Next Article in Journal
Modeling, Simulation and Data Processing for Additive Manufacturing
Next Article in Special Issue
Evolution of Intrinsic and Extrinsic Electron Traps at Grain Boundary during Sintering ZnO Based Varistor Ceramics
Previous Article in Journal
Biocompatibility and Osteogenic Potential of Calcium Silicate-Based Cement Combined with Enamel Matrix Derivative: Effects on Human Bone Marrow-Derived Stem Cells
Previous Article in Special Issue
Designing the Insulation System for Motors in Electrified Aircraft: Optimization, Partial Discharge Issues and Use of Advanced Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simultaneously Enhanced Potential Gradient and Nonlinearity of ZnO Varistor Ceramics by MnO Doping with Nano-Sized ZnO Powders

1
State Key Laboratory of Electrical Insulation and Power Equipment, Xi’an Jiaotong University, Xi’an 710049, China
2
Pinggao Group Co., Ltd., Pingdingshan 467001, China
3
State Grid Zhejiang, Electric Power Co., Ltd., Hangzhou 310000, China
*
Authors to whom correspondence should be addressed.
Materials 2021, 14(24), 7748; https://doi.org/10.3390/ma14247748
Submission received: 7 November 2021 / Revised: 25 November 2021 / Accepted: 4 December 2021 / Published: 15 December 2021
(This article belongs to the Special Issue Advanced Insulating Materials and Technologies)

Abstract

:
ZnO varistor ceramics with a high potential gradient, as well as a high nonlinear coefficient, were reported and analyzed in this paper. With the use of nano-sized ZnO powders, the average grain size was reduced to about 2.6 μm, which successfully raised the potential gradient to 1172 V/mm. Moreover, the nonlinear coefficient increased to 48, and the leakage current was decreased to 8.4 μA/cm2 by doping a moderate amount of MnO (0.9 mol%). This was proven to be caused by the high Schottky barrier height formed at the grain boundary, where the Mn element segregated and, consequently, led to the increased density of interface states. Therefore, this could be considered as a potential method to simultaneously enhance the potential gradient and the nonlinear coefficient of ZnO varistor ceramics.

1. Introduction

ZnO varistor ceramics are widely employed to protect electrical power systems and electronic devices from surges and overvoltage due to their excellent nonlinear current–voltage characteristics and energy absorption capability [1,2]. High-performance ZnO varistor ceramics with a high potential gradient, high nonlinear coefficient and low leakage current are urgently required due to the increasing requirements in the fields of device miniaturization, ultra-high-voltage power systems, high-speed electrified railways, and integrated circuits, etc. [3,4,5,6,7]
The potential gradient (E1mA) of ZnO varistor ceramics is proportional to the number of grain boundaries (GB) per unit length and the breakdown voltage of a single GB, which is approximately constant at ~3 V [1]. Therefore, the commonly accepted approach to enhance the potential gradient is to reduce the grain size, which is generally achieved by optimizing doping [4,8,9,10], sintering [11,12,13] and nano-sized raw materials [5,14,15,16,17,18,19]. Among these methods, the most effective method is using nano-sized ZnO powders to prepare ZnO varistor ceramics. It was reported that the high E1mA of 800 V/mm was achieved by using nano powers with nonlinear coefficients (α) of approximately 20 [5,20]. Similarly, ZnO varistors, with a high α of 96, were prepared but followed by the high leakage current density (JL) of 50 μA/cm2 [21]. This is because a small grain size would increase the number of grain boundaries which would weaken the Schottky barrier and lower its barrier height, resulting in a low α and high JL. Therefore, it seems hard to co-achieve a high potential gradient, high nonlinearity and low leakage current for ZnO varistor ceramics [22,23,24,25]. When the high potential gradient is more than 1000 V/mm, it is difficult for the nonlinear coefficient to reach more than 40, and, at the same time, the leakage current density is seldom lower than 10 μA/cm2 [5,20,26,27,28,29]. The electrical properties of ZnO varistor ceramics are expected to largely depend on their composition. The doping of a transition metal element, e.g., Mn, is acknowledged as a feasible method to improve electrical nonlinearity and reduce the leakage current of ZnO varistor ceramics. Of course, there are many studies on the preparation of ZnO varistor ceramics by doping MnO with micron ZnO powders [20,30,31,32,33]. However, the effects of MnO doping (0.3–1.2 mol%) on the electrical properties of ZnO varistor ceramics prepared by nano sized ZnO powers are inconclusive.
In this paper, ZnO varistor ceramics with a high potential gradient of 1172 V/mm, high nonlinear coefficient of 48 and low leakage current of 8.4 μA/cm2, were obtained by doping MnO with nano-sized ZnO powers. The effects of MnO content on the microstructures, Schottky barriers and electrical properties of ZnO varistor ceramics were investigated. This could provide helpful references for preparing high-performance ZnO varistor ceramics.

