You are on page 1of 9

Ceramics International 47 (2021) 1229–1237

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

CaO doped ZnO–Bi2O3 varistors: Grain growth mechanism, structure and


electrical properties
K. Hembram a, b, *, T.N. Rao a, R.S. Srinivasa b, A.R. Kulkarni b
a
Center for Nanomaterials, International Advanced Research Center for Powder Metallurgy & New Materials (ARCI), Balapur PO, Hyderabad, 500005, A.P., India
b
Department of Metallurgical Engineering and Materials Science, Indian Institute of Technology Bombay, Powai, Mumbai, 400076, India

A R T I C L E I N F O A B S T R A C T

Keywords: In this study, the impact of CaO doping on the microstructure, phase and electrical properties of the ZnO–Bi2O3
CaO doping system was studied. The powder was synthesized by the solution combustion route. Reduction in the crystallite
ZnO Varistor size of ZnO from 23 nm to 17 nm of the powder sample with CaO doping was noted. Conventional sintering was
Grain growth
used to consolidate the powders. Solid state reacted compositions revealed, from X-ray diffraction, formation of
Microstructure
Electrical properties
two phases (a) Ca4Bi6O13 at 750 ◦ C and (b) Ca0.89Bi3.11O5.56 at 900 ◦ C. An increase in density from 4.21 to 5.45
gcc− 1 and a decrease in grain size from 4.50 to 1.54 μm with increasing CaO doping were also observed. A grain
growth mechanism in the ZnO–Bi2O3–CaO system was proposed. The breakdown field increased up to 1 wt %
CaO (27 kV cm− 1) and after that decreased. Impedance spectroscopic investigation, in the temperature range of
100300 ◦ C and frequency range of 0.1 Hz -1M Hz, exhibited two relaxations. These were attributed to defects
formed. The two-stage activation energy of conductivity was noticed, which was similar to that observed in
complex ZnO varistor systems.

1. Introduction that grain size and grain boundary engineering play a crucial role in
achieving a varistor’s performance parameter and design. Bi2O3 melts
ZnO varistor is a non-linear device whose function is to sense and relatively at a lower temperature (650 ◦ C) and insulates the grain
control the surge voltage applied across the terminals; it is used exten­ boundaries of ZnO, which results in electrically insulating grain
sively for over-voltage protection [1]. Commercially ZnO varistor is boundary responsible for non-linear behavior [13]. Bi2O3 also encour­
fabricated using powder metallurgy processing involving mixing ages grain growth of ZnO, which deteriorates the properties of a varistor.
micron-sized ZnO powder with various additives ~4 mol % followed by Controlling the grain growth and microstructure is essential, and this is
pressing and liquid phase sintering at temperatures in the range normally achieved through the addition of dopants, for example, Sb2O3.
1100–1400 ◦ C, which results into grain size of 10–25 μm. Non-linear ZrO2 [14], WO3 [15], SiO2 [16,17], Al2O3 [18]B2O3 [19] CaO [20],and
behavior of a ZnO–Bi2O3 varistor is attributed to segregation of Bi2O3 rare earth oxides (Dy, Sc, Y) [21–23]. However, the influence of CaO
at the grain boundaries which make grain boundaries highly insulating addition on ZnO–Bi2O3 varistor characteristics was reported only
(1010-1012 Ωcm) and ZnO grains core become conducting (0.1–10 Ωcm), recently [20].
the substantial difference in impedance of grains and grain boundaries It was evident that the major portion of CaO reacts with Bi2O3 that
resulting in Double Schottky Barriers (DSB) formations [1–3]. A number forms the bismuth calcium oxide complex phase. However, due to the
of attempts were made to replace Bi2O3, with Pr6O11 [4], V2O5 [5], and complexity of the system in a previous study [20], the influence of CaO
CuO2 [6]; did not show significant enhancement in non-linear behavior addition on structure, grain growth mechanism and properties of
over commercial ones. Furthermore, various materials such as SnO2 [7], varistor was not fully understood [2,24,25]. In this study, an effort has
TiO2 [8], SrTiO3 [9], WO3 [10] and CuCaTiO3 [11] are also being been made to understand the influence of CaO addition on the micro­
investigated as an alternative to ZnO based varistor. However, structure and electrical properties of the ZnO–Bi2O3 system. Different
ZnO–Bi2O3 varistors are still commercially favored due to a combination concentration (X = 0.00, 0.25, 1.00 and 2.50 wt %) of CaO doped (95-X)
of excellent electrical properties at a lower cost [12]. It is well known ZnO–5Bi2O3 powders were made by solution combustion synthesis

* Corresponding author. Center for Nanomaterials, International Advanced Research Center for Powder Metallurgy & New Materials (ARCI), Balapur PO,
Hyderabad, 500005, A.P., India.
E-mail address: kaliyan@arci.res.in (K. Hembram).

https://doi.org/10.1016/j.ceramint.2020.08.242
Received 24 June 2020; Received in revised form 14 August 2020; Accepted 26 August 2020
Available online 28 August 2020
0272-8842/© 2020 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

