), the frit used had the following
composition: 26 wt% SiO2, 62 wt% PbO, 7 wt% B2O3, and
5 wt% ZnO. Previous studies on ZnO–CuO–frit varistor sys- III. Results and Discussion
tems have been reported by Bellini et al.34–37 They showed that
the addition of a large percentage of frit (5%) led to a low Figure 1 shows the typical XRD patterns of the pure ZnO (bot-
breakdown electric field and a low leakage current. tom) and ZnO–CuO (top) powders. Figure 2 shows the XRD
patterns of ZC (bottom), ZCF1 (middle), and ZCF5 (top) va-
ristor pellets. Based on the powder diffraction files of the Joint
Committee on Powder Diffraction Standards all the diffraction
II. Experimental Procedures planes of the maximum relative intensity of the crystalline
Varistor powders of CuO-doped ZnO were prepared by the phases are contained within the range of 101o2yo801, and
freeze-drying process as per Bellini et al.34 Two pellets were ob- only the ZnO phase is visible in Figs. 1 and 2.
tained by pressing the powders at 75 MPa in a uniaxial press No Cu peak is visible in Figs. 1 or 2 due to the low Cu content.
without pressing additives and sintering in air at 10501C for 1 h. In Fig. 1, the unexpected XRD peaks appear at about 2y 5 19
Pellet (a) was used for AFM/EFM characterization with 5 mm and 28.7, which is likely due to impurities or remnants of cupper
diameter and 3 mm thickness and pellet (b) was used for analysis acetate although they were not detected in the sintered samples
of electrical properties with the same diameter but with a 1 mm (Fig. 2). In Fig. 2, the XRD peaks (2y 5 26.71 and 341) show a
thickness. new phase: Zn2SiO4 from the reaction of the vitreous phase at a
For the purposes of identification, the samples were given the high temperature with the zinc grain for both ZCF1 and ZCF5.
following specific names: ZC: ZnO10.5 wt% Cu, ZCF1: However, the amount of the Zn2SiO4 phase developed was not
ZnO10.5 wt% Cu11 wt% frit, and ZCF5: ZnO10.5 wt% sufficient to modify the properties of the ZnO–CuO–frit varistor
Cu15 wt% frit. The frit was prepared as follows: a batch of pre- system. In Fig. 2, a decrease in the intensity of the peaks with an
cursor oxides and isopropyl alcohol was mixed in a ball mill mixer increase in the frit content was observed and attributed to the
with zirconia mixing media for 4 h. The oxide powders were ‘‘diluting effect’’ of the amorphous phase introduced.
batched in the following proportions: 26 wt% SiO21 Figure 3 presents the current density versus applied electric
62 wt% PbO17 wt% B2O315 wt% ZnO. The powder mixture field (J–E) curves of the ZC, ZCF1, and ZCF5 varistors, showing
was dried at 1101C for 24 h and unaggregated before being melted that the a value of ZCF1 (a 5 8.1) was greater than ZCF5
at 9001C in a conventional furnace and fritted in distilled water. (a 5 7.0). It is worth mentioning that variations in the frit con-
The ZnO, ZnO10.5 wt% Cu powders and the crystalline centration affected the nonlinear coefficient (a), improving the
phases of sintered ZC, ZCF1, and ZCF5 samples were analyzed overall varistor material behavior. The ZCF1 resistance is higher
by X-ray diffraction (XRD) using a Siemens D5005 diffracto- than ZCF5 at high voltages. At low voltages, the resistance is
meter with Cua radiation, at 50 kV voltage and 100 mA current. almost the same for all materials, as it should be expected for
The current density versus electric field (J–E) curves were plot- varistors (high resistance at low voltages).
ted from the I–V data of ZnO–CuO-doped varistors measured Based on these results, an analysis was performed of the mi-
with a high-voltage 60 Hz ac power supply (Keithley) whit GPIB crostructural and electrical properties by scanning probe micros-
interface. Electrodes (0.385 cm2) of conductive silver paste were copy (SPM), i.e., AFM combined with EFM. The results were
deposited on the pellets’ surface as-fired and heat treated at analyzed and interpreted using the Digital Instruments software,
1101C for 10 min. The nonlinear coefficient a is considered the which is specific for analyzing SPM images. Figures 4–6 show
quality factor of a varistor, and is defined by the empirical images of 15 mm  15 mm scanned areas of the ZC, ZCF1, and
equation I 5 kVa, where I is the current flowing through the ZCF5 samples obtained via AFM/EFM.
sample, V is the voltage across the sample, k is a constant value
of ZCF1 and ZCF5 samples, which are 2.73  1032
mA  (V  m)1 and 7.93  1027 mA  (V  m)1 respectively,
and a is the nonohmic exponent. For the sake of comparison,
the nonlinear coefficient a can be estimated by the equation
a 5 1/log (E5/V1) using two electric field values, E5 and E1, at
5 and 1 mA/cm2 settings of the current density, respectively.36
The potential barrier was analyzed with a Nanoscope IIIA
AFM (Digital Instruments) operating in the EFM mode, which
was equipped with an electronic extension module (Veeco Instru-
ments, Veeco Metrology Inc., Santa Barbara, CA). Topograph-
ical measurements and electrical data were obtained by the two-
pass technique (Lift Mode). In this configuration and operating
in the Tapping Mode, the probe scans a topographical line during
the first pass. In the second scan, the cantilever is lifted to a pre-
defined distance (75 nm) in order to minimize the effect of the
Van der Waals forces, while the probe detects variations in the
electric force gradient over the same line and the influence of
surface topography is ruled out.38–41 An NSC15 tip (Mi-
kroMasch) was used in all the experiments. Electrostatic force Fig. 1. Powder X-ray diffraction pattern of samples sintered in air at
gradient images were obtained by monitoring the phase and 10501C for 1 h: ZnO (a) and ZnO10.5 mol Cu (b).
November 2008 AFM/EFM Study in ZnO-Based Varistor 3595

Fig. 3. Current density versus. applied electric field (J–E) curves of the
ZC( & ), ZCF1 (J), and ZCF5 (W) samples. The breakdown fields for
the ZCF1 and the ZCF5 are 2500 and 4050 V/cm, respectively.

Fig. 2. X-ray diffraction patterns of samples sintered in air at 10501C


centration of negative charges at the grain boundary, whose be-
for 1 h: ZC (a), ZCF1 (b), and ZCF5 (c).
havior signals the effective presence of the potential barrier.
The increased conductivity resulting from doping was con-
Figures 4(a), 5(a), and 6(a) show the topographic AFM in- firmed by the EFM images, which recorded the potential bar-
formation about the respective microstructures, whose grains riers present at the grain boundaries. Figures 4(b)–(e) to 6(b)–(e)
differ considerably in geometry and size probably due to the show the EFM images obtained using a dc bias voltage of
presence of frit but reveals no chemical specificity or phase in- 0–12 V. In 4 V increments, the grain boundaries are already
formation. Besides, in Figs. 4(a) and 5(a), the heterogeneous visible, in line with the topographic image. The call-out regions,
microstructure exhibits grains and subgrains due to the low sin- which were attributed to semiconducting material, indicated
terability of the mixture and residual porosity due to incomplete that the grains were conductive. The bright regions, especially
densification during the sintering process. Porosity can increase at the grain boundaries, could be ascribed to the electrostatic
the leakage current, favoring electrical degradation. This degra- charge produced by the potential barrier.
dation causes heating during the operation of the device, leading In all EFM images, the surface potential is dependent on the
to its failure. However, the nonlinear coefficient (a) remains un- local electric field. Besides, the surface potential clearly identifies
affected, as shown in Fig. 3. a grain boundary contour. Potential profiles were extracted
The EFM results presented in Figs. 4(b)–(e), 5(b)–(e), and from various positions along the grain boundaries in all the
6(b)–(e) of ZC, ZCF1, and ZCF5 samples, respectively, show samples and, based on these profiles, the dimensions of potential
details of the variations in the electric field gradient across the barriers at the grain boundaries were determined to be 250 nm
area scanned on the sample. The EFM images indicate that an for ZC and ZCF1 and 750 nm for ZCF5 as shown in Fig. 7.
increase in the applied voltage leads to an increase in the con- These results indicated that increasing the frit content to 5%

Fig. 4. Atomic force microscopy (AFM) and electric force microscopy (EFM) profiles of a 15  15 mm area of a pure ZC varistor sample. In (a), an
AFM topographic scan is shown. In (b) through (e), EFM profiles of the same region as (a) are shown but at 0, 4, 8, and 12 V dc, respectively. Grain sizes
are visible (I) as well as the grain boundaries/intergranular regions (II) and porosity (III).
3596 Journal of the American Ceramic Society—Gheno et al. Vol. 91, No. 11