2. Experiments

A series of ZnO varistor ceramic samples were prepared through the solid-state reaction method with the following raw materials: (96.2-x) mol% ZnO, 1.3 mol% Bi2O3, 1.5 mol% Sb2O3, 0.6 mol% Co3O4, 0.4 mol% Ni2O3, 0.008 mol% Al(NO3)3·9H2O and x mol% MnO (x = 0.3, 0.6, 0.9, 1.2). The chemical purity of the raw materials was 99.9%, 99.9%, 99.5%, 99.9%, 99%, 99.9% and 99.5%, respectively. Among these materials, the nano-sized ZnO powders (30 ± 10 nm) were used. The raw materials were milled in polyethylene bottles for 12 h. Then, the mixed slurry was dried at 80 °C. After dying, a large amount of 3 wt% polyvinyl alcohol (PVA) was added into the powders. Then, the slurry was mixed and granulated. The powders were pressed into pellets and pre-sintered at 600 °C. Then, the pellets were sintered at 1050 °C for 2 h with a heating rate of 3.33 °C/min and naturally cooled to room temperature in the air atmosphere. Finally, ZnO varistor ceramic samples with a diameter of about 11.5 mm and a thickness of about 2.5 mm were obtained. They were designated as M1, M2, M3, and M4, with the increasing MnO content.
The surface morphology, element analysis and crystal structure of samples were characterized by scanning electron microscope (SEM, VE-8600S, Keyence, Osaka, Japan), EDS (JSM6390A, JEOL, Tokyo, Japan) and X-ray diffraction (XRD, D8 Advance, Bruker, Berlin, Germany), respectively. The densities of the samples were measured by the Archimedes method. The nonlinear current density-electrical field (J-E) curve was measured at room temperature by using a Keithley DMM 7510 digital multi-meter and a TD2200 precision linear high-voltage DC power. The voltage dependence of barrier capacitance was measured by an impedance analyser (Novocontrol Concept 80, Frankfurt, Germany).