(SCS). The phase, microstructure, electrical and dielectric properties of 2.3. I–V measurement
the sintered samples are investigated. A grain growth mechanism is
proposed for the system. I–V characteristics were measured on a circumference insulated
sintered pellet samples using by 5 KV DC (Rectifiers and Electronics,
2. Experimental and characterizations Delhi, India) breakdown tester at 25 ◦ C. The sample holder, made in-
house, was used for measurements. The breakdown field (Bf) was
2.1. Powder synthesis and consolidations calculated at a current density of 1 mA/cm2. The coefficient of nonlin­
earity (α) was calculated: α = (1/logE2-logE1); E2 > E1, where E1 is the
Zinc nitrate hexahydrate (99% Alfa-Asear make), Calcium nitrate field at current density J1 (0.1 mA/cm2) whereas E2 is the field at current
tetrahydrate (99.9%, Alfa-Asear make), and Cobalt nitrate hexahydrate density J2 (1 mA/cm2).
(>99 %, Sigma make) were mixed in double-distilled water (Sol.-1). To
get complete transparent solution Bismuth nitrate pentahydrate (99%, 2.4. Impedance and dielectric measurements
Sigma make) was heated (100–150 ◦ C) in dilute nitric acid (99.9%,
Renkem make) (Sol.-2). The Sol.-2 and to Sol.-1 were mixed, followed by Complex impedance and dielectric properties measurements were
heating around 100 ◦ C to make a transparent solution. 0.23 mol% of performed on pellets (1 mm thick and 8 mm diameter) with silver
sucrose (98%, Renkem make) was added to the final solution. The electrodes coated on both the sides of the pellet, by impedance analyzer
resulting solution was heated to a temperature in the range of (Novocontrol, Alpha) in the frequency range of 0.1 Hz -1 MHz in the
180–200 ◦ C and stirred until it catches fire. The CaO doped (95-X) wide range of temperatures at 25 ◦ C intervals with an amplitude bias
ZnO–5Bi2O3-XCaO (X = 0.00, 0.25, 1.00, and 2.50) powders were syn­ voltage of 1Vrms− 1. The measurements were performed in the cooling
thesized and calcined at 750 ◦ C for 1 h. The powders were consolidated cycle for better temperature stability. The temperature was controlled
into a 10 mm pellet at the pressure of 150 MPa in a die. The pellets were through a Eurotherm temperature controller. Impedance data were
sintered at two-step sintering (as depicted in Figure S-1): (1) heating up analyzed by using software ZSimpWin 3.21. The impedance data were
to 900 ◦ C and hold it for 0.5 h, then decrease to 825 ◦ C and hold it again analyzed in different formalisms for detailed interpretation.
for 4 h followed by cooling (designated as S-1). (2) Heating up to
1000 ◦ C and hold it for 0.5 h, then decrease to 825 ◦ C and hold it for 4 h
3. Results and discussion
followed by cooling (designated as S-2). (3) Heating up to 1100 ◦ C and
hold it for 0.5 h, then decrease to 825 ◦ C and hold it for 4 h followed by
3.1. Phase and composition analysis of the powder samples
cooling (designated as S-3). The rate of cooling and heating was kept
2 ◦ C min− 1 up to 725 ◦ C and then furnace cooling for all the pellet
A comparison of XRD patterns (at room temperature) of undoped
samples.
(UDZB) and CaO doped (CDZB) powder specimens heated at 750 ◦ C-1h
is shown in Fig. 1. ZnO Wurtzite (JCPDS card no. # 13–7122) and
2.2. Phase and microstructure evaluation
α-Bi2O3 (JCPDS card no. # 52–1007) phases were detected in UDZB
powder sample. Phases of ZnO, Ca4Bi6O13 (JCPDS card no. # 48–0217)
Phases in the powder and sintered samples were studied using
were noticed in the CDZB powder samples. The formation of phases in
Powder X-ray Diffraction (Bruker D8 advance) using CuKα radiation.
CDZB samples is explained elsewhere [27] using ½Bi2O3–CaO phase
Compositions of calcined powders were checked by ICP-OES (Varian
Liberty) analysis. The powders were dissolved in nitric acid to make a
0.1 M solution, and the percentage of elements present in the samples
was estimated. Thermal analysis of the sample was carried out by STA
449 Jupiter (Netzsch GmbH, Germany) with a heating rate of 10 ◦ Cmin-1
from room temperature (RT)-1200 ◦ C in air. Green pellet densities were
measured by weight and dimension analysis technique, whereas sin­
tered samples densities measured by the Archimedes method.
Morphology and microstructure of the powder and sintered specimens
were analyzed by FE-SEM (S-4300-SE/N, Hitachi), and qualitative
composition (elemental analysis) of both the samples were analyzed by
EDS (EDAX, USA) attached to the unit. The average ZnO grain size was
measured by Olympus analySIS software using the following relation­
ship: G = 1.56L, where G is the average ZnO grain size, and L is the
average grain boundary intercept length [26]. The microstructure and
morphology of the powder and sintered samples were studied by TEM
(200 KV, FEI). The powder samples were sonicated for 10–15 min in
ethanol before putting 2–3 drops on a copper grid. Bright-field TEM
images were used for the crystallite size estimation of the powders.
Qualitative compositional (elemental analysis) of the specimens were
also carried out by EDS attached to TEM. SAED studies for the sintered
samples were carried out by TEM attached to the unit. The thickness of
the TEM sample for the sintered sample was reduced using the Isomet
cutter. A 3 mm dimple was prepared by GATAN, 601 ultrasonic cutter on
200 μm thick specimen. The thickness of the sample was reduced to
approximately 60 μm by 2000 no. emery paper flowed by ion milled for
9–10 h at an energy of 4 KeV and angle at 11◦ till a tiny hole was
generated (GATAN, Mod. No. 691). The TEM sample was cleaned (1.5
KeV) for 10–15 min just before analysis. The ’d’ spacing is calculated
from the formula R d = Lλ, where R is the spot spacing, L is the camera Fig. 1. XRD patterns of UDZB and CDZB powder samples calcined at 750 ◦ C for
length, and λ is the electron beam wavelength. 1 h.