Fig. 5. Atomic force microscopy (AFM) and electric force microscopy (EFM) profiles of a 15  15 mm area of a pure ZCF1 sample. In (a), an AFM
topographic scan is shown. In (b) through (e), EFM profiles of the same region as (a) are shown but at 0, 4, 8, and 12 V dc, respectively. Grain sizes are
visible (I) as well as the grain boundaries/intergranular regions (II).

tripled the potential barrier. The width of the resistive layer is a decreases due to a substantial contribution of the glassy phase
function of the material’s chemistry and processing and is not toward the conductivity.
expected to be very sensitive to the voltage applied (Fig. 7). One Several theoretical studies and some scanning probe technical
possible explanation for this behavior is that the contribution to studies are found in the literature and can be used to quantify
the conductivity of the glassy phase increases with an increase in the interfacial properties associated whit local transport prop-
its amount. The width of the potential barrier increases for the erties, grain boundaries, or potential barrier; there are also
same reason while the conductivity at the breakdown voltage some references that associate these properties with different

Fig. 6. Atomic force microscopy (AFM) and electric force microscopy (EFM) profiles of a 15  15 mm area of a pure ZCF5 sample. In (a), an AFM
topographic scan is shown. In (b) through (e), EFM profiles of the same region as (a) but at 0, 4, 8, and 12 V dc, respectively. Grain sizes are visible (I) as
well as the grain boundaries/intergranular regions (II).
November 2008 AFM/EFM Study in ZnO-Based Varistor 3597

The formation of defects in the ZnO matrix presumably gives


rise to potential barriers at the grain boundaries. Potential bar-
rier height and width are both associated with an increase in the
number of electrons and holes in the grain and grain boundary
regions, and increases in the potential barrier can be attributed
to effectively increased grain boundary states, which have been
found to give rise to nonlinear current/voltage characteristics
through the formation of barriers against electrical conduction.
Moreover, Eda23 demonstrated that Schottky barrier deforma-
tions may occur, which can be attributed to the nonuniform
distribution of spatial charges that are necessarily symmetrical
to the Schottky barriers. In other words, the Schottky barriers
symmetrical to the grain boundary must necessarily be sepa-
rated by an intergranular layer of finite distance, thus causing
the asymmetrical deformation of the Schottky barrier.
The results presented in Figs. 4–6 indicate that the frequency
Fig. 7. Dimension of potential barrier at 12 V dc: ZC(n), ZCF1 (&), gradient height is strongly dependent on the applied voltage and
and ZCF5 (J) samples. the Schottky barrier is proportional to the applied field. The
width of the region over which the voltage changes near an in-
dopants.2–5,9–13,17,24–27,34,42–44 Shao et al.43 made an important terface is a function of the material’s parameters and is not ex-
contribution by analyzing the grain boundary behavior in a pected to be very sensitive to the applied voltage. The depletion
commercial polycrystalline ZnO using a nanoimpedance micros- of free charge carriers applies equally to volume and barrier
copy/spectroscopy technique that is based on impedance spec- layers. Sato et al.42 who studied the effect of Pr doping on the
troscopy with a conductive AFM tip. Through this technique, formation of double Schottky barriers at ZnO single grain
which monitors the tip-to-surface interaction, these authors boundaries, proposed that the formation of potential barriers
showed that the electrical properties of polycrystalline ZnO is strongly governed by atomic scale factors, such as the grain
are dominated by the resistive grain boundary effect and con- boundary atomic structure, dopant segregation, and/or forma-
cluded that all grain boundaries between the tip and the bottom tion of native defects. Another important explanation for the
electrode were probed simultaneously. variation in thickness was proposed by Heywang,27 who postu-
Vasconcelos et al.44 studied another electroactive grain lated that the interface charge changes the Fermi level in the
boundary of a highly dense metal oxide SnO2-based polycrys- vicinity of the grain boundary. The potential barrier height and
talline varistor using electrostatic force microscopy (EFM). As width are both associated with an increase in the number of
demonstrated by these authors, the EFM images in the grain electrons and holes in the grain and grain boundary regions. In
boundary region indicated the presence of active potential bar- the grain boundary regions, the number of trapped electrons
riers or active grain junctions. resulting from the segregation of dopants at the grain bound-
In addition, the EFM images provide information about how aries, as well as the creation of positive defects in the depletion
the electrical current in varistors follows the path of least resis- layer and negative defects at the interfaces (metallic vacancies),
tance, i.e., routes with fewer borders (or larger grains) and those are also related to the characteristics of the potential barrier.
with smaller grain boundary barriers. If the grain size is homo- Thus, potential barrier increases can be associated with effec-
geneous, more electric current passes through many parallel tively increased grain boundary states.
paths. The effects of porosity and inhomogeneous grain distri- Grain boundaries have been found to give rise to nonlinear
bution should be avoided, as illustrated in Fig. 4. current/voltage characteristics by forming barriers against elec-
Over the years, a number of models have been developed for trical conduction. In polycrystalline materials in general, most
the barriers to electrical conduction at the interfaces. The first grain boundaries grow randomly and are low-symmetry, high-
model, which was proposed by Matsuoka in 1971,12 considered angle grain boundaries. The basic concept underlying varistor
a space-charge-limited current in the bismuth-rich intergranular action is that the current/voltage characteristics are controlled
layer. In CuO-doped ZnO varistors, the barrier associated with by the existence of electrostatic barriers at the interfaces between
the nonohmic properties of doped-ZnO ceramics can be attrib- grains. The origin of these barriers is the interface charge stem-
uted to the high-resistance intergranular layer between the low- ming from lattice mismatches, defects, and dopants at the grain
resistance CuO-doped ZnO grains, as indicated by the ideal boundary. The interface charge changes the Fermi level in the
model proposed by us in Fig. 8, which shows that superficial vicinity of the grain boundary, resulting in band bending. The
donor states are able to donate electrons. This figure is com- electronic charges stored in an interface represent a repulsive
patible with the nonlinear coefficient and the EFM results, be- potential for the majority carriers—the electrons in the case of
cause the basic concept underlying varistor action is that the an n-doped semiconductor—across the interface.
current/voltage characteristics are controlled by electrostatic
barriers at the grain interfaces.