3. Results

Figure 1a shows the XRD patterns of the ZnO varistor ceramics. The ZnO crystal phase (JCPDS Card No. 89-0510), Bi-rich phase and spinel phase Co(Co4/3Sb2/3)O4 (JCPDS Card No. 78-0718), and Zn2.33Sb0.67O4 (JCPDS Card No. 15-0687) were identified in all of the samples. The β-Bi2O3 phase (JCPDS Card No. 27-0050) and spinel phase were the same in different samples. Those results indicated that the MnO doping content had little effect on the crystal phase of ZnO varistor ceramics, which was in accordance with other reports [20]. In order to compare the content of the intergranular phase of different samples, the Bi2O3 phase and Zn2.33Sb0.67O4 are taken as examples. The ratio of peak height of the Bi2O3 phase and Zn2.33Sb0.67O4 to that of the ZnO main crystal phase in different samples was calculated as shown in Figure 1b. With the increase in MnO content, the relative intensity of Bi2O3 and Zn2.33Sb0.67O4 were both enhanced, which indicated that MnO content influenced the content of the intergranular phase.
To further study the effects of MnO doping on the microstructure of ZnO varistor ceramics, EDS was conducted along the yellow lines in Figure 1c–f. Element distribution was measured traversing several ZnO grains, grain boundaries and intergranular phases. It can be observed that Zn was mainly distributed in the grains (shown in black in the figure), Bi existed in light-colored intergranular phases, while Sb mainly existed in the spinel phase. With the increase in MnO content, the peak of Mn was gradually close to that of Sb, which showed that most of the Mn existed in the intergranular phase and a small portion of it was dissolved in the grain.
Figure 2a–d display the SEM pictures of the samples. The grains, grain boundaries and intergranular phases were clearly observed. The grain size distribution was relatively uniform with a small amount of pores. The relative densities of the samples (ρ) are measured and listed in Table 1 and show a good densification [13]. Because of the usage of the nano-sized ZnO powders, the average grain sizes (d) of the samples were about 2.7 μm, which were much smaller than those of traditional ZnO varistor ceramics prepared from micro powders [34]. With the increase in MnO content, the intergranular phase inhibited the growth of grains, so that the average grain size shown in Figure 2e presented a decreasing trend.
In addition, the uniformity of grain size distribution (GSD) could not be ignored while reducing the average grain size. The GSDs of all the samples were statistically measured, which followed the lognormal distribution, as shown in the insets of Figure 2a–d. The normalized curves of the probability density function were compared, as shown in Figure 2f, and the coefficients of nonuniformity of GSD (ε) were calculated according to the following formula [5]:
E ( X ) = e μ + σ 2 / 2 ,
D ( X ) = e σ 2 1 ( e μ + σ 2 / 2 ) ,
ε = D ( X ) E ( X ) = e σ 2 1 ,
where μ is the logarithmic mean, σ is the logarithmic standard deviation, E(X) is the mean value of the lognormal distribution, D(X) is the standard deviation of the lognormal distribution, and ε is the coefficients of non-uniformity. The coefficients of nonuniformity are listed in Table 1. Combined with the half peak width of probability density function curves and non-uniformity coefficient, it can be seen that the GSD of sample M4 is the most non-uniform. This indicates that the excessive intergranular phase in sample M4 makes the grain non-uniform by the pinning effect.
The nonlinear J-E characteristics of the ZnO varistor ceramics are shown in Figure 3a, from which the potential gradient (E1mA), nonlinear coefficient (α), and the leakage current density (JL) were calculated [35] and shown in Table 1. It is observed that the breakdown fields of all the samples were higher than 1000 V/mm due to the rather small grain sizes of 2.7 μm. The nonlinear coefficients of samples M3 were improved to 47.9 and the leakage current decreased to 8.4 μA/cm2 by doping 0.9 mol% MnO.
Moreover, it is well known that the nonlinear coefficient is relevant to the Schottky barrier height at the grain boundary [1]. Therefore, capacitance–voltage (C-V) characteristics were used to calculate the parameters of Schottky barrier [36,37]:
( 1 C 1 2 C 0 ) 2 = 2 ( ϕ B + U g b ) e ε 0 ε r N D ,
where C is the barrier capacitance at bias voltage, C0 is the barrier capacitance with no bias voltage, ϕB is the barrier height, ND is the donor density, Ugb is DC bias at the single crystal interface, and εr and ε0 are the permittivity of ZnO and vacuum, respectively. Consequently, (1/C − 1/(2C0))2~U curves are plotted in Figure 3b, which shows a good linearity. The barrier height(ϕB), density of the donor (ND), density of the interface state (Ns) and width of the depletion layer(t) were calculated by (1/C−1/(2C0))2~U curve, as shown in Table 1.
It can be observed that the high barrier height of sample M3 was the main reason for its great nonlinearity. Furthermore, the density of the donor and interface state of sample M3 were both larger than those of other samples. It is well known that the Mn2+ dissolved in grains mainly replaced the lattice position of Zn2+ because of the similar radius. Therefore, it had little influence on the electron carrier concentration [31]. However, most of the Mn element was found to segregate at the grain boundary, which was reported to increase the concentration of V″Zn and V′Zn [38]. The depletion layer should satisfy the following the principle of electric neutrality [38]:
n + [ V Z n ] + 2 [ V Z n ] = p + [ Z n i · ] + 2 [ Z n i · · ] + [ V o · ] + 2 [ V o · · ] ,
where n is the electron carrier concentration and p is the hole carrier concentration. In order to satisfy the principle of electric neutrality, ND increased. ND and Ns satisfy the following relationship with ϕB [1]:
ϕ B = e 2 N S 2 2 ε 0 ε r N D .
The interface state density of sample M3 increased, leading to the growth of the barrier height and nonlinear coefficient [39]. However, when MnO was over doped in sample M4, the electrical properties became worse due to the excessive intergranular phase. Therefore, combined with the microstructure, electrical parameters and barrier height, it can be concluded that sample M3 has the best performance when doping 0.9 mol% MnO.

