Next Article in Journal
In Vivo Detection of Tetrodotoxin in Takifugu obscurus Based on Solid-Phase Microextraction Coupled with Ultrahigh-Performance Liquid Chromatography–Tandem Mass Spectrometry
Previous Article in Journal
Propolis: Its Role and Efficacy in Human Health and Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dielectric Responses of (Zn0.33Nb0.67)xTi1−xO2 Ceramics Prepared by Chemical Combustion Process: DFT and Experimental Approaches

by
Theeranuch Nachaithong
1,
Pairot Moontragoon
2,3,4,* and
Prasit Thongbai
4
1
Materials Science and Nanotechnology Program, Faculty of Science, Khon Kaen University, Khon Kaen 40002, Thailand
2
Institute of Nanomaterials Research and Innovation for Energy (IN-RIE), Khon Kaen University, Khon Kaen 40002, Thailand
3
Thailand Center of Excellence in Physics, Commission on Higher Education, Bangkok 10400, Thailand
4
Department of Physics, Faculty of Science, Khon Kaen University, Khon Kaen 40002, Thailand
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(18), 6121; https://doi.org/10.3390/molecules27186121
Submission received: 4 August 2022 / Revised: 7 September 2022 / Accepted: 14 September 2022 / Published: 19 September 2022
(This article belongs to the Section Physical Chemistry)

Abstract

:
The (Zn, Nb)-codoped TiO2 (called ZNTO) nanopowder was successfully synthesized by a simple combustion process and then the ceramic from it was sintered with a highly dense microstructure. The doped atoms were consistently distributed, and the existence of oxygen vacancies was verified by a Raman spectrum. It was found that the ZNTO ceramic was a result of thermally activated giant dielectric relaxation, and the outer surface layer had a slight effect on the dielectric properties. The theoretical calculation by using the density functional theory (DFT) revealed that the Zn atoms are energy preferable to place close to the oxygen vacancy (Vo) position to create a triangle shape (called the ZnVoTi defect). This defect cluster was also opposite to the diamond shape (called the 2Nb2Ti defect). However, these two types of defects were not correlated together. Therefore, it theoretically confirms that the electron-pinned defect-dipoles (EPDD) cannot be created in the ZNTO structure. Instead, the giant dielectric property of the (Zn0.33Nb0.67)xTi1xO2 ceramics could be caused by the interfacial polarization combined with electron hopping between the Zn2+/Zn3+ and Ti3+/Ti4+ ions, rather than due to the EPDD effect. Additionally, it was also proved that the surface barrier-layer capacitor (SBLC) had a slight influence on the giant dielectric properties of the ZNTO ceramics. The annealing process can cause improved dielectric properties, which are properties with a huge advantage to practical applications and devices.

1. Introduction

Recently, giant or colossal dielectric permittivity oxide materials, which are the metal oxides with a high dielectric permittivity (ε′) (more than 103), e.g., BaTiO3, CCTO, LaFeO3 and codoped NiO-based oxides [1,2,3,4,5,6,7], have been extensively researched [8,9,10,11,12,13,14,15]. They have a potential to apply in many devices, i.e., multilayer ceramic capacitors and high-energy-dense storage devices. However, there are still some practical problems about their high dielectric loss tangent (tanδ) values, particularly in a range of low frequency.
In recent years, the giant dielectric properties of (In, Nb)-codoped TiO2 ceramics (called INTO) [16] have been studied. It has been reported that codoped TiO2 ceramics show a very high ε′ (~104), low tanδ (~0.02) and high temperature stability from 80 to 450 K. However, there is still controversy in the origination of this colossal dielectric behavior. In some reports, the electron-pinned defect-dipole (EPDD) was caused by the colossal permittivity of this metal oxide. On the other hand, it was claimed that the colossal permittivity was contributed from the other effects, such as the internal barrier-layer capacitor (IBLC) effect, the surface barrier-layer capacitor (SBLC), the sample–electrode contact effect and polaronic hopping models [16,17,18,19,20]. Therefore, in this work, the mechanics inside of the giant dielectric permittivity of TiO2-based ceramics will be reported and clarified. Although the colossal dielectric response of (Zn, Nb)-codoped TiO2 was widely investigated [21,22], the colossal permittivity properties of these system ceramics synthesized by wet chemical routes has not been widely reported. Thus, in this study, both experimental and theoretical studies were simultaneously performed on TiO2 ceramics codoped with Zn2+/Nb5+ ions. In the experimental part, both the (Zn0.33Nb0.67)xTi1−xO2 powders and sintered ceramics were thoroughly fabricated and characterized. Their colossal permittivity properties at different frequencies and temperatures were investigated to expose the mechanics and behavior inside the microstructures. In the theoretical aspect, the density functional theory (DFT) was employed to evaluate the ground-state properties of the (Zn0.33Nb0.67)xTi1−xO2 to understand the cause of the colossal permittivity in the (Zn, Nb)-codoped TiO2 ceramics. The feasible explanation for the noticed giant permittivity behavior will be reported.