1230
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

could be an excellent candidate for grain boundary phase in ZnO varistor


since the compound contains more oxygen as a result of which increase
in barrier strength of the device is possible, and that would lead to
improvement in the properties [29,30]. The detailed study of ZBC (X =
1.00) was carried out because it demonstrated better varistor properties
than other compositions.
Fig. 2(a) shows a comparison of XRD patterns for the powder (X =
1.000) samples calcined at different temperatures in the range of
750–950 ◦ C for 1 h to understand the grain boundary phase formations.
Interestingly, when as-synthesized powder was calcined at ≤ 650 ◦ C for
1 h, a single Wurtzite ZnO phase was obtained (Figure S-2(a)). However,
when the powder was heated to ≥ 750 ◦ C for 1 h, the Ca4Bi6O13 phase
along with ZnO and ≥ 950 ◦ C for 1 h Ca0.89Bi3.11O5.56 and Ca4Bi6O13
phase along with ZnO was observed (Fig. 2(b)).
For a better understanding of the phase formation of the materials
and supplement the XRD results, thermal analysis of the gel was carried
out. The results are shown in Fig. 2(c). Endothermic peak at 395 ◦ C is an
indication of completion of combustion and formation of ZnO phase;
two endothermic peaks at 750 and 865 ◦ C correspond to Ca4Bi6O13 and
Ca0.89Bi3.11O5.56 phases which suggest that in both cases the secondary
phase formation take place at a temperature below 900 ◦ C. Weight loss
of 30 wt% up to 395 ◦ C is due to the burning of organic and nitrate
compounds, and weight loss of ~10 wt% up to 1200 ◦ C is due to
evaporation of bismuth species [24,43].
The compositions of all the powder samples calcined at 750 ◦ C for 1 h
are given in Table 1. The compositions of the powder samples were
within the experimental error after the synthesis and calcination pro­
cess. As discussed above, the XRD of the powders and thermal analysis of

Fig. 2. (a) XRD patterns of X = 1.00 powder samples calcined at different


Table 1
temperatures for 1 h and (b) Thermograph of a mixture of (Zn, Bi, Ca nitrates,
Compositional analysis by ICP-OES of powder samples calcined at 750 ◦ C for 1 h.
and fuel) gel of X = 1.00 sample.
Sample ID Composition in wt. %

diagram. Less than 10 mol. % CaO yields Ca2Bi4O12 and Bi2O3 phases, ZnO Bi2O3 CaO
and above 10 mol.% Ca4Bi6O13,Ca3Bi8O15, and Ca2Bi2O4 phases are X = 0.00 95.20 4.79 –
reported [27,28]. The composition used for the present study favors the X = 0.25 95.15 4.50 0.21
Ca4Bi6O13 phase formation because the percentage of CaO is marginally X = 1.00 94.25 4.80 0.80
X = 2.50 92.75 4.70 2.54
higher than 20 mol. % of Bi2O3 and CaO phases [27,28]. The Ca4Bi6O13

Fig. 3. Bright-field TEM images; insets are crystallite size histogram of powder samples having different compositions; calcined at 750 ◦ C for 1 h.

1231
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

Fig. 4. (a) Impact of CaO addition on the crystallite size of powder samples. (b) Influence of calcination temperature on crystallite size of X = 1.00 powder samples.

Fig. 5. (a) A comparison of XRD patterns for UDZB and CDZB sintered samples and (b) zoom in view of grain boundary phase.

Fig. 6. (a) Backscattered-SEM (thermally-etched) images; insets are grain size distribution histogram of sintered samples. (b) Influence of CaO doping on the grain
size of sintered samples.

the gel, liquid phase formation at temperature ~750 ◦ C is observed. A 3.2. Morphological studies of powder samples
little deviation of the composition could be due to segregation resulting
from liquid phase formation [31]. These powder samples were com­ Fig. 3 shows the bright-field TEM images; the insets are the crystallite
pacted and sintered into pellets without any further purification and size distribution histograms of the powder samples calcined at 750 ◦ C for
modification. 1 h. An impact of CaO doping on the crystallite size of ZnO for the
powder sample is shown in Fig. 4(a) and the influence of calcination