4. Conclusions

In this study, ZnO varistor ceramics with a high performance were prepared. The increased potential gradient of 1172 V/mm was mainly attributed to the usage of nano powders, with the average grain size reduced to only 2.6 μm. The addition of MnO helped to increase the nonlinear coefficient to 47.96 and lower the leakage current to 8.4 μA/cm2. When doping a moderate amount of MnO, most of the Mn element, which segregated at grain boundary, enhanced the density of interface states, resulting in a high Schottky barrier height.

Author Contributions

Conceptualization, Y.W. and K.W.; methodology, Y.W.; software, Y.W. and Z.H.; validation, Y.W. and K.W.; formal analysis, Y.W., K.W., J.S. and K.L.; investigation, Y.W.; resources, J.L. and C.X.; data curation, Y.W. and Z.H.; writing—original draft preparation, Y.W.; writing—review and editing, Y.W., K.W., J.S., R.C., L.H. and J.L.; supervision, K.W. and J.L.; project administration, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Project of State Grid Corporation of China (5500-201947325A-0-0-00).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gupta, T.K. Application of Zinc Oxide varistors. J. Am. Ceram. Soc. 1990, 73, 1817–1840. [Google Scholar] [CrossRef]
  2. Shaifudin, M.S.; Ghazali, M.S.M. Synergistic Effects of Pr6O11 and Co3O4 on Electrical and Microstructure Features of ZnO-BaTiO3 Varistor Ceramics Muhamad. Materials 2021, 14, 702. [Google Scholar] [CrossRef]
  3. Meng, P.; Zhao, X.; Fu, Z.; Wu, J.; Hu, J.; He, J. Novel zinc-oxide varistor with superior performance in voltage gradient and aging stability for surge arrester. J. Alloys Compd. 2019, 789, 948–952. [Google Scholar] [CrossRef]
  4. Meng, P.; Wu, J.; Yang, X.; Hu, J.; He, J. Electrical properties of ZnO varistor ceramics modified by rare earth-yttrium and gallium dopants. Mater. Lett. 2018, 233, 20–23. [Google Scholar] [CrossRef]
  5. Liu, W.; Zhang, L.; Kong, F.; Wu, K.; Li, S.; Li, J. Enhanced voltage gradient and energy absorption capability in ZnO varistor ceramics by using nano-sized ZnO powders. J. Alloys Compd. 2020, 828, 154252. [Google Scholar] [CrossRef]
  6. Xu, D.; Wu, J.; Jiao, L.; Xu, H.; Zhang, P.; Yu, R.; Cheng, X. Highly nonlinear property and threshold voltage of Sc2O3 doped ZnO-Bi2O3-based varistor ceramics. J. Rare Earths 2013, 31, 158–163. [Google Scholar] [CrossRef]
  7. Guo, M.; Zhao, X.; Shi, W.; Zhang, B.; Wu, K.; Li, J. Simultaneously improving the electrical properties and long-term stability of ZnO varistor ceramics by reversely manipulating intrinsic point defects. J. Eur. Ceram. Soc. 2021, 42, 162–168. [Google Scholar] [CrossRef]
  8. Nahm, C.W. Er2O3 doping effect on electrical properties of ZnO-V2O5-MnO2-Nb2O5 varistor ceramics. J. Am. Ceram. Soc. 2011, 94, 3227–3229. [Google Scholar] [CrossRef]
  9. Nahm, C.W. Effect of Yb2O3 addition on varistor properties and aging characteristics of ZnO-V2O5-Mn3O4 system. J. Mater. Sci. Mater. Electron. 2018, 29, 2958–2965. [Google Scholar] [CrossRef]
