Next Article in Journal
Experimental Investigations on the Effect of Axial Homogenous Magnetic Fields on Propagating Vortex Flow in the Taylor–Couette System
Previous Article in Journal
Methodologies in Spectral Tuning of DSSC Chromophores through Rational Design and Chemical-Structure Engineering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Temperature Dielectric Relaxation Behaviors in Mn3O4 Polycrystals

1
School of Materials Science and Engineering, Guilin University of Electronic Technology, Guilin 541004, China
2
Guangxi Key Laboratory of Information Materials, Guilin University of Electronic Technology, Guilin 541004, China
*
Author to whom correspondence should be addressed.
Materials 2019, 12(24), 4026; https://doi.org/10.3390/ma12244026
Submission received: 19 October 2019 / Revised: 14 November 2019 / Accepted: 29 November 2019 / Published: 4 December 2019
(This article belongs to the Section Materials Physics)

Abstract

:
High temperature dielectric relaxation behaviors of single phase Mn3O4 polycrystalline ceramics prepared by spark plasma sintering technology have been studied. Two dielectric relaxations were observed in the temperature range of 200 K–330 K and in the frequency range of 20 Hz–10 MHz. The lower temperature relaxation is a type of thermally activated relaxation process, which mainly results from the hopping of oxygen vacancies based on the activation energy analysis. There is another abnormal dielectric phenomenon that is different from the conventional thermally activated behavior and is related to a positive temperature coefficient of resistance (PTCR) effect in the temperature region. In line with the impedance analyses, we distinguished the contributions of grains and grain boundaries. A comparison of the frequency-dependent spectra of the imaginary impedance with imaginary electric modulus suggests that both the long range conduction and the localized conduction are responsible for the dielectric relaxations in the Mn3O4 polycrystalline samples.

1. Introduction

The relation between physical properties and microstructure (such as grains, grain boundaries, sample-electrode interfaces, and so on) is an important aspect for ceramic materials and is helpful for better understanding their electrical properties [1,2,3,4,5]. Dielectric, modulus, and impedance measurements are the most widely used characterization methods for investigating the microstructure-property relation and relaxation mechanism. According to the temperature and frequency dependence of the dielectric peaks, the nature of the anomalies may be attributed to a thermally activated behavior, a ferroelectric phase transition, or other mechanisms. Furthermore, the contributions of grains, grain boundaries, and sample-electrode interfaces can be distinguished by impedance spectrum analysis [1,2,3,4].
The PTCR (positive temperature coefficient of resistance) effect is characterized by an increase in resistance with temperature, which is in contrast to the thermally activated behavior in which resistance decreases with temperature. The papers concerning the study of PTCR mainly focus on BaTiO3 and donor-doped BaTiO3 [6,7,8]. Goodman pointed out that the PTCR effect in BaTiO3 was related to the grain boundary [6]. Sinclair et al. suggested that the PTCR effect in BaTiO3 stems from the resistances of the grain and grain boundary [3]. The PTCR effect of the donor-doped BaTiO3 was demonstrated by Heywang-Jonker model [7,8] and was attributed mainly to the donor dopants, which resulted in the difference of the resistances between the grain and grain boundary.
The strong couplings among spin, charge, lattice, and orbital have received much attention in strongly correlated Mn3O4 systems. In addition, Mn3O4 is widely used in the electronics industry and is a raw material for the production of soft magnetic oxyferrites [9,10,11,12,13,14]. The dielectric and magneto-dielectric effects in low temperature (<43 K) for Mn3O4 have been studied in a few reports [15,16,17,18]. Previous studies indicated that the low temperature magnetoelectric coupling mechanism originated from spin-phonon coupling or the modulation of Mn3+ orbital states through the inverse process of single-ion spin anisotropy [15,16]. In this paper, we studied the microstructure-property relation of Mn3O4 polycrystalline sample in the high temperature range (200–330 K) and demonstrate the comprehensive understanding of dielectric relaxation of Mn3O4 ceramics by using dielectric and impedance spectroscopy. The results show there are two types of relaxations. The relaxation at lower temperature is a normal thermally activated relaxation process, which is associated with the hopping of oxygen vacancies. The relaxation at higher temperature is attributed to the PTCR effect, caused by the difference of the resistances between the grain and grain boundary. This work is helpful for understanding the dielectric relaxation behaviors in manganese oxides materials.