1232
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

temperature on the crystallite size of ZnO for ZBC (X = 1.00) powder 3.3.2. Microstructure analysis
sample is shown in Fig. 4(b). FE-SEM images of polished and thermally etched UDZB and CDZB
A decrease in crystallite size with increasing CaO doping was samples sintered by S-2 are depicted in Fig. 6 (a). The insets of Fig. 6 (a)
observed until X = 1.00. Thereafter, a slight increase can be seen clearly. show ZnO grain size distribution histogram. The influence of CaO
Uniform distribution and decrease in crystallite size with CaO doping doping on grain size is depicted in Fig. 6 (b). To confirm the liquid phase
are due to grain boundary phase formation of Ca4Bi6O13 at the surface of distribution throughout the samples, FESEM and EDS were carried out
ZnO. Bright-field TEM images (insets are particle size distribution and on the fracture surface of all sintered samples, as shown in Figure S-3(a)
dark field TEM images) of ZBC (X = 1.00) powder sample calcined at and (b). It is clearly seen that the liquid phase is homogeneously
different temperatures are shown in Figure S-2(b). As expected, it can be distributed throughout the microstructure, and a thin layer is covered on
seen that crystallite size increased from 16 nm (550 ◦ C) to 650 nm the ZnO grains, and the bulk of liquid phases are segregated at triple
(850 ◦ C) with increasing temperature. However, an abnormal increase junctions and grain boundaries. The liquid phase in the samples
in grain size for the powder calcined at 850 ◦ C for 1 h is attributed to increased as the concentration of CaO is increased. Decrease in grain size
liquid phase formation, as discussed earlier. from 4.45 to 1.54 μm and narrow grain size distribution with larger CaO
addition till X = 1.00 sintered sample, thereafter increase in grain size
due to the formation of a more liquid phase. It may be noted that the
3.3. Phase and microstructure studies of sintered samples homogenous distribution of grain boundary phase Ca4Bi6O13 and
Ca0.89Bi3.11O5.56 throughout the microstructure of the sintered samples.
3.3.1. Phase analysis The round shape of ZnO grain suggests that a liquid phase is formed
Fig. 5 (a–b) show the comparison of XRD patterns with zoomed-in during sintering, which is well known for ZnO–Bi2O3 based varistor
view on 2θ values in the range of 25–31◦ of UDZB and CDZB samples composition [33,34]. The effect of sintering temperature on ZnO grain
sintered by S-2. The phases of α- Bi2O3 (JCPDS card no. # 41–1449) and size (X = 1.00) is shown in Figure S-4. An increase in grain size and
ZnO (JCPDS card no. # 13–7122) were clearly noticed in UDZB sintered liquid phase with CaO doping are observed. Grain growth retardation is
sample whereas the secondary phase of Ca4Bi6O13 (JCPDS card no. # attributed to an increase in viscosity of liquid phase and Zener pining by
48–0217) and Ca0.89Bi3.11O5.56 (JCPDS card no. # 40–0317) along with grain boundary phases of Ca4Bi6O13 and Ca0.89Bi3.11O5.56. Grain growth
ZnO matrix phase were observed in case of CDZB sintered samples. The of ZnO with increasing sintering temperature, which is attributed to the
intensity of XRD peaks for Ca0.89Bi3.11O5.56 phase is higher than the formation of a more liquid phase as expected [35].
Ca4Bi6O13 phase of X = 1.00 compared to X = 2.50 CDZB, suggesting
that Ca0.89Bi3.11O5.56 phase favored the lower concentration of CaO. 3.3.3. TEM, EDS and SAED studies of sintered specimen
Phase formation of these sintered samples also could be explained by the Fig. 7 (a) and (b) display the bright-field and dark-field TEM images
1/2Bi2O3–CaO phase diagram, as discussed in the earlier section [27, of X = 1.00 sintered sample, respectively. The EDS of grain boundary
28]. The Ca4Bi6O13 and Ca0.89Bi3.11O5.56 phases appeared in the tem­ phase (white oval, as shown in dark-field TEM image) is shown in Fig. 7
perature range 750–900 ◦ C (Fig. 6); the present findings are in good (c). HR-TEM image of the grain boundary interface is shown in Fig. 7(d).
agreement with earlier studies [27,28]. However, the influence of these The SAED patterns at the inter-granular area (oval marked in Fig. 7(b))
secondary phases in ZnO varistor is hitherto unreported, except in our from zone axes [001] and [011] of the sample are shown in Fig. 7 (e) and
previous study [20]. It is believed that a few monolayer thin oxygen rich (f), respectively. Spot EDS analysis from the white location confirms that
grain boundary phase between two ZnO grains is good enough to exhibit the grain boundary contains Bi, Ca, Zn and O species. HR-TEM image
ZnO varistor action, which helps in improving tunneling and switching shows a ~2 nm amorphous grain boundary similar to that of commercial
of charge carriers through the double Schottky barrier (DSB), resulting ZnO varistors [42,52]. The SAED patterns (hkl) planes corresponding to
in a better coefficient of nonlinearity [30,32]. Therefore, small amount (020), (220) and (220) from zone axis [001], and (200), (111)) and
of Ca4Bi6O13 and Ca0.89Bi3.11O5.56 grain boundary phases are ideal for (111) from zone axis [011] which turns out be a FCC structure (a =
high performance ZnO varistor.

Fig. 7. (a) Bright-field and (b) Dark-field TEM images at the same location; (c) EDS of grain boundary phase confirming presence of Bi, Ca, O and Zn elements (white
oval), (d) HR-TEM image at the inter-granular area, (e) SAED pattern from zone axes [001] and [011], confirming an FCC structure (a = ~0.55 nm) of
Ca0.89Bi3.11O5.56 phase of X = 0.00 sintered sample.

1233
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

Fig. 8. A grain growth mechanism comparison with commercial and CaO doped ZnO varistor system.

~0.55 nm). The grain boundary phase is confirmed to be


Ca0.89Bi3.11O5.56 (JCPDS No # 40–0317) phase. The phase confirmation
by SAED is in good agreement with the XRD results.

3.4. Understanding the grain growth mechanism

Before elucidating the grain growth mechanism of the CaO doped


ZnO–Bi2O3 system, it is necessary to revisit the traditional Sb2O3 added
ZnO–Bi2O3 system and compare it with this new system. It is well known
that the densification in a ZnO varistor is due to the liquid phase sin­
tering. Bismuth oxide liquid phase is arranged by a diffusion process due
to solution precipitation and grain boundary reactions, as shown in
Fig. 8(a) [35–37]. Generally, Sb2O3 is added as a grain growth inhibitor,
which forms spinel (Zn7Sb2O12) and pyrochlore (Bi3Zn2Sb3O14) during
heating, creating the Zener drag and inhibiting ZnO grain growth [38].
Grain size is inversely proportional to the concentration of the spinel
phase. It is also evident that a small amount of Sb2O3 (< 75 ppm) pro­
Fig. 9. Influence of CaO addition on I–V characteristics; inset is the influence of duces a coarse-grained microstructure rather than smaller grain in ZnO
I–V characteristics on sintering temperature for X = 0.00 samples. based varistor system [39]. In this system, ZnO grain growth retardation
is due to Ca4Bi6O13 and Ca0.89Bi3.11O5.56 phases at the grain boundary of
ZnO grain and their interface. The ZnO grain growth inhibition is
attributed to pinning by a small solute particle (~ 5 nm) and an increase

Fig. 10. (a) Normalized Cole-Cole plots and the data are fitted to a two parallel RC circuit model (inset); (b) Electrical modulus plots and the data are fitted into two
semicircles, and insert is the frequency dependent Z′′ and M′′ plots for UDZB and CDZB sintered samples.