  10. Zhang, L.; Liu, W.; Gao, J.; Kong, F.; Li, Y.; Li, S. Effects of the Er2O3 doping on the microstructure and electrical properties of ZnO–Bi2O3 based varistor ceramics. Ceram. Int. 2021, 47, 32349–32356. [Google Scholar] [CrossRef]
  11. Xu, D.; Shi, L.; Wu, Z.; Zhong, Q.; Wu, X. Microstructure and electrical properties of ZnO-Bi2O3-based varistor ceramics by different sintering processes. J. Eur. Ceram. Soc. 2009, 29, 1789–1794. [Google Scholar] [CrossRef]
  12. Mazaheri, M.; Hassanzadeh-Tabrizi, S.A.; Sadrnezhaad, S.K. Hot pressing of nanocrystalline zinc oxide compacts: Densification and grain growth during sintering. Ceram. Int. 2009, 35, 991–995. [Google Scholar] [CrossRef]
  13. Savary, E.; Marinel, S.; Gascoin, F.; Kinemuchi, Y.; Pansiot, J.; Retoux, R. Peculiar effects of microwave sintering on ZnO based varistors properties. J. Alloys Compd. 2011, 509, 6163–6169. [Google Scholar] [CrossRef]
  14. Hembram, K.; Rao, T.N.; Srinivasa, R.S.; Kulkarni, A.R. High performance varistors prepared from doped ZnO nanopowders made by pilot-scale flame spray pyrolyzer: Sintering, microstructure and properties. J. Eur. Ceram. Soc. 2015, 35, 3535–3544. [Google Scholar] [CrossRef]
  15. Meng, L.; Zheng, L.; Cheng, L.; Li, G.; Huang, L.; Zhang, F. Synthesis of novel core-shell nanocomposites for fabricating high breakdown voltage ZnO varistors. J. Mater. Chem. 2011, 21, 11418–11423. [Google Scholar] [CrossRef]
  16. Boumezoued, A.; Guergouri, K. Synthesis and characterization of ZnO-based nano-powders: Study of the effect of sintering temperature on the performance of ZnO-Bi2O3 varistors. J. Mater. Sci. Mater. Electron. 2021, 32, 3125–3139. [Google Scholar] [CrossRef]
  17. Li, Y.; Li, G.; Yin, Q. Preparation of ZnO varistors by solution nano-coating technique. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2006, 130, 264–268. [Google Scholar] [CrossRef]
  18. Cheng, L.H.; Zheng, L.Y.; Meng, L.; Li, G.R.; Gu, Y.; Zhang, F.P.; Chu, R.Q.; Xu, Z.J. Electrical properties of Al2O3-doped ZnO varistors prepared by sol-gel process for device miniaturization. Ceram. Int. 2012, 38, S457–S461. [Google Scholar] [CrossRef]
  19. Gao, L.; Li, Q.; Luan, W. Preparation and Electric Properties of Dense Nanocrystalline Zinc Oxide Ceramics. J. Am. Ceram. Soc. 2002, 18, 2001–2003. [Google Scholar] [CrossRef]
  20. Abrishami, M.E.; Kompany, A. Varistor behavior of Mn doped ZnO ceramics prepared from nanosized precursors. J. Electroceramics 2012, 29, 125–132. [Google Scholar] [CrossRef]
  21. Maleki, M.; Ali, S.; Ali, M.; Sani, F.; Nemati, A.; Safaee, I. Two-step sintering of ZnO varistors. Solid State Ionics 2011, 190, 99–105. [Google Scholar] [CrossRef]
  22. Durán, P.; Tartaj, J.; Moure, C. Fully dense, fine-grained, doped zinc oxide varistors with improved nonlinear properties by thermal processing optimization. J. Am. Ceram. Soc. 2003, 86, 1326–1329. [Google Scholar] [CrossRef]