III. Conclusions
CuO-doped ZnO varistors are highly disordered materials and
their grain boundaries are nonhomogeneous, showing structural
variations along the interface. Our EFM results showed that the
height and width of potential barriers in all the samples were
dependent on the bias voltage. Moreover, the potential barriers
associated with the increased number of electrons and holes in
the grain and at the grain boundary may be related to a resistive
layer. This layer may be associated with spatially charged re-
gions resulting from charge redistribution, indicating that the
segregated atoms of the dopant (frit) may stabilize the ZnO
grain boundary that is rich in disordered oxygen species, thereby
Fig. 8. Ideal energy band model of ZnO–CuO-doped ceramics. enhancing the varistor properties.
3598 Journal of the American Ceramic Society—Gheno et al. Vol. 91, No. 11
22
Acknowledgment U. Schwing and B. Hoffmann, ‘‘Model Experiments Describing the Micro-
contact of ZnO Varistors,’’ J. Appl. Phys., 57 [12] 5372–9 (1985).
23
The authors gratefully acknowledge the LNLS (Brazilian National Laboratory K. Eda, ‘‘Conduction Mechanism of Non-Ohmic Zinc Oxide Ceramics,’’
of Synchrotron Light) in Campinas, SP, Brazil for the AFM/EFM analyses. J. Appl. Phys., 49 [5] 2964–72 (1978).
24
E. R. Leite, J. A. Varela, and E. Longo, ‘‘Barrier Voltage Deformation on of
ZnO Varistors by Current Pulse,’’ J. Appl. Phys., 72 [1] 147–50 (1992).
25
References J.-H. Choi, N.-M. Hwang, and D.-Y. Kim, ‘‘Pore–Boundary Separation Be-
havior During Sintering of Pure and Bi2O3-Doped ZnO Ceramics,’’ J. Am. Ceram.
1
M. R. Cássia-Santos, V. C. Sousa, M. M. Oliveira, F. R. Sensato, W. K. Soc., 84 [6] 1398–400 (2001).
26
Bacelar, J. W. Gomesa, E. Longo, E. R. Leite, and J. A. Varela, ‘‘Recent Research H. S. Domingos, J. M. Carlsson, P. D. Bristowe, and B. Hellsing, ‘‘The For-
Developments in SnO2-Based Varistors,’’ Mater. Chem. Phys., 90 [1] 1–9 (2005). mation of Defect Complexes in a ZnO Grain Boundary,’’ Int. Sci., 12, 227–34
2
M. R. Cássia-Santos, V. C. Sousa, M. M. Oliveira, P. R. Bueno, W. K. Bacelar, (2004).
27
M. O. Orlandi, C. M. Barrado, J. W. Gomes, E. Longo, E. R. Leite, and J. A. W. Heywang, ‘‘Resistivity Anomaly in Doped Barium Titanate,’’ J. Am.
Varela, ‘‘SnO2 and TiO2 Based Electronic Ceramics,’’ Cerâmica, 47, 136–43 (2001). Ceram. Soc., 47 [10] 484–90 (1964).
3 28
V. C. Sousa, J. W. Gomes, P. R. Bueno, M. R. Cássia-Santos, A. L. Araujo, B.-S. Chiou and M.-C Chung, ‘‘Effect of Copper Additive on the Microstruc-
E. R. Leite, E. Longo, and J. A. Varela, ‘‘Effect of the Addition of ZnO Seeds on ture and Electrical Properties of Polycrystalline Zinc Oxide,’’ J. Am. Ceram. Soc.,
the Electrical properties of ZnO-Based Varistors,’’ Mater. Chem. Phys., 80, 512–6 75 [12] 3363–8 (1992).
29
(2003). T. R. N. Kutty and N. Raghu, ‘‘Varistors Based on Polycristalline ZnO:Cu,’’
4
S. L. M. Brito, D. Gouvêa, and R. Ganzella, ‘‘Study of Polyacrylate Adsorp- Appl. Phys. Lett., 54 [18] 1796–8 (1989).
30
tion in a Commercial Varistor System: Characterization of the Physical–Chemistry Y.-S. Lee and T.-Y. Tseng, ‘‘Phase Identification and Electrical Properties in
Properties,’’ Cerâmica, 51, 30–6 (2005). ZnO–Glass Varistors,’’ J. Am. Ceram. Soc., 75 [6] 1636–40 (1992).
5 31
L. Hozer, Semiconductor Ceramics: Grain Boundary Effects. Ellis Horwood: Y.-S. Lee and T.-Y. Tseng, ‘‘Influence of Processing Parameters on the
Polish Scientific Publishers, West Sussex, UK, 1994. Microstructure and Electrical Properties of Multilayer-Chip ZnO Varistors,’’
6
C.-C. Hwang and T.-Y. Wu, ‘‘Synthesis and Characterization of Nanocrystal- J. Mater. Sci.: Mater. Electron., 6, 90–6 (1995).
32
line ZnO Powders by a Novel Combustion Synthesis Method,’’ Mater. Sci. Eng. B, Y.-S. Lee and T.-Y. Tseng, ‘‘Effects of Spinel Phase Formation in the Calci-
111 [2, 3] 197–206 (2004). nation Process on the Characteristics of ZnO–Glass Varistor,’’ J. Mater. Sci.:
7
T. K. Gupta, R. E. Tressler, G. L. Messing, C. G. Pantino, and R. H. Newhan Mater. Electron., 8, 115–23 (1997).
33
(eds.) Tailoring Multiphase and Composite Ceramics, pp. 439–507. Plenum Press, Y.-S. Lee and T.-Y. Tseng, ‘‘Correlation of Boundaries Characteristics with
New York, 1986. Electrical Properties in ZnO–Glass Varistors,’’ J. Mater. Sci.: Mater. Electron., 9,
8
T. K. Gupta, ‘‘Application of Zinc Oxide Varistors,’’ J. Am. Ceram. Soc., 73 [7] 65–76 (1998).
34
1817–40 (1990). J. V. Bellini, M. R. Morelli, and R. H. G. A. Kiminami, ‘‘Varistores de ZnO-
9
R. C. Buchanan, Ceramic Materials for Eletronics—Processing, Properties, and CuO-Vidro’’; PI0204858-2, 31/10/2002, INPI, Rio de Janeiro, Brazil.
35
Applications, pp. iii. Marcel Dekker Inc, New York, USA, 1986. J. V. Bellini, M. R. Morelli, and R. H. G. A. Kiminami, ‘‘Electrical Properties
10
M.-H. Wang, K.-A. Hu, B.-Y. Zhao, and N.-F. Zhang, ‘‘Electrical Charac- of Polycrystalline ZnO: CuO Obtained from Freeze-Dried ZnO1Copper (II)
teristics and Stability of Low Voltage ZnO Varistors Doped with Al,’’ Mater. Acetate Powders,’’ J. Mater. Sci.: Mater. Electron., 13 [8] 485–9 (2002).
36
Chem. Phys., 100 [1] 142–6 (2006). J. V. Bellini, M. R. Morelli, and R. H. G. A. Kiminami, ‘‘Ceramic
11
L. M. Levinson and H. R. Philipp, ‘‘Zinc Oxide Varistors—A Review,’’ Am. System Based on ZnO–CuO Obtained by Freeze-Drying,’’ Mater. Lett., 57,
Ceram. Soc. Bull., 65 [4] 639–46 (1986). 3775–8 (2003).
12 37
M. Matsuoka, ‘‘Nonohmic Properties of Zinc Oxide Ceramics,’’ Jpn. J. Appl. J. V. Bellini, M. R. Morelli, and R. H. G. A. Kiminami, ‘‘Physical Changes of
Phys., 10 [6] 736–46 (1971). Sintered Ceramics Obtained from Freeze-Dried ZnO1(CH3COO)2Cu  H2O Pow-
13
N. Yamaoka, M. Masuyama, and M. Fukui, ‘‘SrTiO3-Based Boundary-Layer ders,’’ Mater. Lett., 57, 3325–9 (2003).
38
Capacitor Having Varistor Characteristics,’’ Am. Ceram. Soc. Bull., 62 [6] 698–700 D. Sarid, Scanning Force Microscopy with Applications to Electric, Magnetic,
(1983). and Atomic Forces, pp. 129–51. Oxford University Press, New York, 1991.
14 39
M. F. Yan and W. W. Rhodes, ‘‘Preparation and Properties of TiO2 Va- S. N. Magonov and M. H. Whangbo, Surface Analysis with STM and AFM—
ristors,’’ Appl. Phys. Lett., 40, 536–7 (1982). Experimental and Theorical Aspects of Image Analysis, pp. 116–28. VCH,
15
V. Makarov and M. Tontelj, ‘‘Novel Varistor Material Based on Tungsten Weinheim , 1996.
40
Oxide,’’ J. Mater. Sci. Lett., 13, 937–9 (1994). E. Meyer, H. J. Hug, and R. Bennewitz, Scanning Probe Microscopy—The
16
S. A. Pianaro, P. R. Bueno, E. Longo, and J. A. Varela, ‘‘A New SnO2-Based Lab on a Tip, pp. 29. Springer, New York, 2003.
41
Varistor System,’’ J. Mater. Sci. Lett., 14 [10] 692–4 (1995). S. M. Gheno, H. L. Hasegawa, and P. I. Paulin Filho, ‘‘Direct Observation of
17
S. Li, F. Xie, F. Liu, J. Li, and M. A. Alim, ‘‘The Relation Between Residual Potential Barrier Behavior in Yttrium–Barium Titanate Observed by Electrostac-
Voltage Ratio and Microstructural Parameters of ZnO Varistors,’’ Mater. Lett., tic Force Microscopy,’’ Scr. Mater., 56 [6] 545–8 (2007).
42
59, 302–7 (2005). Y. Sato, J. P. Buban, T. Mizoguchi, N. Shibata, M. Yodogawa, T. Yama-
18
G. Q. Lu, Nanoporous Materials: Science and Engineering, pp. 700. Imperial moto, and Y. Ikuhara, ‘‘Role of Pr Segregation in Acceptor-State Formation at
College, London, 2001. ZnO Grain Boundaries,’’ Phys. Rev. Lett., 67, 106802 (2006).
19 43
M. Rittner, ‘‘Market Analysis of Nanostructured Materials,’’ Am. Ceram. R. Shao, S. V. Kalinin, and D. A. Bonnell, ‘‘Local Impedance Imaging and
Soc. Bull., 81 [3] 33–6 (2002). Spectroscopy of Polycrystalline ZnO Using Contact Atomic Force Microscopy,’’
20
M. Matsuoka, T. Masuyama, and Y. Iida, ‘‘Voltage Nonlinearity of Zinc Appl. Phys. Lett., 82, 1869 (2003).
44
Oxide Ceramics Doped with Alkali Earth Metal Oxide,’’ Jpn. J. Appl. Phys., 8, J. S. Vasconcelos, N. S. L. S. Vasconcelos, M. O. Orlandi, P. R. Bueno, J. A.
1275–6 (1969). Varela, E. Longo, C. M. Barrado, and E. R. Leite, ‘‘Electrostatic Force
21
H. C. Ling, M. F. Yan, and W. W. Rhodes, ‘‘Monolithic Device with Microscopy as a Tool to Estimate the Number of Active Potential Barriers in
Dual Capacitor and Varistors Functions,’’ J. Am. Ceram. Soc., 72 [7] 1274–6 Dense Non-Ohmic Polycrystalline SnO2 Devices,’’ Appl. Phys. Lett., 89, 152102
(1989). (2006). &

You might also like