2. Experimental

Single-phase Mn3O4 polycrystalline samples were prepared using spark plasma sintering technology by adjusting the sintering temperature, applied static pressure, and holding time [19]. Chemical composition and the elemental maps were measured by a Quanta 450 FEG field emission scanning electron microscope (FESEM) and the available energy-dispersive X-ray spectroscopy (EDX) equipment (FEI, Hillsboro, OR, USA). The sintered pellet was polished and then coated with silver glue. The permittivities of samples were measured using a precise impendence analyzer (Wayne Kerr Electronics 6500B, Cavendish Square, London) with an applied voltage of 1 V in the temperature range from 200 K to 330 K and in the frequency range of 20 Hz–10 MHz. The temperature was controlled by a physical properties measurement system (Quantum Design 9T, San Diego, CA, USA).

3. Results and Discussion

The X-ray diffraction patterns of the Mn3O4 polycrystalline samples exhibit the single phase character as reported previously [18]. In order to further determine the chemical compositions and elemental maps, the EDX and energy dispersive X-ray analysis (EDXA) spectra measurements were carried out, as shown in Figure 1. The results show only Mn and O elements present in the as-prepared sample and the ratio of Mn:O = 0.73 ± 0.006, which is further evidence that the prepared Mn3O4 has a single phase. We also measured the current-density versus electric-field curve of the sample with silver electrode, as shown in Figure 1f. The nearly linear slope indicates the electrode is a good ohmic contact with ceramics. The silver glue as an electrode has some influence on the dielectric properties, which is helpful for studying these properties.
Figure 2 shows the temperature dependence of the real part (ε’) of the complex dielectric constant (å*) at various frequencies for Mn3O4 polycrystalline sample. There are two dielectric relaxation peaks. The peak at lower temperature moves slightly to the higher temperature with the frequency increasing. The other peak position at higher temperature is almost unaffected by the frequency. The electric modulus can be expressed as M* = 1/å*, which suggests that the modulus can largely reduce the background and provide information about the relaxation mechanism [20,21]. Figure 3 shows the temperature dependence of the imaginary part of the modulus (M”). The M”(T) curve shows two pronounced relaxations, from low temperature to high temperature, marked as AM1 and AM2, respectively. As the frequency increase, AM1 shifts to higher temperature, which indicates a well-known thermally activated behavior. However, the AM2 peak, which is different from the general thermal activation behavior, shifts to lower temperature with the frequency increasing. Therefore, we refer to it as an abnormal thermally activated behavior.
Generally speaking, for a thermally activated relaxation process, the variation of peak position can be described by the Arrhenius law [22]:
ƒ = ƒ0 exp (−Ea/kBTM)
where ƒ0 is pre-exponential and Ea is the activation energy. According to the Arrhenius law, it is clear that lnƒ is proportional to 1/TM. The activation energy can be obtained according to the slope. We can make a preliminary judgment on the mechanism of relaxation peaks based on the activation energy. Figure 4 shows the Arrhenius plots of M” for the two types of relaxations (AM1 and AM2). The solid line shows the fitting to the experiment data of AM1 by Equation (1). The activation energy was derived to be about 1.44 eV. Similar results were also reported in SrTiO3 ceramics [23], PbZr1−xTixO3 single crystals [24], and Mg doped PZT [25], etc. The type of dielectric relaxation is attributed to the mobility of oxygen vacancies [20,26,27]. Therefore, AM1 can be ascribed to the hopping of oxygen vacancies. For the abnormal dielectric relaxation AM2, the peak position as a function of frequency seems also to follow the Arrhenius law mathematically, but the derived value of Ea is −2.31 eV. Activation energy is the energy required to move a crystal atom away from an equilibrium position to another new equilibrium or unbalanced position. That is to say it is the energy needed to be overcome in order to start a physicochemical process. Therefore, it is difficult to understand a negative value of active energy.
Impedance spectrum analysis is a common method for analyzing the contributions of different microstructural components to the relaxation in ceramic materials [4,28]. In order to get a deep insight into the nature of the relaxation process, the impedance spectrum has been studied. Figure 5a shows the imaginary part of the impedance Z” versus the imaginary part Z’ of the impedance plots (Nyquist plots) below 260 K (for AM1). The irregular semicircular arc radius decreases with the temperature increasing, which indicates that Mn3O4 ceramics have smaller resistivity at higher temperatures between 230 K and 260 K. The irregular semicircular coil may suggest the existence of multiple relaxations in a Mn3O4 polycrystalline sample [4]. The Nyquist plots can be analyzed by using an ideal equivalent electrical circuit consisting of resistance and capacitance. This circuit can set up a connection between the microstructure and physical properties. The Nyquist plots at different temperatures have been well fitted with an equivalent circuit [29,30]. As shown in the inset of Figure 5b, the circuit consists of two sub-circuits in series. (Cgb, Cg) and (Rgb, Rg) represent the capacitances and resistances of grain boundaries and grains, respectively. CPE denotes a constant phase element with an impedance Z*CPE = A(jù)n, where A is the scale factor and n decides the departure from an ideal capacitor. Figure 5b shows a representative result at 235 K and Table 1 provides the fitted parameters. The circuit made up of two sub-circuits in series indicates that there are two relaxations [30]. The relaxation at low frequency is related to grain boundaries and the one at high frequency is duo to the grains [4,31]. The electrode has little influence on the dielectric properties, which is consistent with the above conclusion. As shown in Table 1, the resistance of the grain is smaller than that of grains boundaries, which is similar to the results of reference [1].
It is necessary to clarify the origin of the abnormal dielectric relaxation AM2 shown in Figure 3. It is well-known that the Vogel-Fulcher relation, the Arrhenius relation, or a complicated relaxation time distribution function is usually used to derive the relaxation time for the normal thermally activated phenomena [20,32]. It is difficult to understand that the relaxation behavior that the peak position shifts to low temperature with the frequency increasing for the abnormal thermally activated behavior, as shown in Figure 3. Similar phenomena were shown in BaTiO3 [3], Gd2SiO5 laser crystals [20], and BaTi0.85Zr0.15O3 ceramics [33], which are related to the PTCR effect. Therefore, the abnormal dielectric behavior in Mn3O4 might be associated with the PTCR effect in a similar way to for the above materials. The resistance R of Mn3O4 polycrystalline at different temperatures was derived according to the impedance spectrum (Z”f) studies, since the impedance peak intensity yields the value of R/2 [3]. Figure 6 displays the temperature dependence of the resistance (R). As expected, there is a critical point at 260 K. The resistance decreases with the increasing of temperature below 260 K, and increases above 260 K. The results show there is a positive temperature coefficient resistor above 260 K (the PTCR effect). The temperature region of the PTCR effect and that of the abnormal dielectric phenomenon matches perfectly. This result implies that the abnormal dielectric phenomenon stems from the PTCR effect in Mn3O4 polycrystalline. We also studied the Nyquist plots above 260 K. The semicircular arc radius of the Nyquist plot decreases as the temperature decreases (shown in Figure 7a), which shows that the resistance of the Mn3O4 ceramics increases with the temperature. The impedance data can also be fitted with the equivalent circuit, as shown in Figure 7b, and Table 1 gives the fitted parameters at 265 K. The resistance of the grain boundaries Rgb is about 9 × 1015 Ù, which is much larger than that of the grain. According to the Heywang-Jonker model, the PTCR effect can be explained by the difference of the resistances between the grain and grain boundary.
The normalized functions of M″/M″max and Z″/Z″max are shown in Figure 8 measured at 242 and 265 K. For the same temperature, the Z″/Z″max and M″/M″max peaks locate near to each other but not overlap. As reference [34] states, the overlapping of the peak position of M″/M″max and Z″/Z″max curves or not is a criterion of delocalized or long-range motions of charge carriers. Therefore, there are long-range and localized conduction below and above 260 K for the Mn3O4 polycrystalline.

4. Conclusions

In summary, the temperature and frequency dependences of dielectric constant/electric modulus/impedance spectrums have been investigated in a Mn3O4 polycrystalline sample. There are two types of dielectric relaxations. The low-temperature relaxation is due to the hopping of oxygen vacancies. The other dielectric relaxation occurs above 260 K and is different from the general thermal activation behavior, where the resistance increases with the increasing of the temperature. The temperature region of the PTCR effect and that of the abnormal dielectric behavior matches perfectly with each other. This result implies that the abnormal dielectric behavior can be ascribed to the PTCR effect in Mn3O4 polycrystalline. In line with the normalized functions of electric modulus and impedance spectrums, it can be concluded that there are long-range and localized forms of conduction below and above 260 K for Mn3O4 polycrystalline.

Author Contributions

Conceptualization, X.Z. and H.Z.; methodology, S.W.; validation, S.W., R.Y. and L.F.; formal analysis, S.W. and X.Z.; investigation, S.W.; data curation, X.Z.; writing—original draft preparation, S.W.; writing—review and editing, X.Z.

Funding

This work was supported by the National Science Foundation of China (Grant No. 51802052, 11464007), National Science Foundation of Guangxi (Grant No. 2017GXNSFAA198373), Guangxi Key Laboratory of Information Material Foundation (Grant No. 171020-Z, 171025-Z), Guilin University of Electronic Technology Talents Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bai, Y.; Wang, S.W.; Zhang, X.; Zhao, Z.K.; Shao, Y.P.; Yao, R.; Yang, M.M.; Gao, Y.B. Negative magnetization, dielectric and magnetodielectric properties of EuCrO3. Mater. Res. Express 2019, 6, 026101. [Google Scholar] [CrossRef]
  2. Alexander, B.; Alan, M.; Ralf, W.; Robert, K.; Joachim, B.; Reda, C.; Horst, H.; Oliver, C. Proton conduction in grain-boundary-free oxygen-deficient BaFeO2.5+δ thin films. Materials 2018, 11, 52. [Google Scholar]
  3. Sinclair, D.C.; West, A.R. Impedance and modulus spectroscopy of semiconducting BaTiO3 showing positive temperature-coefficient of resistance. J. Appl. Phys. 1989, 66, 3850. [Google Scholar] [CrossRef]
  4. Rehman, S.D.F.; Li, J.B.; Dou, Y.K.; Zhang, J.S.; Zhao, Y.J.; Rizwan, M.; Khalid, S.; Jin, H.B. Dielectric relaxations and electrical properties of Aurivillius Bi3.5La0.5Ti2Fe0.5Nb0.5O12 ceramics. J. Alloy. Compd. 2016, 654, 315–320. [Google Scholar] [CrossRef]
  5. Pogrebnjak, A.D.; Rogoz, V.M.; Bondar, O.V.; Erdybaeva, N.K.; Plotnikov, S.V. Structure and physicomechanical properties of NbN-based protective nanocomposite coatings: A review. Prot. Met. Phys. Chem. Surf. 2016, 52, 802–813. [Google Scholar] [CrossRef] [Green Version]
  6. Goodman, G. Electrical Conduction Anomaly in SamariumDoped Barium Titanate. J. Am. Ceram. Soc. 1963, 46, 48. [Google Scholar] [CrossRef]
  7. Wang, W.H. Resistivity anomaly in doped borium titanate. J. Am. Ceram. Soc. 1964, 47, 484. [Google Scholar]
  8. Jonker, G.H. Some aspects of semiconducting barium titanate. Solid-State Electron. 1964, 7, 895. [Google Scholar] [CrossRef]
  9. Shen, Y.F.; Zerger, R.P.; Deguzman, R.N.; Suib, S.L.; Mccurdy, L.; Potter, D.I.; Oyoung, C.L. Manganese oxide octahedral modecular-sieves-preparation, characterization, and applications. Science 1993, 260, 511. [Google Scholar] [CrossRef]
  10. Chen, B.; Rao, G.H.; Wang, S.W.; Lan, Y.A.; Pan, L.J.; Zhang, X. Facile synthesis and characterization of Mn3O4 nanoparticles by auto-combustion method. Mater. Lett. 2015, 154, 160–162. [Google Scholar] [CrossRef]
  11. Armstrong, A.R.; Bruce, P.G. Synthesis of layered LiMnO2 as an electrode for rechargeable lithium batteries. Nature 1996, 381, 499–500. [Google Scholar] [CrossRef]
  12. Wang, S.W.; Zhang, X.; Yao, R.; Rao, G.H. Size-dependent exchange bias in single phase Mn3O4 nanoparticles. Chin. Phys. B 2016, 25, 117502. [Google Scholar] [CrossRef]
  13. Kim, M.; Chen, X.M.; Wang, X.; Nelson, C.S.; Budakian, R.; Abbamonte, P.; Cooper, S.L. Pressure and field tuning the magnetostructural phases of Mn3O4: Raman scattering and X-ray diffraction studies. Phys. Rev. B 2011, 84, 174424. [Google Scholar] [CrossRef] [Green Version]
  14. Guillou, F.; Thota, S.; Prellier, W.; Kumar, J.; Hardy, V. Magnetic transitions in Mn3O4 and an anomaly at 38 K in magnetization and specific heat. Phys. Rev. B 2011, 83, 094423. [Google Scholar] [CrossRef]
  15. Tackett, R.; Lawes, G.; Melot, B.C.; Grossman, M.; Toberer, E.S.; Seshadri, R. Magnetodielectric coupling in Mn3O4. Phys. Rev. B 2007, 76, 024409. [Google Scholar] [CrossRef] [Green Version]
  16. Suzuki, T.; Katsufuji, T. Magnetodielectric properties of spin-orbital coupled system Mn3O4. Phys. Rev. B 2008, 77, 220402. [Google Scholar] [CrossRef]
  17. Dwivedi, G.D.; Kumar, A.; Yang, K.S.; Chen, B.Y.; Liu, K.W.; Chatterjee, S.; Yang, H.D.; Chou, H. Structural phase transition, Neel temperature enhancement, and persistent magneto-dielectric coupling in Cr-substituted Mn3O4. J. Appl. Phys. 2014, 116, 103906. [Google Scholar] [CrossRef]
  18. Thota, S.; Singh, K.; Nayak, S.; Simon, C.; Kumar, J.; Prellier, W. The ac-magnetic susceptibility and dielectric response of complex spin ordering processes in Mn3O4. J. Appl. Phys. 2014, 116, 103906. [Google Scholar] [CrossRef]
  19. Yao, R.; Zhang, X.; Wang, S.W.; Shao, Y.P.; Zhao, Z.K. The properties of Mn3O4 synthesized by spark plasma sintering. Powd. Meta. Tech. 2016, 34, 434–439. [Google Scholar]
  20. Liu, L.N.; Wang, C.C.; Zhang, D.M.; Zhang, Q.L.; Ning, K.J.; Wang, J.; Sun, X.H. Dielectric relaxations and phase transition in laser crystals Gd2SiO5 and Yb-doped Gd2SiO5. J. Am. Ceram. Soc. 2014, 97, 1823–1828. [Google Scholar] [CrossRef]
  21. Moynihan, C.T.; Boesch, L.P.; Laberage, N.L. Decay function for the electric field relaxation in vitreous ionic conductors. Phys. Chem. Glasses 1973, 14, 122–125. [Google Scholar]
  22. Koltunowicz, T.N.; Zukowski, P.; Czarnacka, K.; Bondariev, V.; Boiko, O.; Svito, I.A.; Fedotov, A.K. Dielectric properties of nanocomposite (Cu)x(SiO2)(100−x) produced by ion-beam sputtering. J. Alloy. Compd. 2015, 652, 444–449. [Google Scholar] [CrossRef]
  23. Wang, C.C.; Lei, C.M.; Wang, G.J.; Sun, X.H.; Li, T. Oxygen-Vacancy-Related Dielectric Relaxations in SrTiO3 at High Temperatures. J. Appl. Phys. 2013, 113, 094103. [Google Scholar] [CrossRef]
  24. Sumara, I.J.; Roleder, K.; Dec, J.; Miga, S. Ti-induced and modified dielectric relaxations in PbZrl−x, TixO3 single crystals (x ≤ 0.03) in the frequency range 10 Hz–10 MHz. J. Phys. Condens. Mat. 1995, 7, 6137–6149. [Google Scholar] [CrossRef]
  25. Guiffard, B.; Boucher, E.; Eyraud, L.; Lebrun, L.; Guyomar, D. Influence of donor co-doping by niobium or fluorine on the conductivity of Mn doped and Mg doped PZT ceramics. J. Eur. Ceram. Soc. 2005, 25, 2487–2490. [Google Scholar] [CrossRef]
  26. Ang, C.; Yu, Z.; Cross, L.E. Oxygen-vacancy-related low-frequency dielectric relaxation and electrical conduction in Bi:SrTiO3. Phys. Rev. B 2000, 62, 228–236. [Google Scholar] [CrossRef] [Green Version]
  27. Han, F.F.; Deng, J.M.; Liu, X.Q.; Yan, T.X.; Ren, S.K.; Ma, X.; Liu, S.S.; Peng, B.L.; Liu, L.J. High-temperature dielectric and relaxation behavior of Yb-doped Bi0.5Na0.5TiO3 ceramics. Cera. Inter. 2017, 43, 5564. [Google Scholar] [CrossRef]
  28. Sridarane, R.; Subramanian, S.; Janani, N.; Murugan, R. Investigation on microstructure, dielectric and impedance properties of Sr1−xBi2+(2/3)x(VxTa1−x)2O9 [x = 0, 0.1 and 0.2] ceramics. J. Alloys. Compd. 2010, 492, 642–648. [Google Scholar] [CrossRef]
  29. Raymond, O.; Font, R.; Suarez-Almodovar, N.; Portelles, J.; Siqueiros, J.M. Frequency-temperature response of ferroelectromagnetic Pb(Fe1/2Nb1/2)O3 ceramics obtained by different precursors. Part II. Impedance spectroscopy characterization. J. Appl. Phys. 2005, 97, 084108. [Google Scholar] [CrossRef]
  30. Tang, R.J.; Jiang, C.; Qian, W.; Jian, J.; Zhang, X.; Wang, H.Y.; Yang, H. Dielectric relaxation, resonance and scaling behaviors in Sr3Co2Fe24O41 hexaferrite. Sci. Rep. 2015, 5, 13645. [Google Scholar] [CrossRef]
  31. Idrees, M.; Nadeem, M.; Atif, M.; Siddique, M.; Mehmood, M.; Hassan, M.M. Origin of colossal dielectric response in LaFeO3. Acta. Mater. 2011, 59, 1338–1345. [Google Scholar] [CrossRef]
  32. Wang, C.C.; Lu, H.B.; Jin, K.J.; Yang, G.Z. Temperature-dependent dielectric strength of a Maxwell-Wagner type relaxation. Mod. Phys. Lett. B 2008, 22, 1297–1305. [Google Scholar] [CrossRef]
  33. Xu, J.; Itoh, M. Unusual dielectric relaxation in lightly doped n-type rhombohedral BaTi0.85Zr0.15O3: Ta ferroelectric ceramics. Chem. Mater. 2005, 17, 1711–1716. [Google Scholar] [CrossRef] [Green Version]
  34. Gerhardt, R. Impedance and dielectric-spectroscopy revisited distinguishing localized relaxation from long-range conductivity. J. Phys. Chem. Solids. 1994, 55, 1491–1506. [Google Scholar] [CrossRef]
Figure 1. (Color online) (a) electronic image, (bd) X-ray mapping, (e) energy-dispersive X-ray spectroscopy (EDX) spectrogram and (f) current-density versus electric-field curve at 300 K of Mn3O4 polycrystalline sample.
Figure 1. (Color online) (a) electronic image, (bd) X-ray mapping, (e) energy-dispersive X-ray spectroscopy (EDX) spectrogram and (f) current-density versus electric-field curve at 300 K of Mn3O4 polycrystalline sample.
Materials 12 04026 g001
Figure 2. (Color online) Temperature dependence of ε’ for Mn3O4 polycrystalline sample measured with various frequencies.
Figure 2. (Color online) Temperature dependence of ε’ for Mn3O4 polycrystalline sample measured with various frequencies.
Materials 12 04026 g002
Figure 3. (Color online) The electric modulus imaginary part (M”) versus the temperature plots at different frequencies.
Figure 3. (Color online) The electric modulus imaginary part (M”) versus the temperature plots at different frequencies.
Materials 12 04026 g003
Figure 4. (Color online) Arrhenius plots of M” for two types of relaxations (AM1 and AM2). Symbols are the experimental points and solid line represents the fitting.
Figure 4. (Color online) Arrhenius plots of M” for two types of relaxations (AM1 and AM2). Symbols are the experimental points and solid line represents the fitting.
Materials 12 04026 g004
Figure 5. (Color online) (a) Complex impedance below 260 K. (b) Nyquist plots at 235 K for the circuit shown.
Figure 5. (Color online) (a) Complex impedance below 260 K. (b) Nyquist plots at 235 K for the circuit shown.
Materials 12 04026 g005
Figure 6. (Color online) The resistance R versus the temperature.
Figure 6. (Color online) The resistance R versus the temperature.
Materials 12 04026 g006
Figure 7. (Color online) (a) Complex impedance above 260 K, (b) Nyquist plots for the circuit at 265 K.
Figure 7. (Color online) (a) Complex impedance above 260 K, (b) Nyquist plots for the circuit at 265 K.
Materials 12 04026 g007
Figure 8. (Color online) Normalized functions of electric modulus and impedance versus frequency at 242 and 265 K.
Figure 8. (Color online) Normalized functions of electric modulus and impedance versus frequency at 242 and 265 K.
Materials 12 04026 g008
Table 1. The fitting parameters obtained according to the experimental data by the equivalent circuit.
Table 1. The fitting parameters obtained according to the experimental data by the equivalent circuit.
Temp.(K)Rgb (MΩ)Cgb (pF)CPE (10−8 S·sn)nRg (MΩ)Cg (pF)CPE (10−8 S·sn)n
2352.628189.324.540.4730.951108.21.5010.564
2659 × 109108127380.2780.07471.953.9260.568

Share and Cite

MDPI and ACS Style

Wang, S.; Zhang, X.; Yao, R.; Fan, L.; Zhou, H. High-Temperature Dielectric Relaxation Behaviors in Mn3O4 Polycrystals. Materials 2019, 12, 4026. https://doi.org/10.3390/ma12244026

AMA Style

Wang S, Zhang X, Yao R, Fan L, Zhou H. High-Temperature Dielectric Relaxation Behaviors in Mn3O4 Polycrystals. Materials. 2019; 12(24):4026. https://doi.org/10.3390/ma12244026

Chicago/Turabian Style

Wang, Songwei, Xin Zhang, Rong Yao, Liguo Fan, and Huaiying Zhou. 2019. "High-Temperature Dielectric Relaxation Behaviors in Mn3O4 Polycrystals" Materials 12, no. 24: 4026. https://doi.org/10.3390/ma12244026

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