  23. Viswanath, R.N.; Ramasamy, S.; Ramamoorthy, R.; Jayavel, P.; Nagarajan, T. Preparation and characterization of nanocrystalline ZnO based materials for varistor applications. Nanostructured Mater. 1995, 6, 993–996. [Google Scholar] [CrossRef]
  24. Anas, S.; Metz, R.; Sanoj, M.A.; Mangalaraja, R.V.; Ananthakumar, S. Sintering of surfactant modified ZnO-Bi2O3 based varistor nanopowders. Ceram. Int. 2010, 36, 2351–2358. [Google Scholar] [CrossRef]
  25. Boumezoued, A.; Guergouri, K.; Barille, R.; Rechem, D.; Zaabat, M.; Rasheed, M. ZnO nanopowders doped with bismuth oxide, from synthesis to electrical application. J. Alloys Compd. 2019, 791, 550–558. [Google Scholar] [CrossRef]
  26. Mirzayi, M. The effect of TiO2 concentration on the electrical and microstructural properties of ZnO-base varistor ceramic prepared from nanosize ZnO particles. Adv. Appl. Ceram. 2020, 119, 373–379. [Google Scholar] [CrossRef]
  27. Macary, L.S.; Kahn, M.L.; Estournès, C.; Fau, P.; Flash, D.F.; Trémouilles, D.; Bafleur, M.; Renaud, P.; Chaudret, B. Size effect on properties of varistors made from zinc oxide nanoparticles through low temperature spark plasma sintering. Adv. Funct. Mater. 2009, 19, 1775–1783. [Google Scholar] [CrossRef] [Green Version]
  28. Cheng, L.; Li, G.; Yuan, K.; Meng, L.; Zheng, L. Improvement in nonlinear properties and electrical stability of ZnO varistors with B2O3 additives by nano-coating method. J. Am. Ceram. Soc. 2012, 95, 1004–1010. [Google Scholar] [CrossRef]
  29. Rohini, R.; Pugazhendhi Sugumaran, C. Enhancement of Electro-Thermal Characteristics of Micro/Nano ZnO Based Surge Arrester. J. Electr. Eng. Technol. 2021, 16, 469–481. [Google Scholar] [CrossRef]
  30. Hong, Y.W.; Kim, J.H. The electrical properties of Mn3O4-doped ZnO. Ceram. Int. 2004, 30, 1301–1306. [Google Scholar] [CrossRef]
  31. Han, J.; Senos, A.M.R.; Mantas, P.Q. Varistor behaviour of Mn-doped ZnO ceramics. J. Eur. Ceram. Soc. 2002, 22, 1653–1660. [Google Scholar] [CrossRef]
  32. Toplan, Ö.; Günay, V.; Özkan, O.T. Grain growth in the MnO added ZnO-6 wt% Sb2O3 ceramic system. Ceram. Int. 1997, 23, 251–255. [Google Scholar] [CrossRef]
  33. Ohashi, N.; Terada, Y.; Ohgaki, T.; Tanaka, S.; Tsurumi, T.; Fukunaga, O.; Haneda, H.; Tanaka, J. Synthesis of ZnO bicrystals doped with Co or Mn and their electrical properties. Jpn. J. Appl. Phys. 1999, 38, 5028–5032. [Google Scholar] [CrossRef]
  34. Huang, Y.; Wu, K.; Tang, Z.; Xin, L.; Zhang, L.; Li, J. Investigation of electrical inhomogeneity in ZnO varistor ceramics based on electronic relaxations. Ceram. Int. 2019, 45, 1110–1114. [Google Scholar] [CrossRef]
  35. Guo, M.; Wu, K.; Zhang, L.; Ying, L. Revisiting the effects of Co2O3 on multiscale defect structures and relevant electrical properties in ZnO varistors. High Volt. 2020, 5, 241–248. [Google Scholar] [CrossRef]