1234
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

Table 3
Electrical, dielectric (200 ◦ C and 50 Hz) properties and Ea of sintered specimens.
Sample ID Cgb1 (pF) Rgb1 (MΩ-cm) Cgb2 (pF) Rgb2 (MΩ-cm) ε’ (200 ◦ C, 50 Hz) Tan δ (200 ◦ C & 50 Hz) Ea. (eV)

Region I Region I

X = 0.00 089 0.136 102 3.21 1163 ± 10 6.13 ± 0.5 0.56 1.01
X = 0.25 101 0.055 104 3.31 1037 ± 10 6.66 ± 0.5 0.62 0.79
X = 1.00 136 0.325 514 6.11 0444 ± 10 5.45 ± 0.5 0.61 0.85
X = 2.50 154 4.610 554 22.10 0395 ± 10 4.65 ± 0.5 0.84 0.97

in viscosity of the liquid phase, as shown in Fig. 8 (b). Neither inverse properties compared to other samples; this is due to the small grain size
grain boundary nor spinel and pyrochlore phase formation was observed and narrow grain size distribution of ZnO. The X = 1.00 sample was
in any of the present sintered samples. For SiO2 doped ZnO varistor, the sintered at different temperatures to optimize sintering parameters. The
ZnSiO4 phase is responsible for increasing the viscosity of the liquid breakdown field (Bf) of 26.6 kVcm− 1 with a moderate coefficient of
phase and inhibiting the ZnO grain growth [40,41]. In the present study, nonlinearity of 7 was obtained sample sintered by S-1 sample. Excep­
even a small (X = 0.25) doping of CaO also retarded grain growth, and tional high breakdown voltage is due to the uniform distribution of
the effect is significant. Therefore, it can be inferred that the grain
growth retardation is more dominated by an increase in the liquid
Table 2
phase’s viscosity rather than Zener pining by secondary grain boundary
The grain size, density and electrical properties of sintered pellet specimens.
phases. The impression of the present study is that the grain growth
Sample ID Grain Size Density Breakdown field Coefficient of
pinning for CaO doped ZnO varistor is similar to that of SiO2 [41,42].
(μm) (gcc− 1) (kVcm− 1) nonlinearity

X = 0.00 4.50 ± 5.23 ± 01.72 ± 0.50 02.90 ± 0.5


3.5. Electrical properties studies of sintered samples 0.50 0.05
X = 0.25 4.40 ± 5.40 ± 02.70 ± 0.50 03.15 ± 0.5
3.5.1. I–V characteristics 0.50 0.05
Fig. 9 shows the comparison I–V characteristic of UDZB and CDZB X = 1.00 1.54 ± 5.32 ± 12.88 ± 1.00 12.00 ± 1
(S-2) 0.25 0.05
sintered samples; inset is the X = 1.00 sample sintered at different
S-3 3.91 ± 5.01 ± 02.60 ± 0.50 08.58 ± 1
temperatures. The summary of varistor properties for all the samples is 0.50 0.08
given in Table 2. An increase in breakdown field with increasing CaO S-1 1.35 ± 5.21 ± 26.60 ± 5.00 06.65 ± 1
addition till X = 1.00 12.88 kVcm− 1 thereafter decrease was observed. 0.25 0.05
The coefficient of nonlinearity as high as 12 was obtained for X = 1.00 X = 2.50 4.75 ± 5.43 ± 07.78 ± 3.00 3.00 ± 0.5
0.50 0.05
sintered sample. The X = 1.00 sample exhibited better varistor

Fig. 11. Frequency dependent (a) dielectric constant, (b) tan δ (inset is the tan δ at higher frequencies region) at 200 ◦ C; (c) Temperature dependent exponent and
(d) Arrhenius plot of DC conductivity for UDZB and CDZB sintered samples.

1235
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

secondary phases of Ca4Bi6O13 and Ca0.89Bi3.11O5.56 at the grain that the conduction of charge carrier is due to long-range hopping [51].
boundary and interface area and smaller ZnO grain size. The Bf is an It is clearly seen that two stages of Ea in the Arrhenius plot, the Ea is in
order of magnitude higher than the commercial ZnO varistor [32,43]. the range 0.80–1.01 eV for region II, the Ea is in the range 0.30–0.60 eV
for region I, the trend of these values are similar to that of complex ZnO
3.5.2. Impedance and electrical modulus studies varistor system [44,52,53]. It may be noted that the Ea values for these
Fig. 10 (a) shows normalized impedance plots fitted to a two parallel compositions are very much close to the complex varistor composition.
RC circuit model, and (b) electrical modulus and inset is the frequency It is known that all traps at grain boundaries are due to Bi2O3 and are
dependent Z′′ and M′′ plot at 200 ◦ C for sintered (S-2) samples. The responsible for the varistor effect; other dopants are added to enhance
impedance plots are elongated semicircular shaped whereas electric properties of the ZnO–Bi2O3 based varistors [54].
modulus plots show two semicircles for all the samples, thereby justi­
fying the two parallel RC circuit model for data interpretation. The Z′′ 4. Conclusions
versus f plots show one relaxation peak whereas M′′ versus f plots show
two relaxation peaks, one at lower (500 Hz) frequency regime and Understanding of CaO doping on microstructure, phase and prop­
another at higher frequency region (2 × 105 Hz) (lower concentration erties in a simple ZnO–Bi2O3 system was established. A reduction in
CaO) sintered samples. However, as the concentration of CaO increased crystallite size from 23 to 17 nm for the powder was observed with CaO
to X = 2.50, the second relaxation peak moves beyond 1 MHz due to doping. The phase formation of Ca4Bi6O13 and Ca0.89Bi3.11O5.56 at a
small capacitance values. It is to be noted that the same trend was temperature of 750 ◦ C and 900 ◦ C was observed. An increase in density
observed in the complex composition of ZnO varistor as well [20,44,45]. from 4.21 to 5.45 gcc− 1 and a decrease in grain size from 4.50 to 1.54 μm
The electrical properties of fitted data are shown in Table 3. An increase due to CaO doping were discerned. This is due to grain growth retar­
in grain boundary resistivity (Rgb1 and Rgb2) with increasing CaO doping dation caused by an increase in viscosity of the liquid phase and Zener
was observed. An increase in capacitance (Cgb1 and Cgb2) values with pinning mechanism due to the presence of Ca4Bi6O13 and
increasing CaO concentration was noted. Resistivity and capacitance Ca0.89Bi3.11O5.56 grain boundary secondary phases. Increase in break­
increased are attributed to decreases in carrier concentrations of the down field having CaO content till 1wt.% (as high as 27 Kvcm− 1)
samples, as reported in our previous studies [46,47]. thereafter decrease was observed. The coefficient of nonlinearity as high
as 12 for 1 wt% CaO sample was obtained without transition metal oxide
doping. Two defect relaxations in a wide range of temperature and
3.6. Dielectric studies
frequency ranges were observed similar to that of commercial ZnO
varistor. In the case of ≥ 1 wt% samples, the carrier traps development
To correlate the I–V and impedance results, and to appreciate the
are due to the presence of Ca4Bi6O13 and Ca0.89Bi3.11O5.56 grain
influence of CaO doping on the electrical properties of sintered samples,
boundary phases. Two stage activation energy of conductivity was dis­
the complex impedance raw data are analyzed. Complex impedance
cerned; 0.30–0.60 eV (Region I) and 0.80–1.01 eV (Region I). The
data are analyzed using the relationship [48,49]:
activation energy of DC conductivity values is similar to that of the
[ ]− 1 [ ]− 1
1 1 complex ZnO varistor system.
Z * = Z ’ − Z ’’ = + jωCg + + jωCgb1
Rg Rgb1
+ [1/ (Rgb2 + Ao (jω)n )]− 1
(1) Declaration of competing interest

Where ω and n are the angular frequency and depression of impedance The authors declare that they have no known competing financial
plots, the capacitance and resistance of grain boundary type 1 (GB1) and interests or personal relationships that could have appeared to influence
grain boundary type 2 (GB2) are assigned as Cgb1, Rgb1, Cgb2 and Rgb2. the work reported in this paper.
Relaxation takes place when ωRC = 1. GB1 is assigned to resistance
having lower values, whereas GB2 is to the highest resistance compo­ Acknowledgments
nent. Fig. 11 (a–b) displays the frequency dependent tan δ and ε′ for
different compositions of sintered (S-2) specimens at 200 ◦ C and 50 Hz, One of the author like to extend gratitude to DST-SERB, Govt. of
are summarized in Table 3. Decrease in dielectric constant with India (Grant No. SB/EMEQ-221/2013), for partially supporting this
increasing frequency; the value remains almost constant beyond 105 Hz. work. Director ARCI is duly acknowledged for giving permission to
Tan δ decreases with increasing frequency around 103 Hz, and beyond publish the manuscript.
that, it remains constant. A decrease in dielectric constant with
increasing CaO doping was observed [Table- 3], it is due to the fact that Appendix A. Supplementary data
dielectric constant decreases with a decrease in grain size [50]. It may be
noted that a decrease in Tan δ with increasing CaO addition is due to a Supplementary data to this article can be found online at https://doi.
reduction in carrier concentration of the samples. org/10.1016/j.ceramint.2020.08.242.

3.7. Conductivity studies References

[1] M. Matsuoka, Nonohmic properties of zinc oxide ceramics, Jpn. J. Appl. Phys. 10
The activation energy of DC conductivity is calculated using a rela­ (1971) 736.
tionship: [2] T.K. Gupta, Application of zinc oxide varistors, J. Am. Ceram. Soc. 73 (1990)
1817–1840.
Ea [3] D. Clarke, Varistor ceramics, J. Am. Ceram. Soc. 82 (1999) 485–502.
σ dc = σo exp (2) [4] S. Bachir, C. Sandouly, J. Kossanyi, J.C. Ronfard-Haret, Rare earth-doped
kT
polycrystalline zinc oxide electroluminescent ceramics, J. Phys. Chem. Solid. 57
(1996) 1869–1879.
Where σo is the pre-exponential factor, Ea is the activation energy, and k
[5] J.K. Tsai, T.B. Wu, Non-ohmic characteristics of ZnO-V2O5 ceramics, J. Appl. Phys.
is the Boltzmann’s constant. Frequency dependent AC conductivity are 76 (1994) 4817–4822.
fitted to the Jonscher’s law at different temperature and extracted the [6] T.R.N. Kutty, N. Raghu, Varistors based on polycrystalline ZnO:Cu, Appl. Phys.
exponent “n” and DC conductivity values of the samples [51]. The “n” Lett. 54 (1989) 1796–1798.
[7] P.R. Bueno, S.A. Pianaro, E.C. Pereira, L.O.S. Bulhões, E. Longo, J.A. Varela,
and Arrhenius plots for DC conductivity are shown in Fig. 11 (c–d). The Investigation of the electrical properties of SnO2 varistor system using impedance
“n” values are in the range of 0.90–0.30 at 125–450 ◦ C, which suggests spectroscopy, J. Appl. Phys. 84 (1998) 3700–3705.

1236
K. Hembram et al. Ceramics International 47 (2021) 1229–1237

[8] M.F. Yan, W.W. Rhodes, Preparation and properties of TiO2 varistors, Appl. Phys. [31] J.-R. Lee, Y.-M. Chiang, Bi segregation at ZnO grain boundaries in equilibrium with
Lett. 40 (1982) 536–537. Bi2O3-ZnO liquid, Solid State Ionics 75 (1995) 79–88.
[9] T.R.N. Kutty, S. Philip, Low voltage varistors based on SrTiO3 ceramics, Mater. Sci. [32] F. Greuter, Electrically active interfaces in ZnO varistors, Solid State Ionics 75
Eng., B 33 (1995) 58–66. (1995) 67–78.
[10] V. Makarov, M. Trontelj, Novel varistor material based on tungsten oxide, J. Mater. [33] E. Olsson, G.L. Dunlop, The effect of Bi2O3 content on the microstructure and
Sci. Lett. 13 (1994) 937–939. electrical properties of ZnO varistor materials, J. Appl. Phys. 66 (1989)
[11] S.Y. Chung, I.D. Kim, S.J. Kang, Strong non-linear current-voltage behaviour in 4317–4324.
perovskite-derivative calcium copper titanate, Nat. Mater. 3 (2004) 774–778. [34] J.P. Gambino, W.D. Kingery, G.E. Pike, L.M. Levinson, H.R. Philipp, Effect of heat
[12] J. Li, S. Li, P. Cheng, M. Alim, Advances in ZnO–Bi2O3 based varistors, J. Mater. treatments on the wetting behavior of bismuth-rich intergranular phases in ZnO:Bi:
Sci. Mater. Electron. 26 (2015) 4782–4809. Co varistors, J. Am. Ceram. Soc. 72 (1989) 642–645.
[13] D.R. Clarke, Grian-boundary segregation in a commercial ZnO-based varistor S, [35] T.M. Shaw, Liquid phase redistribution dueing liquid-phase sintering, J. Am.
J. Appl. Phys. 50 (1979) 6829–6832. Ceram. Soc. 69 (1986) 27–34.
[14] Y. Akiyama, M. Takada, A. Fukumori, Y. Sato, S. Yoshikado, Effect of ZrO2 addition [36] W.D. Kingery, Introduction to Ceramics, 1960.
on tolerance characteristics of the electrical degradation of ZnO varistors, Electr. [37] D. Dey, R.C. Bradt, Grain growth of ZnO during Bi2O3 liquid-phase sintering, J. Am.
Eng. Jpn. 179 (2012) 11–19. Ceram. Soc. 75 (1992) 2529–2534.
[15] X. Xiao, L. Zheng, L. Cheng, T. Tian, X. Ruan, M. Podlogar, et al., Influence of WO3- [38] T. Senda, R.C. Bradt, Grain growth of zinc oxide during the sintering of zinc
doping on the microstructure and electrical properties of ZnO–Bi2O3 varistor oxide—antimony oxide ceramics, J. Am. Ceram. Soc. 74 (1991) 1296–1302.
ceramics sintered at 950◦ C, J. Am. Ceram. Soc. 98 (2015) 1356–1363. [39] S. Bernik, N. Daneu, A. Rečnik, Inversion boundary induced grain growth in TiO2
[16] H. Bai, M. Zhang, Z. Xu, R. Chu, J. Hao, H. Li, et al., The effect of SiO2 on electrical or Sb2O3 doped ZnO-based varistor ceramics, J. Eur. Ceram. Soc. 24 (2004)
properties of low-temperature-sintered ZnO–Bi2O3–TiO2–Co2O3–MnO2-based 3703–3708.
ceramics, J. Am. Ceram. Soc. 100 (2017) 1057–1064. [40] T. Takemura, M. Kobayashi, Y. Takada, K.E.N. Sato, Effects of antimony oxide on
[17] Z-h Wu, J-h Fang, D. Xu, Q-d Zhong, L-y Shi, Effect of SiO2 addition on the the characteristics of ZnO varistors, J. Am. Ceram. Soc. 70 (1987) 237–241.
microstructure and electrical properties of ZnO-based varistors, Int. J. Miner. [41] H.-Q. Shao, X.-H. Gao, Z.-C. Cao, Effect of annealing on phase structure and
Metall. Mater. 17 (2010) 86–91. degradation of a zinc oxide varistor with Si-additive, J. Eur. Ceram. Soc. 17 (1997)
[18] L.-H. Cheng, L.-Y. Zheng, L. Meng, G.-R. Li, Y. Gu, F.-P. Zhang, et al., Electrical 55–59.
properties of Al2O3-doped ZnO varistors prepared by sol–gel process for device [42] K. Eda, M. Inada, M. Matsuoka, Grain growth control in ZnO varistors using seed
miniaturization, Ceram. Int. 38 (Supplement 1) (2012) S457–S461. grains, J. Appl. Phys. 54 (1983) 1095–1099.
[19] L. Cheng, G. Li, K. Yuan, L. Meng, L. Zheng, Improvement in nonlinear properties [43] P. Durán, F. Capel, J. Tartaj, C. Moure, A strategic two-stage low-temperature
and electrical stability of ZnO varistors with B2O3 additives by nano-coating thermal processing leading to fully dense and fine-grained doped-ZnO varistors,
method, J. Am. Ceram. Soc. 95 (2012) 1004–1010. Adv. Mater. 14 (2002) 137–141.
[20] K. Hembram, T.N. Rao, M. Ramakrishana, R.S. Srinivasa, A.R. Kulkarni, Influence [44] M. Andres-Verges, A.R. West, Impedance and modulus spectroscopy of ZnO
of CaO doping on phase, microstructure, electrical and dielectric properties of ZnO varistors, J. Electroceram. 1 (1997) 125–132.
varistors, J. Alloys Compd. 817 (2020) 152700. [45] L.M. Levinson, H.R. Philipp, AC properties of metal-oxide varistors, J. Appl. Phys.
[21] G-h Chen, J-l Li, Y. Yang, C-l Yuan, C-r Zhou, Microstructure and electrical 47 (1976) 1117–1122.
properties of Dy2O3-doped ZnO–Bi2O3 based varistor ceramics, Mater. Res. Bull. 50 [46] M. Barsoum, M. Barsoum, Fundamentals of Ceramics, CRC press, 2002.
(2014) 141–147. [47] Jacobs K.F. A. Kröger, North-Holland Publishing Company - Amsterdam/London
[22] P. Williams, O.L. Krivanek, G. Thomas, M. Yodogawa, Microstructure-property 1973 American Elsevier Publishing Company, Inc. - New York 313 Seiten,
relationships of rare-earth–zinc-oxide varistors, J. Appl. Phys. 51 (1980) zahlreiche Abbildungen und Tabellen, Kunstleder Preis Dfl. 70.00. Krist. Tech., in:
3930–3934. The Chemistry of Imperfect Crystals, Preparation, Purification, Crystal Growth and
[23] D. Xu, X. Cheng, H. Yuan, J. Yang, Y. Lin, Microstructure and electrical properties Phase Theory, 2nd Revised Edition 1-9, 1974. K67-K68
of Y(NO3)3⋅6H2O-doped ZnO–Bi2O3-based varistor ceramics, J. Alloys Compd. 509 [48] K. Hembram, T.N. Rao, M. Ramakrishna, R.S. Srinivasa, A.R. Kulkarni, A novel
(2011) 9312–9317. economical grain boundary engineered ultra-high performance ZnO varistor with
[24] M. Peiteado, M.A. de la Rubia, M.J. Velasco, F.J. Valle, A.C. Caballero, Bi2O3 lesser dopants, J. Eur. Ceram. Soc. 38 (2018) 5021–5029.
vaporization from ZnO-based varistors, J. Eur. Ceram. Soc. 25 (2005) 1675–1680. [49] R. Tang, C. Jiang, W. Qian, J. Jian, X. Zhang, H. Wang, et al., Dielectric relaxation,
[25] R. Metz, H. Delalu, J.R. Vignalou, N. Achard, M. Elkhatib, Electrical properties of resonance and scaling behaviors in Sr3Co2Fe24O41 hexaferrite, Sci. Rep. 5 (2015)
varistors in relation to their true bismuth composition after sintering, Mater. Chem. 13645.
Phys. 63 (2000) 157–162. [50] J. Wu, C.-W. Nan, Y. Lin, Y. Deng, Giant dielectric permittivity observed in Li and
[26] M.I. Mendelson, Average grain size in polycrystalline ceramics, J. Am. Ceram. Soc. Ti doped NiO, Phys. Rev. Lett. 89 (2002) 217601.
52 (1969) 443–446. [51] A.K. Jonscher, The ‘universal’ dielectric response, Nature 267 (1977) 673–679.
[27] R.S. Roth, N.M. Hwang, C.J. Rawn, B.P. Burton, J.J. Ritter, Phase equilibria in the [52] X. Zhao, R. Liao, N. Liang, L. Yang, J. Li, J. Li, Role of defects in determining the
systems CaO–CuO and CaO-Bi2O3, J. Am. Ceram. Soc. 74 (1991) 2148–2151. electrical properties of ZnO ceramics, J. Appl. Phys. 116 (2014), 014103.
[28] E.Y. Vstavskaya, A.Y. Zuev, V.A. Cherepanov, The phase diagram of the bismuth- [53] M.A. Alim, Admittance-frequency response in zinc oxide varistor ceramics, J. Am.
calcium oxide system, Mater. Res. Bull. 29 (1994) 1233–1238. Ceram. Soc. 72 (1989) 28–32.
[29] T.R.N. Kutty, S. Ezhilvalavan, The influence of Bi2O3 non-stoichiometry on the [54] P. Cheng, S. Li, L. Zhang, J. Li, Characterization of intrinsic donor defects in ZnO
non-linear property of ZnO varistors, Mater. Chem. Phys. 38 (1994) 267–276. ceramics by dielectric spectroscopy, Appl. Phys. Lett. 93 (2008), 012902.
[30] F. Stucki, F. Greuter, Key role of oxygen at zinc oxide varistor grain boundaries,
Appl. Phys. Lett. 57 (1990) 446–448.

1237

You might also like