2. Experimental Details

(Zn0.33Nb0.67)xTi1−xO2 (ZNTO) nanoparticles (x = 0.5, 1, 2.5, 5.0 and 10%) were synthesized by a simple combustion method. C16H28O6Ti (Sigma-Aldrich, Bangkok, Thailand), NbCl5 (Sigma-Aldrich, >99.9%), N2O6Zn·xH2O (Sigma-Aldrich, 99.999%) were weighed corresponding to each doped content. Firstly, NbCl5 and N2O6Zn·xH2O were mixed in citric acid and then a C16H28O6Ti solution was mixed at 130 °C and stirred until a gel was obtained. Secondly, all doping concentrations gels were calcined, then they were pressed into pellets with diameter of 9.5 mm and thickness of ~1.3 mm. Finally, they were sintered at 1400 °C for 5 h. (Zn0.33Nb0.67)xTi1−xO2 ceramics with x = 0.5 and 0.25% are referred to as the 0.5% ZNTO and 2.5% ZNTO ceramics, respectively.
The phase structure of all samples was characterized by X-ray diffraction technique (XRD, PANalytical, EMPYREAN). The field emission scanning electron microscopy technique (FE-SEM, FEI, Hileos Nanolab G3CX) was performed to reveal the surface morphologies of the homogeneity of all elements in the sintered samples. To confirm existing oxygen vacancies in the microstructures, the Raman spectra of sintered ceramics were measured with a Raman System (NT-MDT Ntegra Spectra), using laser wavelength of 532 nm. Finally, the dielectric properties or permittivity responses were measured by an impedance analyzer technique (KEYSIGHT E4990A). In this study, the temperature dependence of the dielectric constant (ε′) and loss tangent (tanδ) was measured in the frequency and temperature ranges of 102–106 Hz and −60 to 200 °C. In order to thoroughly understand inside the nanoscale level, the stable configuration of the periodic boundary conditions of 2 × 2 × 6 super-cell Zn and Nb-codoped rutile–TiO2 structure was studied using the DFT calculation, performed under Vienna Ab initio Simulation Package (VASP) with Projector-augmented plane-wave pseudopotential method (PAW) and the Perdew–Burke–Ernzerhof (PBE) form of exchange–correlation functional. In this model, firstly, one oxygen vacancy was created and then replaced two Ti atoms with Zn atoms to form 2ZnVo triangular defect and another two Ti atoms with Nb atoms to form 2NbTi diamond defect. Additionally, cutoff energy of 600 eV and 3 × 3 × 3 k-point meshes in Monkhorst–Pack k-point are also employed to optimize the structures to obtain the lowest energy-preferable configuration, the conjugate-gradient algorithm and Hellmann–Feynman theorem were carried out to calculate the force acting on each ion and 5 × 5 × 5 k-point meshes in Monkhorst–Pack k-point are used to calculate the electronic structures. The orbitals of Ti(3p6 4s2 3d2), Zn(3p6 4s2 3d10), O(2s2 2p4) and Nb(4p6 5s1 4d4) were treated as valence electrons.

3. Results and Discussion

According to Figure 1, the XRD patterns of the ZNTO powders prepared by a chemical combustion method confirmed the rutile-TiO2 (JCPDS 21-1276) [23,24] phase and no impurity phase, which is in good agreement with other works [23,24,25]. The a and c values, extracted from the XRD patterns, are shown in Table 1, and they increased with the rising doping content. These results confirmed that the Zn2+ and Nb5+ can substitute in the Ti sites of the rutile structure of TiO2. The increase in the lattice parameters was due to the larger ionic radii of the dopants compared to that of the Ti4+ host ion.
According to Figure 2, the surface morphologies of the ZNTO ceramic sintered at 1400 °C for 5 h with 0.5 and 2.5% were revealed. It shows that the grains with grain sizes of 5.7 ± 2.4 and 3.7 ± 1.3 µm, respectively, and the grain boundaries were clearly observed. The microstructure of the sintered ceramics was highly dense without pores in the 0.5% ZNTO and 2.5% ZNTO ceramics. The mean grain size of the ZNTO ceramics decreased with an increasing codoping concentration. This result is similar to those reported in the literature by Nachaithong et al. [26] and Yang et al. [27]. It was explained that the decreased mean grain size of the codoped TiO2 ceramics was caused by the solute drag mechanism.
As shown in Figure 3, the Raman spectra of the pure TiO2 and ZNTO ceramics with the Eg and A1g modes were presented. The Eg peaks of the TiO2, 0.5% Nb-doped TiO2, 0.5% Zn-doped TiO2, 0.5% ZNTO and 2.5% ZNTO samples appeared at 446.5, 446.5, 446.5, 444.5 and 444.5 cm−1, respectively, whereas the A1g modes appeared at 610.5, 610.0, 611.0, 611.5 and 610.5 cm−1, respectively. It was shown that the oxygen vacancies and O–Ti–O bonds were associated with the Eg and A1g modes [28].
According to Figure 4, the SEM-mapping method was used to reveal a distribution of the elements in the ZNTO ceramics (such as Zn, Nb, Ti and O), and it was found that there is a homogeneous dispersion of the doped elements in the microstructure.
Figure 5 shows the ε′ permittivity at room temperature of the Pure-TiO2, 0.5% ZNTO and 2.5% ZNTO ceramics. All the samples exhibited a giant dielectric permittivity in the frequency range of 40–106 Hz. The ε′ value of a pure-TiO2 was the lowest. Although a rutile-TiO2 had the largest ε′ value among the simple oxides, its ε′ value is very low compared to those of many complex oxides, such as BaTiO3 and CaCu3Ti4O12-based materials. Nevertheless, the ε′ value of TiO2 can be significantly enhanced by codoping metal ions. In the tanδ spectrum, the samples showed that the tanδ peaks in the frequency range of 103–104 and ε′ slightly changed. This dielectric relaxation behavior could imply that the Nb5+ and Zn2+ codoping ions have slightly affected the ionic polarization. However, when it was considered in a low-frequency range, the tanδ increased with an increasing codopant concentration. It shows that the codopants have an effect on the interfacial polarization which is usually induced at the internal insulating layer, i.e., a semiconducting region or surface of the sample–electrode layer of ceramics. The ε′ values of the 0.5% ZNTO and 2.5% ZNTO ceramics were about ≈ 9 × 104 and 3 × 104 with tanδ ≈ 0.26 and 1.25 at 1 kHz and RT, respectively.
To study the effects of the surface on the electrical properties of ZNTO ceramics in Figure 6, after the dielectric properties of the as-fired sample were measured, both sides of the electrodes and the outer surface layers of the pellet were removed by polishing them with SiC paper (referred to as the polished-sample), and after that, the polished-sample was measured, the electrodes were removed by the SiC paper and annealed at 1200 °C in air for 30 min (referred to as the annealed-sample). In comparison with the as-sample, polished-sample and annealed-sample, at room temperature, all the samples exhibited very high ε′ values of ≈104–105. The change in the ε′ and tanδ was not considerable compared with that of the polished-sample, but the dielectric permittivity slightly increased in frequency, more than 103 Hz. The tanδ decreased significantly when compared with the as-sample. After the annealing process, the greatly reduced tanδ value and greatly increased ε′ value in the anneal-sample were primarily due to the filling of the oxygen vacancies on the surface. Therefore, the SBLC effect was a key factor for the anneal-sample. In this experiment, it was clearly shown that the outer surface layer influences the dielectric responses of the ZNTO ceramics. The annealing method was suggested to be one of the most important for the improvement of the dielectric properties of ZNTO ceramics by creating a resistive outer surface layer. Furthermore, this method could be a new outer surface design approach for other codoped TiO2 ceramic systems.
To explain the behavior of the colossal dielectric properties in ZNTO ceramics, the effect of temperature on the dielectric behavior at different frequencies and temperatures was investigated, as shown in Figure 7. It is seen that the ε′ step likely decreases with an increasing frequency, but increases with an increasing temperature, corresponding to a decrease in the resistance (R). Moreover, the corresponding tanδ peak also increases. This confirmed the thermally activated giant dielectric relaxation behavior. According to Figure 8, the Arrhenius plots of the 0.5% ZNTO and 2.5% ZNTO ceramics show the activation energy (Ea) which is the temperature dependence of the relaxation peak of the dielectric loss (fmax), following the Arrhenius law:
f max = f 0 exp ( E a K B T )
It is required energy for the dielectric relaxation in the ceramics and calculated from the slope of the ln fmax vs. 1000/T plots. The Ea values of the 0.5% ZNTO and 2.5% ZNTO ceramics were 0.213 and 0.175 eV., respectively. It is implied that the giant dielectric response could be related to the Zn2+/Zn3+ or Ti3+/Ti4+ electron hopping. Increasing the Nb5+ doping concentration could result in a Ti3+/Ti4+ ions ratio increase. The concentration of the free electrons in the Nb5+-doped TiO2 is generally proportional to the Nb5+ dopant concentration, following the equations [16]:
2 TiO 2 + Nb 2 O 5 4 TiO 2 2 Ti Ti + 2 Nb Ti + 8 O O + 1 2 O 2 ,
Ti 4 + + e Ti 3 +
Therefore, the electron hopping can be readily stimulated when the Ti3+ or Zn3+ content rises, giving rise to a decrease in the Ea values for the ZNTO ceramics with x = 0.5 and 2.5%.
According to the theoretical investigations, we placed 2Nb atoms preferentially forming a diamond-shaped structure (called 2Nb2Ti defect) and the 2ZnVo triangular defect into the TiO2 structure simultaneously in three configurations: near, opposite and far. The lowest total energy can be obtained when the 2Nb diamond defects and 2ZnVo triangular defects are opposite (as shown in Figure 9), corresponding to the EPDD model and not in the ZNTO structure. It can be clearly suggested that the giant dielectric relaxation behavior of the ZNTO ceramics did not originate from the EPDD effect. Moreover, the electron hopping mechanism between the Ti3+/Ti4+ ions and Zn2+/Zn3+ is also the most likely mechanism to be related to the giant dielectric relaxation in the ZNTO ceramics.

4. Conclusions

In this work, the (Zn + Nb)-codoped TiO2 ceramics prepared by a combustion process show a colossal dielectric permittivity and a low loss behavior. Moreover, their dielectric characteristics show a good frequency stability. The annealing process can cause improved dielectric properties, which are properties with a huge advantage in practical applications and devices. In addition, the giant dielectric relaxation behavior of the ZNTO originated from both the IBLC and the Ti3+/Ti4+ ions and Zn2+/Zn3+ electron hopping mechanism. It indicates that there is giant dielectric relaxation behavior in the (Zn, Nb)-codoped TiO2 systems.

Author Contributions

Conceptualization, P.T. and P.M.; methodology, T.N.; software, T.N.; validation, T.N., P.M. and P.T.; formal analysis, T.N.; investigation, T.N.; resources, T.N.; data curation, T.N.; writing—original draft preparation, T.N.; writing—review and editing, P.M.; visualization, P.M.; supervision, P.M.; project administration, P.M.; funding acquisition, P.T. All authors have read and agreed to the published version of the manuscript.

Funding

The Fundamental Fund Khon Kaen University, the Basic Research Fund of Khon Kaen University, the National Research Council of Thailand (NRCT) and Khon Kaen University: N42A650294. Thailand Research Fund under The Royal Golden Jubilee Ph.D. Program [Grant Number PHD/0114/2559].

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

The study did not report any data.

Acknowledgments

T. Nachaithong would like to thank the Thailand Research Fund under The Royal Golden Jubilee Ph.D. Program [Grant Number PHD/0114/2559] for his Ph.D. scholarship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, J.; Nan, C.W.; Lin, Y.; Deng, Y. Giant dielectric permittivity observed in Li and Ti doped NiO. Phys. Rev. Lett. 2002, 89, 217601. [Google Scholar] [CrossRef] [PubMed]
  2. Schmidt, R.; Stennett, M.C.; Hyatt, N.C.; Pokorny, J.; Prado-Gonjal, J.; Li, M.; Sinclair, D.C. Effects of sintering temperature on the internal barrier layer capacitor (IBLC) structure in CaCu3Ti4O12 (CCTO) ceramics. J. Eur. Ceram. Soc. 2012, 32, 3313–3323. [Google Scholar] [CrossRef]
  3. Liu, L.; Fan, H.; Fang, P.; Chen, X. Sol–gel derived CaCu3Ti4O12 ceramics: Synthesis, characterization and electrical properties. Mater. Res. Bull. 2008, 43, 1800–1807. [Google Scholar] [CrossRef]
  4. Li, J.; Wu, K.; Jia, R.; Hou, L.; Gao, L.; Li, S. Towards enhanced varistor property and lower dielectric loss of CaCu3Ti4O12 based ceramics. Mater. Des. 2016, 92, 546–551. [Google Scholar] [CrossRef]
  5. Xu, T.; Wang, C.-A. Effect of two-step sintering on micro-honeycomb BaTiO3 ceramics prepared by freeze-casting process. J. Eur. Ceram. Soc. 2016, 36, 2647–2652. [Google Scholar] [CrossRef]
  6. Liu, Y.; Wang, W.; Huang, J.; Zhu, C.; Wang, C.; Cao, Y. High dielectric permittivity of Li and Sc co-doped NiO ceramics. J. Mater. Sci. Mater. Electron. 2014, 25, 1298–1302. [Google Scholar] [CrossRef]
  7. Ni, W.; Ye, J.; Guo, Y.; Cheng, C.; Lin, Z.; Li, Y.; Wang, H.; Yu, Y.; Li, Q.; Huang, S.; et al. Decisive role of mixed-valence structure in colossal dielectric constant of LaFeO3. J. Am. Ceram. Soc. 2017, 100, 3042–3049. [Google Scholar] [CrossRef]
  8. Ni, L.; Chen, X.M. Dielectric relaxations and formation mechanism of giant dielectric constant step in CaCu3Ti4O12 ceramics. Appl. Phys. Lett. 2007, 91, 122905. [Google Scholar] [CrossRef]
  9. Ke, S.; Fan, H.; Huang, H. Dielectric relaxation in A2FeNbO6 (A = Ba, Sr, and Ca) perovskite ceramics. J. Electroceramics 2009, 22, 252–256. [Google Scholar] [CrossRef]
  10. Zhang, C.-h.; Zhang, K.; Xu, H.-x.; Song, Q.; Yang, Y.-t.; Yu, R.-h.; Xu, D.; Cheng, X.-n. Microstructure and electrical properties of sol–gel derived Ni-doped CaCu3Ti4O12 ceramics. Trans. Nonferrous Met. Soc. China 2012, 22, s127–s132. [Google Scholar] [CrossRef]
  11. Shi, Y.; Hao, W.; Wu, H.; Sun, L.; Cao, E.; Zhang, Y.; Peng, H. High dielectric-permittivity properties of NaCu3Ti3Sb0.5Nb0.5O12 ceramics. Ceram. Int. 2015, 42, 116–121. [Google Scholar] [CrossRef]
  12. Wang, Z.; Cao, M.; Zhang, Q.; Hao, H.; Yao, Z.; Wang, Z.; Song, Z.; Zhang, Y.; Hu, W.; Liu, H. Dielectric Relaxation in Zr-Doped SrTiO3 Ceramics Sintered in N2 with Giant Permittivity and Low Dielectric Loss. J. Am. Ceram. Soc. 2015, 98, 476–482. [Google Scholar] [CrossRef]
  13. Liu, Y.; Zhao, X.; Zhang, C. Dielectric and impedance characteristics of Na1/3Ca1/3Y1/3Cu3Ti4O12 ceramic prepared by solid-state reaction method. J. Mater. Sci. Mater. Electron. 2016, 27, 11757–11761. [Google Scholar] [CrossRef]
  14. Nautiyal, A.; Autret, C.; Honstettre, C.; De Almeida-Didry, S.; Amrani, M.E.; Roger, S.; Negulescu, B.; Ruyter, A. Local analysis of the grain and grain boundary contributions to the bulk dielectric properties of Ca(Cu3−yMgy)Ti4O12 ceramics: Importance of the potential barrier at the grain boundary. J. Eur. Ceram. Soc. 2016, 36, 1391–1398. [Google Scholar] [CrossRef]
  15. Jumpatam, J.; Thongbai, P.; Yamwong, T.; Maensiri, S. Effects of Bi3+ doping on microstructure and dielectric properties of CaCu3Ti4O12/CaTiO3 composite ceramics. Ceram. Int. 2015, 41, S498–S503. [Google Scholar] [CrossRef]
  16. Hu, W.; Liu, Y.; Withers, R.L.; Frankcombe, T.J.; Noren, L.; Snashall, A.; Kitchin, M.; Smith, P.; Gong, B.; Chen, H.; et al. Electron-pinned defect-dipoles for high-performance colossal permittivity materials. Nat. Mater. 2013, 12, 821–826. [Google Scholar] [CrossRef]
  17. Nachaithong, T.; Kidkhunthod, P.; Thongbai, P.; Maensiri, S. Surface barrier layer effect in (In + Nb) co-doped TiO2 ceramics: An alternative route to design low dielectric loss. J. Am. Ceram. Soc. 2017, 100, 1452–1459. [Google Scholar] [CrossRef]
  18. Li, J.; Li, F.; Zhuang, Y.; Jin, L.; Wang, L.; Wei, X.; Xu, Z.; Zhang, S. Microstructure and dielectric properties of (Nb + In) co-doped rutile TiO2 ceramics. J. Appl. Phys. 2014, 116, 074105. [Google Scholar] [CrossRef]
  19. Zhang, L.; Tang, Z.-J. Polaron relaxation and variable-range-hopping conductivity in the giant-dielectric-constant material CaCu3Ti4O12. Phys. Rev. B 2004, 70, 174306. [Google Scholar] [CrossRef]
  20. Lunkenheimer, P.; Fichtl, R.; Ebbinghaus, S.; Loidl, A. Nonintrinsic origin of the colossal dielectric constants in CaCu3Ti4O12. Phys. Rev. B 2004, 70, 172102. [Google Scholar] [CrossRef] [Green Version]
  21. Wei, X.; Jie, W.; Yang, Z.; Zheng, F.; Zeng, H.; Liu, Y.; Hao, J. Colossal permittivity properties of Zn,Nb co-doped TiO2 with different phase structures. J. Mater. Chem. C 2015, 3, 11005–11010. [Google Scholar] [CrossRef]
  22. Thongyong, N.; Tuichai, W.; Chanlek, N.; Thongbai, P. Effect of Zn2+ and Nb5+ co-doping ions on giant dielectric properties of rutile-TiO2 ceramics. Ceram. Int. 2017, 43, 15466–15471. [Google Scholar] [CrossRef]
  23. Nachaithong, T.; Thongbai, P.; Maensiri, S. Colossal permittivity in (In1/2Nb1/2)xTi1−xO2 ceramics prepared by a glycine nitrate process. J. Eur. Ceram. Soc. 2017, 37, 655–660. [Google Scholar] [CrossRef]
  24. Zhu, X.; Yang, L.; Li, J.; Jin, L.; Wang, L.; Wei, X.; Xu, Z.; Li, F. The dielectric properties for (Nb,In,B) co-doped rutile TiO2 ceramics. Ceram. Int. 2017, 43, 6403–6409. [Google Scholar] [CrossRef]
  25. Zhang, J.; Yue, Z.; Zhou, Y.; Peng, B.; Zhang, X.; Li, L. Temperature-dependent dielectric properties, thermally-stimulated relaxations and defect-property correlations of TiO2 ceramics for wireless passive temperature sensing. J. Eur. Ceram. Soc. 2016, 36, 1923–1930. [Google Scholar] [CrossRef]
  26. Nachaithong, T.; Tuichai, W.; Kidkhunthod, P.; Chanlek, N.; Thongbai, P.; Maensiri, S. Preparation, characterization, and giant dielectric permittivity of (Y3+ and Nb5+) co–doped TiO2 ceramics. J. Eur. Ceram. Soc. 2017, 37, 3521–3526. [Google Scholar] [CrossRef]
  27. Yang, C.; Wei, X.; Hao, J. Disappearance and recovery of colossal permittivity in (Nb+Mn) co-doped TiO2. Ceram. Int. 2018, 11, 12395–12400. [Google Scholar] [CrossRef]
  28. Hu, W.; Lau, K.; Liu, Y.; Withers, R.L.; Chen, H.; Fu, L.; Gong, B.; Hutchison, W. Colossal Dielectric Permittivity in (Nb+Al) Codoped Rutile TiO2 Ceramics: Compositional Gradient and Local Structure. Chem. Mater. 2015, 27, 4934–4942. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of single doped (a) Zn, (b) Nb in TiO2 powder and (Zn0.33Nb0.67)xTi1−xO2 powder with x = (c) 0.5%, (d) 2.5% and (e) 0.5% ZNTO, (f) 2.5% ZNTO ceramics.
Figure 1. XRD patterns of single doped (a) Zn, (b) Nb in TiO2 powder and (Zn0.33Nb0.67)xTi1−xO2 powder with x = (c) 0.5%, (d) 2.5% and (e) 0.5% ZNTO, (f) 2.5% ZNTO ceramics.
Molecules 27 06121 g001
Figure 2. Surface morphologies of the (Zn0.33Nb0.67)xTi1−xO2 ceramics with x = 0.5 and 2.5%.
Figure 2. Surface morphologies of the (Zn0.33Nb0.67)xTi1−xO2 ceramics with x = 0.5 and 2.5%.
Molecules 27 06121 g002
Figure 3. Raman spectra of rutile-TiO2, single doped of Nb5+ and Zn2+ and ZNTO ceramics with difference codoping levels.
Figure 3. Raman spectra of rutile-TiO2, single doped of Nb5+ and Zn2+ and ZNTO ceramics with difference codoping levels.
Molecules 27 06121 g003
Figure 4. The elements mapping of 2.5% ZNTO ceramic; the Zn, Nb, Ti and O dopants are homogeneously dispersed in the grains and grain boundaries.
Figure 4. The elements mapping of 2.5% ZNTO ceramic; the Zn, Nb, Ti and O dopants are homogeneously dispersed in the grains and grain boundaries.
Molecules 27 06121 g004
Figure 5. Frequency dependence of (a) ε′ and (b) tanδ at RT for ZNTO ceramics with various codopant concentrations (x = 0.5%, 0.25%).
Figure 5. Frequency dependence of (a) ε′ and (b) tanδ at RT for ZNTO ceramics with various codopant concentrations (x = 0.5%, 0.25%).
Molecules 27 06121 g005
Figure 6. Frequency dependence of ε′ at RT of the as-fired sample, polished-sample and annealed-sample (in air) for (Zn0.33Nb0.67)xTi1−xO2 ceramics with x = 0.5% (a), 2.5% (b); inset shows frequency dependence of tanδ at RT.
Figure 6. Frequency dependence of ε′ at RT of the as-fired sample, polished-sample and annealed-sample (in air) for (Zn0.33Nb0.67)xTi1−xO2 ceramics with x = 0.5% (a), 2.5% (b); inset shows frequency dependence of tanδ at RT.
Molecules 27 06121 g006
Figure 7. Temperature dependence of the (a,c) dielectric relaxation and (b,d) tanδ at different frequencies for ZNTO ceramics with x = 0.5 and 2.5% respectively.
Figure 7. Temperature dependence of the (a,c) dielectric relaxation and (b,d) tanδ at different frequencies for ZNTO ceramics with x = 0.5 and 2.5% respectively.
Molecules 27 06121 g007
Figure 8. Arrhenius plots of the dielectric relaxation process for 0.5% and 2.5% ZNTO ceramics.
Figure 8. Arrhenius plots of the dielectric relaxation process for 0.5% and 2.5% ZNTO ceramics.
Molecules 27 06121 g008
Figure 9. Energy-preferable structure of the 2ZnVo triangular defect and 2Nb diamond defect.
Figure 9. Energy-preferable structure of the 2ZnVo triangular defect and 2Nb diamond defect.
Molecules 27 06121 g009
Table 1. Lattice parameter of ZNTO powders and ceramics with difference codoping levels.
Table 1. Lattice parameter of ZNTO powders and ceramics with difference codoping levels.
SampleLattice Parameter (Å)
ac
(a)0.5% Zn-TiO2 powder4.59612.9618
(b)0.5% Nb-TiO2 powder4.59642.9619
(c)(Zn0.33Nb0.67)xTi1−xO2 powder with x = 0.5%4.59712.9621
(d)(Zn0.33Nb0.67)xTi1−xO2 powder with x = 2.5%4.59942.9624
(e)0.5% ZNTO4.59672.9625
(f)2.5% ZNTO4.59972.9647
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nachaithong, T.; Moontragoon, P.; Thongbai, P. Dielectric Responses of (Zn0.33Nb0.67)xTi1−xO2 Ceramics Prepared by Chemical Combustion Process: DFT and Experimental Approaches. Molecules 2022, 27, 6121. https://doi.org/10.3390/molecules27186121

AMA Style

Nachaithong T, Moontragoon P, Thongbai P. Dielectric Responses of (Zn0.33Nb0.67)xTi1−xO2 Ceramics Prepared by Chemical Combustion Process: DFT and Experimental Approaches. Molecules. 2022; 27(18):6121. https://doi.org/10.3390/molecules27186121

Chicago/Turabian Style

Nachaithong, Theeranuch, Pairot Moontragoon, and Prasit Thongbai. 2022. "Dielectric Responses of (Zn0.33Nb0.67)xTi1−xO2 Ceramics Prepared by Chemical Combustion Process: DFT and Experimental Approaches" Molecules 27, no. 18: 6121. https://doi.org/10.3390/molecules27186121

Article Metrics

Back to TopTop