  36. Wu, K.; Huang, Y.; Hou, L.; Tang, Z.; Li, J.; Li, S. Effects of dc bias on dielectric relaxations in CaCu3Ti4O12 ceramics. J. Mater. Sci. Mater. Electron. 2018, 29, 4488–4494. [Google Scholar] [CrossRef]
  37. Wang, X.; Ren, X.; Li, Z.; You, W.; Jiang, H.; Yu, W.; Jin, L.; Yao, Z.; Shi, L. A unique tuning effect of Mg on grain boundaries and grains of ZnO varistor ceramics. J. Eur. Ceram. Soc. 2021, 41, 2633–2640. [Google Scholar] [CrossRef]
  38. Han, J.; Mantas, P.Q.; Senos, A.M.R. Defect chemistry and electrical characteristics of undoped and Mn-doped ZnO. J. Eur. Ceram. Soc. 2002, 22, 49–59. [Google Scholar] [CrossRef]
  39. Wu, K.; Wang, Y.; Hou, Z.; Li, S.; Li, J.; Tang, Z.; Lin, Y. Colossal permittivity due to electron trapping behaviors at the edge of double Schottky barrier. J. Phys. D Appl. Phys. 2021, 54, 045301. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns, (b) relative intensity of Bi2O3 and (cf) Line-scanning EDS results covering several grains and grain boundaries of ZnO varistor samples.
Figure 1. (a) XRD patterns, (b) relative intensity of Bi2O3 and (cf) Line-scanning EDS results covering several grains and grain boundaries of ZnO varistor samples.
Materials 14 07748 g001
Figure 2. (ad) SEM pictures, (e) average grain size and (f) normalized probability plots of the ZnO varistor ceramics.
Figure 2. (ad) SEM pictures, (e) average grain size and (f) normalized probability plots of the ZnO varistor ceramics.
Materials 14 07748 g002aMaterials 14 07748 g002b
Figure 3. (a) Nonlinear J-E curves and (b) (1/C − 1/(2C0))2~U curves of ZnO varistor ceramics.
Figure 3. (a) Nonlinear J-E curves and (b) (1/C − 1/(2C0))2~U curves of ZnO varistor ceramics.
Materials 14 07748 g003
Table 1. Parameters of ZnO varistor ceramics.
Table 1. Parameters of ZnO varistor ceramics.
Samplesd
(µm)
ερE1mA
(V/mm)
αJL
(μA/cm2)
ϕB
(eV)
ND
(1022 m−3)
Ns
(1016 cm−2)
t
(nm)
M12.800.31497.1%116315.2540.11.427.876.28399
M22.770.30897.3%115115.2830.71.438.416.54388
M32.660.31697.6%117247.868.41.6912.598.68345
M42.690.35097.4%117935.1422.41.536.726.03449
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Y.; Hou, Z.; Li, J.; Wu, K.; Song, J.; Chen, R.; Li, K.; Hao, L.; Xu, C. Simultaneously Enhanced Potential Gradient and Nonlinearity of ZnO Varistor Ceramics by MnO Doping with Nano-Sized ZnO Powders. Materials 2021, 14, 7748. https://doi.org/10.3390/ma14247748

AMA Style

Wang Y, Hou Z, Li J, Wu K, Song J, Chen R, Li K, Hao L, Xu C. Simultaneously Enhanced Potential Gradient and Nonlinearity of ZnO Varistor Ceramics by MnO Doping with Nano-Sized ZnO Powders. Materials. 2021; 14(24):7748. https://doi.org/10.3390/ma14247748

Chicago/Turabian Style

Wang, Yao, Zongke Hou, Jianying Li, Kangning Wu, Jiguang Song, Rui Chen, Kai Li, Liucheng Hao, and Chenbo Xu. 2021. "Simultaneously Enhanced Potential Gradient and Nonlinearity of ZnO Varistor Ceramics by MnO Doping with Nano-Sized ZnO Powders" Materials 14, no. 24: 7748. https://doi.org/10.3390/ma14247748

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop