Next Article in Journal
Precipitation Evolution in the Austenitic Heat-Resistant Steel HR3C upon Creep at 700 °C and 750 °C
Next Article in Special Issue
Broadband Electromagnetic Absorption Effect of Topological Structure Using Carbon Nanotube Based Hybrid Material
Previous Article in Journal
Simultaneous Substitution of Fe and Sr in Beta-Tricalcium Phosphate: Synthesis, Structural, Magnetic, Degradation, and Cell Adhesion Properties
Previous Article in Special Issue
Dual-Scale Porosity Alumina Structures Using Ceramic/Camphene Suspensions Containing Polymer Microspheres
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Colossal Permittivity Characteristics of (Nb, Si) Co-Doped TiO2 Ceramics

1
Department of Physics, College of Science, King Faisal University, P.O. Box 400, Al-Ahsa 31982, Saudi Arabia
2
Physics Department, Faculty of Science, Assiut University, Assiut 71516, Egypt
3
Department of Physics, Faculty of Science, The New Valley University, El-Kharga 72511, Egypt
*
Author to whom correspondence should be addressed.
Materials 2022, 15(13), 4701; https://doi.org/10.3390/ma15134701
Submission received: 8 June 2022 / Revised: 24 June 2022 / Accepted: 1 July 2022 / Published: 5 July 2022
(This article belongs to the Special Issue High-Performance Structural Ceramics and Hybrid Materials)

Abstract

:
(Nb5+, Si4+) co-doped TiO2 (NSTO) ceramics with the compositions (Nb0.5Si0.5)xTi1−xO2, x = 0, 0.025, 0.050 and 0.1 were prepared with a solid-state reaction technique. X-ray diffraction (XRD) patterns and Raman spectra confirmed that the tetragonal rutile is the main phase in all the ceramics. Additionally, XRD revealed the presence of a secondary phase of SiO2 in the co-doped ceramics. Impedance spectroscopy analysis showed two contributions, which correspond to the responses of grain and grain-boundary. All the (Nb, Si) co-doped TiO2 showed improved dielectric performance in the high frequency range (>103 Hz). The sample (Nb0.5Si0.5)0.025Ti0.975O2 showed the best dielectric performance in terms of higher relative permittivity (5.5 × 104) and lower dielectric loss (0.18), at 10 kHz and 300 K, compared to pure TiO2 (1.1 × 103, 0.34). The colossal permittivity of NSTO ceramics is attributed to an internal barrier layer capacitance (IBLC) effect, formed by insulating grain-boundaries and semiconductor grains in the ceramics.

1. Introduction

Titanium dioxide (TiO2) is an interesting inorganic material used in several important applications such as photocatalysis, gas sensing and medical applications, due to its chemical stability at room temperature and its wide band gap [1]. The room temperature relative permittivity (ε′) of pure TiO2 is ~100, which is not suitable for practical energy storage applications. Interestingly, recent studies reported colossal relative permittivity (CP) for co-doped TiO2 in the form (A, B)xTi1−xO2, where A is pentavalent and B is tri- or bivalent cations [2,3,4,5,6]. The origin of CP of the co-doped TiO2 ceramics is controversially discussed in the literature. Several models were proposed to explain the colossal relative permittivity and low dielectric loss (tanδ = ε″/ε′, where ε′ and ε″ are the real and imaginary part of the complex relative permittivity) of co-doped TiO2. It seems there is no universal model that can be applied for all cases. For example, the electron-pinned defect dipole (EPDD) model was successful for (Nb5+, In3+) co-doped TiO2 [7,8,9,10]. This model suggests that the electrons produced by the donors (Nb5+) would reduce Ti4+ to Ti3. In these conditions, defect complexes such as the triangular shaped In 2 3 + V O Ti 3 + are created in the rutile structure of TiO2 and are able to inhibit the free electrons from long-range motion. Therefore, CP and low tanδ were expected for these co-doped TiO2. In addition to the EPDD model, the internal barrier layer capacitance (IBLC) model [11,12] has also been suggested as a possible reason for the colossal permittivity of (donor, acceptor) co-doped TiO2 [2,13]. The IBLC model is widely accepted for the interpretation of CP of electrically heterogeneous ceramics. In IBLC, conducting grains are separated by the electrically insulating grain-boundaries. The applied alternating voltage on the sample displaces the charge carriers from the semiconductor grains where they pile up at the thin resistive grain boundaries. The capacitors, thus formed, lead to a colossal effective relative permittivity (ε′) of the ceramic. Considering the IBLC model, the use of acceptors is not compulsory in order to obtain CP in co-doped TiO2. For example, Yang et al. reported on the colossal permittivity in (Nb5+, Zr4+) co-doped TiO2 [14]. Doping of CCTO with SiO2 has been reported as promoting the dielectric properties of CCTO, due to abnormal grain growth [15]. Doping of NiO with Si4+ was effective in increasing the grain boundary resistance due to the formation of a Si-rich secondary phase at the grain-boundaries, which resulted in decreasing the dielectric loss in co-doped NiO ceramics [16]. The phase purity of co-doped TiO2 is dependent on the solubility of dopants in TiO2, which can be predicted by the Hume–Rothery rule [14,17]. According to this rule, a substitutional-solid solution is more likely to form if the difference between the ionic radius of the doping element and the matrix element is less than 15%. In cases where an ionic radius mismatch is greater than 15%, the formation of an interstitial solid solution is more favorable. Most studies opted to choose dopants of ionic radii close to the Ti4+ to avoid the formation of a secondary phase in the prepared ceramics. Nevertheless, the effect of a secondary phase on the dielectric properties of ceramics with a targeted crystalline phase is not easy to predict in most cases [18,19]. The smaller ionic radius of Si4+ (radius = 40 pm) compared to Ti4+ (radius = 60.5 pm) would result in a secondary phase in the resulting ceramics. In the present work, we investigated the dielectric and structural properties of (Nb0.5Si0.5)xTi1−xO2 (x = 0, 0.025, 0.050 and 0.1) over a wide range of frequencies and temperatures. It was found that NSTO ceramics have improved dielectric performance compared to pure TiO2 while the sample (Nb0.5Si0.5)0.025Ti0.975O2 ceramic showed the best improved dielectric performance.

2. Experiment Procedure

2.1. Materials

Anatase TiO2 (99.8%, Aldrich, St. Louis, MI, USA), Nb2O5 (99.99%, Aldrich) and SiO2 (99.99%, Aldrich) were used for the synthesis of the powder of (Nb0.5Si0.5)xTi1−xO2, x = 0, 0.025, 0.050 and 0.1. These samples will be referred to as NSTO.

2.2. Synthesis of (Nb0.5Si0.5)xTi1−xO2 Ceramics

Stoichiometric amounts of the elemental oxides were weighed according to the molar ratio of (Nb0.5Si0.5)xTi1−xO2, then milled in a Retsch PM400 machine for 20 h with a rotation speed of 200 rpm using pots and balls made of tungsten carbide. The obtained powder was pressed (320 MPa) into pellets, 10 mm in diameter and 2 mm in thickness. Subsequently, the pellets were sintered in air inside an electric tubular furnace at a temperature of 1500 °C for 10 h.

2.3. Characterization Methods

Field-emission scanning electron microscope (FE-SEM) (Joel, Tokyo, Japan, SM7600F) and powder X-ray diffraction (XRD) (Stoe and Cie GmbH, Darmstadt, Germany, Cu Kα1 radiation) techniques were used to study the morphology and phase purity of the prepared ceramics. XRD measurements were taken across the range 20° ≤ 2θ ≤ 80° with steps of 0.02°. Raman spectra were recorded with a fully integrated confocal Raman microscope (HORIBA Scientific, Piscataway, NJ, USA) in a back-scattering configuration. All the samples were excited by a He-Cd laser (wavelength 442 nm, 100 mW). The number of gratings in the Raman spectrometer was 1800 L/mm. Impedance spectroscopy (IS) measurements were conducted in dry a nitrogen atmosphere using the system turnkey concept 50 from Novocontrol in the temperature range 200–400 K over the frequency range 1 Hz–1 MHz at an oscillation voltage of 0.1 V. The temperature of the sample was automatically controlled by Quatro Cryosystem. Silver paint was applied to both sides of each pellet before the electrical measurements were taken.

3. Results and Discussion

The XRD patterns of the sintered NSTO ceramics are shown in Figure 1. All the samples present the tetragonal rutile TiO2 (JCPDS 21-1276) as the main phase. The lattice parameters as well as the cell volume were calculated for each sample from the angle and hkl values of the main diffraction peaks using UNITCELL software. These values are summarized in Table 1. Compared to the un-doped TiO2, (Nb, Si) co-doped ceramics (x > 0) show higher values for lattice parameters and the cell volume. These results highlight the effect of the substitution of small Ti4+ ions (radius = 60.5 pm) by the larger Nb5+ ions (radius = 64 pm) in the TiO2 matrix. Moreover, a tiny impurity peak is observed in the XRD of doped samples at 21.82° and corresponds to SiO2 (JCPDS 47-718). Compared to the (110) peak of the main TiO2 rutile phase, the intensity of the impurity peak is considerably low, but increases with increasing the doping content.
Raman spectroscopy is known for its usefulness in confirming the anatase or rutile phase of TiO2 [20]. Figure 2 represents the room temperature Raman spectroscopy over the range of 100–1100 cm−1 for NSTO ceramics. Four peaks are observed at ~143 cm−1, ~447 cm−1, ~612 cm−1 and ~826 cm−1. These peaks correspond to the typical first-order Raman-active modes of rutile TiO2: B1g, Eg, A1g and B2g, respectively [21]. In addition, a peak is observed at ~237 cm−1, which is a second-order scattering feature [21]. Therefore, Raman spectroscopy confirmed that all NSTO ceramics possess the rutile phase structure, which is in agreement with the XRD results. As seen in Figure 2, the Eg slightly shifted toward a lower wave number for the sample with x = 0.1, which might be due to the evolution of lattice distortion and the movement of oxygen along c-axis with the increase in dopants content [22].
Figure 3 shows FE-SEM micrographs of the scratched surface of NSTO ceramics. The grains and grain-boundaries are clearly seen for all the samples. Using the lineal intercept method [23,24], the average grain size was found to be 9.1 ± 1.6, 5.6 ± 1.1, 4.8 ± 0.9 and 9.4 ± 1.6 µm for the samples with x = 0, 0.025, 0.05 and 0.1, respectively. Therefore, the average grain size tends to decrease with increasing the dopants content up to x = 0.05. Additionally, the grain in the co-doped samples seems to be composed of smaller sub-grains as shown by the red-dashed line in the SEM micrograph of the sample x = 0.025 (Figure 3). Similar observations on the effect of doping content on the grain size have been reported [25,26,27]. This behavior was attributed to the effect of the secondary phase in inhibiting grain growth [25] and/or the increased number of nuclei centers with increasing doping content [27].
Figure 4 shows the frequency dependency of relative permittivity (ε′) and dielectric loss (tanδ) of NSTO ceramics. It can be seen that (Nb, Si) co-doped TiO2 (x > 0) ceramics have improved dielectric properties in terms of higher ε′ and lower tanδ compared to the un-doped sample (x = 0). Among these ceramics, the sample with x = 0.025 showed the minimum tanδ of ~0.11 and colossal permittivity (ε′~4.8 × 104) at 10 kHz and room temperature. This dielectric performance is close to that of (Nb, La) [25], (Ta, In) [6] and (Nb, Eu) [28] co-doped TiO2 but lower than (Ta, Mg) [29], (Ta, Y) [30] (Ta, Ho) [31] (Nb, Zr) [14] co-doped TiO2.
Figure 5 depicts the room temperature complex impedance (Z* = Z′ + iZ″) plots for the four ceramic samples. The impedance spectrum of every sample is comprised of a large semicircular arc that covers most of the frequency range. However, a close examination at the high-frequency region reveals that the aforementioned arc does not pass through the origin. A second circular arc and non-zero intercept with the real Z′ axis is observed in the high frequency region for the samples with x = 0 and x > 0, respectively. This behavior is similar to what is reported for CaCu3Ti4O12 [32] and co-doped TiO2 [2,13]. The arc at high frequency (/non zero intercept) is attributed to the response of the semiconductor component (grains), while the larger arc represents the response of electrically insulating component (grain-boundaries). The resistances of grains (Rg) and grain-boundaries (Rgb) can be calculated from the intercepts of the corresponding arcs with the axis of the real part of the impedance (Z′) [11]. The room temperature values of Rg and Rgb are summarized in Table 2. It can be seen that the NSTO ceramics have an electrically heterogeneous structure of semiconducting grains and resistive grain-boundaries. It is observed that the change in Rgb with increasing doping content is more pronounced than in Rg. This result indicates that (Nb5+, Si4+) co-doping affects, principally, the grain-boundary resistance of TiO2. Moreover, Rg and Rgb decrease gradually with increasing doping content because of the increase in the concentration of free carriers as a result of doping. The sintering temperature is reported to affect the dielectric performance of pure rutile TiO2 [18,33,34]. For example, ε′ was found to increase from 400 to 1.3 × 104 with increasing the sintering temperature from 1200 to 1450 °C [33]. In our case, the undoped TiO2 has a performance comparable to that found in the literature for the rutile TiO2, which is sintered at high temperature [18]. In this reference, the dielectric constant and tanδ of pure TiO2 ceramics at 1 kHz and room temperature were found to be 103 and ~0.4, respectively. As seen in Figure 4, the co-doped NSTO samples (x > 0) have comparatively higher relative permittivity (ε′) and lower dielectric loss (tanδ), in the frequency range > 103 Hz, than the pure TiO2 (x = 0). As shown by the FE-SEM micrograph in Figure 3, the grain of the co-doped ceramic is composed of smaller grains. Therefore, we believe that the co-doping results in more grain-boundary barriers in front of the charge carriers, which causes the reduced tanδ values for co-doped ceramics in the high frequency range.
Figure 6 shows the variation in the imaginary part of the impedance (Z″) with frequency at selected temperatures for NSTO ceramics. At a given temperature, the spectrum of Z″ shows a peak with maximum value Z max at a frequency fmax. With increasing temperature, Z max decreases while fmax shifts to higher frequencies. According to the inter barrier layer capacitance (IBLC), the dependence of the imaginary part of the impedance (Z″) on the angular frequency (ω) is given by [35,36]:
Z ω = R g ω R g C g 1 + ω R g C g 2 + R g . b ω R gb C gb 1 + ω R gb C gb 2
where Rg (/Rgb) and Cg (/Cgb) are the resistances and capacitances of the grain (/grain-boundaries), respectively. Consequently, for electrically heterogeneous materials with Rgb >> Rg, the overall resistance of the sample R ≈ Rgb and can be calculated from Z max by the relationship:
Z max = R gb 2
Moreover, the relaxation time (τ) is related to the frequency fmax by the relationship:
τ = 1 ω = 1 2 π f max
The activation energy for conduction (E) and for relaxation (U) were calculated from the Arrhenius plots in Figure 7 using the relationships:
σ = σ 0   e E K B T
where σ0 is the pre-exponential factor, KB is Boltzmann constant and T is absolute temperature.
τ =   τ 0   e U k B T
where τ0 is the pre-factor and U is the activation energy for the relaxation process.
The calculated values of activation energies from the spectra of M″ are given in Table 3. It can be seen that the co-doped NSTO samples (x > 0) have activation-energy values considerably lower than the pure TiO2 (x = 0), which correlate with the lower resistivity of these samples.
Moreover, the modulus formalism has been deployed to further study the conductivity relaxation behavior of the NSTO ceramics. Figure 8 shows the spectra of the imaginary part M″ of the electric modulus M* = M′ + iM″ = 1/ε*, where ε* is the complex relative permittivity. At room temperature, the spectrum of M″ of the un-doped TiO2 (x = 0) shows two resolved peaks. For the co-doped samples (x > 0), only the low frequency peak is observed within the studied frequency range. The low and high frequency M″ peaks are attributed to the contribution of grain-boundary and grain, respectively. Furthermore, the observed peaks shift towards higher frequency with increasing temperature. The relaxation time τ is related to the peak frequency (fmax) as τ = 1/2πfmax. The insets of Figure 4 show the log(τ) versus 1000/T curves, which are fitted according to the Arrhenius equation. The calculated activation energy of the grain-boundary is 0.351 eV, 0.093 eV, 0.089 eV and 0.112 eV for the samples with x = 0, 0.025, 0.05 and 0.1, respectively. These values are comparable to the activation energies obtained from the spectra of Z″, as previously presented in Table 3.
The activation-energy values for the doped ceramics (x > 0) are comparable to the value of 0.1 eV reported for the dipolar relaxation due to electrons hopping between Ti3+ and Ti4+ ions in co-doped TiO2 [37]. Substituting of Nb5+ for Ti4+ resulting in a distorted lattice and the reduction of some Ti4+ to Ti3+. This process may be expressed using Kroger-Vink notations as follows [38,39]:
Nb 2 O 5 + 2 TiO 2 = 2 Nb Ti + 2 Ti Ti + 8 O O + 1 2 O 2
2 Ti 4 + + 2 e = 2 Ti 3 +

4. Conclusions

(Nb0.5Si0.5)xTi1−xO2, (x = 0, 0.025, 0.050 and 0.1) (NSTO) ceramics were fabricated using a conventional, high-temperature, solid state reaction technique. X-ray diffraction (XRD) patterns and Raman spectra confirmed that the tetragonal rutile is the main phase in all the ceramics. Additionally, XRD revealed the presence of a small amount of secondary phase of SiO2 in the co-doped ceramics. NSTO ceramics showed room temperature colossal relative permittivity (>104) over the studied frequency range (1–1 MHz). (Nb, Si) co-doped (Nb0.5Si0.5)0.025Ti0.975O2 ceramics showed the most improved dielectric performance as higher ε′ (~5.5 × 104) and lower tanδ (~0.18), at 10 kHz and 300 K, compared to pure TiO2 (1.1 × 103, 0.34). The colossal permittivity of NSTO ceramics is attributed to an internal barrier layer capacitance (IBLC) effect, formed by insulating grain-boundaries and semiconductor grains in the ceramics. The electrical heterogeneity of the NSTO ceramics was confirmed by the impedance-spectroscopy analysis. The activation-energy values for conduction and relaxation of the co-doped NSTO ceramics are close to 0.1 eV which is reported for the dipolar relaxation due to electrons hopping between Ti3+ and Ti4+ ions in co-doped TiO2

Author Contributions

Conceptualization, H.M.K.; Formal analysis, H.M.K., M.M.A. and J.M.; Funding acquisition, H.M.K.; Investigation, H.M.K., M.M.A. and A.A.; Project administration, H.M.K. and M.M.A.; Software, H.M.K., J.M. and N.M.S.; Validation, M.M.A.; Writing—original draft, H.M.K. and M.M.A.; Writing—review and editing, H.M.K., M.M.A. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported through the Annual Funding track by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia [Project No. AN000101].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available up on request.

Acknowledgments

The authors thank the DSR at King Faisal University for financial and technical support (Project No. AN000101).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, H.; Zhu, W.; Zhang, Z.; Cai, W.; Zhou, X. Er and Mg co-doped TiO2 nanorod arrays and improvement of photovoltaic property in perovskite solar cell. J. Alloys Compd. 2018, 771, 649–657. [Google Scholar] [CrossRef]
  2. Fan, J.; Cheng, Y.; Zheng, L.; Tong, L.; Hu, Z.; He, G. Giant dielectric response and relaxation behavior in (Tm + Ta) co-doped TiO2 ceramics. Phys. Chem. Chem. Phys. 2022, 24, 4759–4768. [Google Scholar] [CrossRef] [PubMed]
  3. Siriya, P.; Chanlek, N.; Srepusharawoot, P.; Thongbai, P. Excellent giant dielectric properties over wide temperatures of (Al, Sc)3+ and Nb5+ doped TiO. Results Phys. 2022, 36, 105458. [Google Scholar] [CrossRef]
  4. Siriya, P.; Pengpad, A.; Srepusharawoot, P.; Chanlek, N.; Thongbai, P. Improved microstructure and significantly enhanced dielectric properties of Al(3+)/Cr(3+)/Ta(5+) triple-doped TiO2 ceramics by Re-balancing charge compensation. RSC Adv. 2022, 12, 4946–4954. [Google Scholar] [CrossRef]
  5. Thanamoon, N.; Chanlek, N.; Moontragoon, P.; Srepusharawoot, P.; Thongbai, P. Microstructure, low loss tangent, and excellent temperature stability of Tb+Sb-doped TiO2 with high dielectric permittivity. Results Phys. 2022, 37, 105536. [Google Scholar] [CrossRef]
  6. Wang, Z.; Peng, P.; Zhang, L.; Wang, N.; Tang, B.; Cui, B.; Liu, J.; Xu, D. Effect of electric field on the microstructure and electrical properties of (In + Ta) co-doped TiO2 colossal dielectric ceramics. J. Mater. Sci. Mater. Electron. 2022, 33, 6283–6293. [Google Scholar] [CrossRef]
  7. Hu, W.; Liu, Y.; Withers, R.; Frankcombe, T.J.; Norén, 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]
  8. Mandal, S.; Pal, S.; Kundu, A.K.; Menon, K.S.R.; Hazarika, A.; Rioult, M.; Belkhou, R. Direct view at colossal permittivity in donor-acceptor (Nb, In) co-doped rutile TiO2. Appl. Phys. Lett. 2016, 109, 092906. [Google Scholar] [CrossRef]
  9. Tsuji, K.; Han, H.; Guillemet-Fritsch, S.; Randall, C.A. Dielectric relaxation and localized electron hopping in colossal dielectric (Nb,In)-doped TiO2 rutile nanoceramics. Phys. Chem. Chem. Phys. 2017, 19, 8568–8574. [Google Scholar] [CrossRef] [Green Version]
  10. Zhao, L.; Wang, J.; Gai, Z.; Li, J.; Liu, J.; Wang, J.; Wang, C.; Wang, X. Annealing effects on the structural and dielectric properties of (Nb + In) co-doped rutile TiO2 ceramics. RSC Adv. 2019, 9, 8364–8368. [Google Scholar] [CrossRef] [Green Version]
  11. Sinclair, D.C.; Adams, T.B.; Morrison, F.; West, A. CaCu3Ti4O12: One-step internal barrier layer capacitor. Appl. Phys. Lett. 2002, 80, 2153–2155. [Google Scholar] [CrossRef]
  12. Adams, T.; Sinclair, D.; West, A. Giant Barrier Layer Capacitance Effects in CaCu3Ti4O12 Ceramics. Adv. Mater. 2002, 14, 1321–1323. [Google Scholar] [CrossRef]
  13. Mingmuang, Y.; Chanlek, N.; Thongbai, P. Ultra–Low Loss Tangent and Giant Dielectric Permittivity with Excellent Temperature Stability of TiO2 Co-Doped with Isovalent-Zr4+/Pentavalent-Ta5+ Ions. J. Mater. 2022, in press. [Google Scholar] [CrossRef]
  14. Yang, C.; Wei, X.; Hao, J. Colossal permittivity in TiO2 co-doped by donor Nb and isovalent Zr. J. Am. Ceram. Soc. 2017, 101, 307–315. [Google Scholar] [CrossRef]
  15. Kim, K.-M.; Kim, S.-J.; Lee, J.-H.; Kim, D.-Y. Microstructural evolution and dielectric properties of SiO2-doped CaCu3Ti4O12 ceramics. J. Eur. Ceram. Soc. 2007, 27, 3991–3995. [Google Scholar] [CrossRef]
  16. Kotb, H.M.; Ahmad, M.M.; Alraheem, N.A. Study of the structural, impedance spectroscopy and dielectric properties of Na and Si co-doped NiO ceramics. J. Phys. D Appl. Phys. 2017, 50, 435304. [Google Scholar] [CrossRef] [Green Version]
  17. Peng, P.; Chen, C.; Cui, B.; Li, J.; Xu, D.; Tang, B. Influence of the electric field on flash-sintered (Zr + Ta) co-doped TiO2 colossal permittivity ceramics. Ceram. Int. 2021, 48, 6016–6023. [Google Scholar] [CrossRef]
  18. Cheng, X.; Li, Z.; Wu, J. Colossal permittivity in ceramics of TiO2 Co-doped with niobium and trivalent cation. J. Mater. Chem. A 2015, 3, 5805–5810. [Google Scholar] [CrossRef]
  19. Song, Y.-C.; Liu, P.; Zhao, X.; Guo, B.; Cui, X. Dielectric properties of (Bi0.5Nb0.5) Ti1-O2 ceramics with colossal permittivity. J. Alloys Compd. 2017, 722, 676–682. [Google Scholar] [CrossRef]
  20. Hanaor, D.A.H.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2010, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  21. Tuichai, W.; Danwittayakul, S.; Manyam, J.; Chanlek, N.; Takesada, M.; Thongbai, P. Giant dielectric properties of Ga3+–Nb5+ Co-doped TiO2 ceramics driven by the internal barrier layer capacitor effect. Materialia 2021, 18, 101175. [Google Scholar] [CrossRef]
  22. Zhao, X.-G.; Liu, P. Dielectric and electric relaxations induced by the complex defect clusters in (Yb+Nb) co-doped rutile TiO2 ceramics. J. Am. Ceram. Soc. 2017, 100, 3505–3513. [Google Scholar] [CrossRef]
  23. Wurst, J.C.; Nelson, J.A. Lineal Intercept Technique for Measuring Grain Size in Two-Phase Polycrystalline Ceramics. J. Am. Ceram. Soc. 1972, 55, 109. [Google Scholar] [CrossRef]
  24. Chen, G.-H.; Li, J.-L.; Chen, X.; Kang, X.-L.; Yuan, C.-L. Sintering temperature dependence of varistor properties and impedance spectroscopy behavior in ZnO based varistor ceramics. J. Mater. Sci. Mater. Electron. 2015, 26, 2389–2396. [Google Scholar] [CrossRef]
  25. Wang, X.W.; Zheng, Y.P.; Liang, B.K.; Zhang, G.; Shi, Y.C.; Zhang, B.H.; Xue, L.L.; Shang, S.Y.; Shang, J.; Yin, S.Q.; et al. Preparation and properties of La and Nb co-doped TiO2 colossal dielectric ceramic materials. J. Mater. Sci. Mater. Electron. 2020, 31, 16044–16052. [Google Scholar] [CrossRef]
  26. Jiao, L.; Guo, P.; Kong, D.; Huang, X.; Li, H. Dielectric properties of (Yb0.5Ta0.5)xTi1−xO2 ceramics with colossal permittivity and low dielectric loss. J. Mater. Sci. Mater. Electron. 2020, 31, 3654–3661. [Google Scholar] [CrossRef]
  27. Garai, M.; Reka, A.A.; Karmakara, B.; Mollaa, A.R. Microstructure–mechanical properties of Ag°/Au° doped K–Mg–Al–Si–O–F glass-ceramics. RSC Adv. 2021, 11, 11415–11424. [Google Scholar] [CrossRef]
  28. Wang, Z.; Chen, H.; Wang, T.; Xiao, Y.; Nian, W.; Fan, J. Enhanced relative permittivity in niobium and europium co-doped TiO2 ceramics. J. Eur. Ceram. Soc. 2018, 38, 3847–3852. [Google Scholar] [CrossRef]
  29. Thongyong, N.; Chanlek, N.; Srepusharawoot, P.; Thongbai, P. Origins of Giant Dielectric Properties with Low Loss Tangent in Rutile (Mg1/3Ta2/3)0.01Ti0.99O2 Ceramic. Molecules 2021, 26, 6952. [Google Scholar] [CrossRef]
  30. Fan, J.; Long, Z.; Hu, Z. Interface effects and defect clusters inducing thermal stability and giant dielectric response in (Ta+Y)-co-doped TiO2 ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 26232–26240. [Google Scholar] [CrossRef]
  31. Fan, J.; Long, Z.; Zhou, H.; He, G.; Hu, Z. Colossal dielectric behavior of (Ho, Ta) co-doped rutile TiO2 ceramics. J. Mater. Sci. Mater. Electron. 2021, 32, 14780–14790. [Google Scholar] [CrossRef]
  32. De Almeida-Didry, S.; Merad, S.; Autret-Lambert, C.; Nomel, M.M.; Lucas, A.; Gervais, F. A core-shell synthesis of CaCu3Ti4O12 (CCTO) ceramics showing colossal permittivity and low electric losses for application in capacitors. Solid State Sci. 2020, 109, 106431. [Google Scholar] [CrossRef]
  33. Li, Z.; Wu, J.; Xiao, D.; Zhu, J.; Wu, W. Colossal permittivity in titanium dioxide ceramics modified by tantalum and trivalent elements. Acta Mater. 2015, 103, 243–251. [Google Scholar] [CrossRef]
  34. Li, Z.; Luo, X.; Wu, W.; Wu, J. Niobium and divalent-modified titanium dioxide ceramics: Colossal permittivity and composition design. J. Am. Ceram. Soc. 2017, 100, 3004–3012. [Google Scholar] [CrossRef]
  35. Nachaithong, T.; Thongbai, P. Preparation, characterization, electrical properties and giant dielectric response in (In + Nb) co-doped TiO2 ceramics synthesized by a urea chemical-combustion method. J. Mater. Sci. Mater. Electron. 2017, 28, 10914–10920. [Google Scholar] [CrossRef]
  36. Liu, J.; Duan, C.-G.; Yin, W.-G.; Mei, W.N.; Smith, R.W.; Hardy, J.R. Large dielectric constant and Maxwell-Wagner relaxation in Bi2/3Cu3Ti4O12. Phys. Rev. B 2004, 70, 144106. [Google Scholar] [CrossRef]
  37. Si, R.J.; Li, T.Y.; Sun, J.; Wang, J.; Wang, S.T.; Zhu, G.B.; Wang, C.C. Humidity sensing behavior and its influence on the dielectric properties of (In + Nb) co-doped TiO2 ceramics. J. Mater. Sci. 2019, 54, 14645–14653. [Google Scholar] [CrossRef]
  38. Li, J.; Zeng, Y.; Fang, Y.; Chen, N.; Du, G.; Zhang, A. Synthesis of (La + Nb) co-doped TiO2 rutile nanoparticles and dielectric properties of their derived ceramics composed of submicron-sized grains. Ceram. Int. 2020, 47, 8859–8867. [Google Scholar] [CrossRef]
  39. Dong, W.; Hu, W.; Berlie, A.; Lau, K.; Chen, H.; Withers, R.L.; Liu, Y. Colossal Dielectric Behavior of Ga+Nb Co-Doped Rutile TiO2. ACS Appl. Mater. Interfaces 2015, 7, 25321–25325. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of the sintered NSTO ceramics.
Figure 1. The XRD patterns of the sintered NSTO ceramics.
Materials 15 04701 g001
Figure 2. Raman spectra of the sintered ceramics of NSTO.
Figure 2. Raman spectra of the sintered ceramics of NSTO.
Materials 15 04701 g002
Figure 3. FE-SEM morphology of NSTO ceramics.
Figure 3. FE-SEM morphology of NSTO ceramics.
Materials 15 04701 g003aMaterials 15 04701 g003b
Figure 4. Frequency dependence of ε′ and tanδ at room temperature for NSTO ceramics.
Figure 4. Frequency dependence of ε′ and tanδ at room temperature for NSTO ceramics.
Materials 15 04701 g004
Figure 5. Impedance complex plane plot (Z*) at room temperature for NSTO ceramics.
Figure 5. Impedance complex plane plot (Z*) at room temperature for NSTO ceramics.
Materials 15 04701 g005
Figure 6. Frequency dependence of −Z″ at different temperatures for NSTO ceramics.
Figure 6. Frequency dependence of −Z″ at different temperatures for NSTO ceramics.
Materials 15 04701 g006
Figure 7. The Arrhenius plots of the conductivity σ (left) and relaxation time τ (right) in the grain-boundaries for NSTO ceramics. The dashed line represents the line of best fit.
Figure 7. The Arrhenius plots of the conductivity σ (left) and relaxation time τ (right) in the grain-boundaries for NSTO ceramics. The dashed line represents the line of best fit.
Materials 15 04701 g007
Figure 8. Frequency dependence of M′′ at different temperatures for NSTO ceramics.
Figure 8. Frequency dependence of M′′ at different temperatures for NSTO ceramics.
Materials 15 04701 g008
Table 1. Lattice parameters and volume cell for NSTO ceramics.
Table 1. Lattice parameters and volume cell for NSTO ceramics.
Lattice Parameters (Å)Cell Volume (Å3)
ac
x = 04.592642.9566862.3634
x = 0.0254.597032.9585262.5214
x = 0.054.596692.9565562.4706
x = 0.014.606472.9617062.8476
Table 2. The room temperature values of grain resistivity (Rg), grain-boundary resistivity (Rgb), ε′ and tanδ at 1.1 kHz and the minimum dielectric loss value (tanδ)min for NSTO ceramics.
Table 2. The room temperature values of grain resistivity (Rg), grain-boundary resistivity (Rgb), ε′ and tanδ at 1.1 kHz and the minimum dielectric loss value (tanδ)min for NSTO ceramics.
Rg (Ω.cm)Rgb (Ω.cm)ε′tanδ(tanδ)min
at 1.1 kHz
x = 01.2 × 1049.0 × 1061.1 × 1030.650.34 (at 58 Hz)
x = 0.02515.33.0 × 1055.5 × 1040.180.11 (at 10 KHz)
x = 0.059.71.0 × 1056.0 × 1040.400.11 (at 24 KHz)
x = 0.135.94.3 × 1045.3 × 1041.370.13 (at 47 KHz)
Table 3. Activation-energy values for conduction (E) and for relaxation process (U) for NSTO ceramics.
Table 3. Activation-energy values for conduction (E) and for relaxation process (U) for NSTO ceramics.
E
(eV)
U
(eV)
x = 00.3380.341
x = 0.0250.0890.095
x = 0.050.0870.078
x = 0.10.1010.099
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kotb, H.M.; Alshoaibi, A.; Mazher, J.; Shaalan, N.M.; Ahmad, M.M. Colossal Permittivity Characteristics of (Nb, Si) Co-Doped TiO2 Ceramics. Materials 2022, 15, 4701. https://doi.org/10.3390/ma15134701

AMA Style

Kotb HM, Alshoaibi A, Mazher J, Shaalan NM, Ahmad MM. Colossal Permittivity Characteristics of (Nb, Si) Co-Doped TiO2 Ceramics. Materials. 2022; 15(13):4701. https://doi.org/10.3390/ma15134701

Chicago/Turabian Style

Kotb, Hicham Mahfoz, Adil Alshoaibi, Javed Mazher, Nagih M. Shaalan, and Mohamad M. Ahmad. 2022. "Colossal Permittivity Characteristics of (Nb, Si) Co-Doped TiO2 Ceramics" Materials 15, no. 13: 4701. https://doi.org/10.3390/ma15134701

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