Next Article in Journal
Stable and Oriented Liquid Crystal Droplets Stabilized by Imidazolium Ionic Liquids
Next Article in Special Issue
Concentration-Dependent Multi-Potentiality of L-Arginine: Antimicrobial Effect, Hydroxyapatite Stability, and MMPs Inhibition
Previous Article in Journal
Progress and Future Directions with Peptide-Drug Conjugates for Targeted Cancer Therapy
Previous Article in Special Issue
Effects of Temperature on Enantiomerization Energy and Distribution of Isomers in the Chiral Cu13 Cluster
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Significantly Improved Colossal Dielectric Properties and Maxwell—Wagner Relaxation of TiO2—Rich Na1/2Y1/2Cu3Ti4+xO12 Ceramics

by
Pariwat Saengvong
1,
Narong Chanlek
2,
Bundit Putasaeng
3,
Atip Pengpad
1,4,
Viyada Harnchana
1,4,
Sriprajak Krongsuk
1,4,
Pornjuk Srepusharawoot
1,4 and
Prasit Thongbai
1,4,*
1
Giant Dielectric and Computational Design Research Group (GD–CDR), Department of Physics, Faculty of Science, Khon Kaen University, Khon Kaen 40002, Thailand
2
Synchrotron Light Research Institute (Public Organization), 111 University Avenue, Muang District, Nakhon Ratchasima 30000, Thailand
3
National Metal and Materials Technology Center, National Science and Technology Development Agency, Thailand Science Park, Pathum Thani 12120, Thailand
4
Institute of Nanomaterials Research and Innovation for Energy (IN–RIE), Khon Kaen University, Khon Kaen 40002, Thailand
*
Author to whom correspondence should be addressed.
Submission received: 7 September 2021 / Revised: 26 September 2021 / Accepted: 30 September 2021 / Published: 5 October 2021

Abstract

:
In this work, the colossal dielectric properties and Maxwell—Wagner relaxation of TiO2–rich Na1/2Y1/2Cu3Ti4+xO12 (x = 0–0.2) ceramics prepared by a solid-state reaction method are investigated. A single phase of Na1/2Y1/2Cu3Ti4O12 is achieved without the detection of any impurity phase. The highly dense microstructure is obtained, and the mean grain size is significantly reduced by a factor of 10 by increasing Ti molar ratio, resulting in an increased grain boundary density and hence grain boundary resistance (Rgb). The colossal permittivities of ε′ ~ 0.7–1.4 × 104 with slightly dependent on frequency in the frequency range of 102–106 Hz are obtained in the TiO2–rich Na1/2Y1/2Cu3Ti4+xO12 ceramics, while the dielectric loss tangent is reduced to tanδ ~ 0.016–0.020 at 1 kHz due to the increased Rgb. The semiconducting grain resistance (Rg) of the Na1/2Y1/2Cu3Ti4+xO12 ceramics increases with increasing x, corresponding to the decrease in Cu+/Cu2+ ratio. The nonlinear electrical properties of the TiO2–rich Na1/2Y1/2Cu3Ti4+xO12 ceramics can also be improved. The colossal dielectric and nonlinear electrical properties of the TiO2–rich Na1/2Y1/2Cu3Ti4+xO12 ceramics are explained by the Maxwell–Wagner relaxation model based on the formation of the Schottky barrier at the grain boundary.

1. Introduction

Improvement of electronic device efficiency through the development of materials with enhanced electrical properties is significant. Colossal dielectric oxides (CDOs) with very high dielectric constant (ε′) are widely used to manufacture critical components in electronic devices, especially for multilayer ceramic capacitors (MLCCs) [1,2,3]. The ε′ of a CDO influences the geometry and performance of the MLCCs. The size of an MLCC can be miniaturized by using the insulating oxide between the metallic electrodes with a dielectric oxide with a higher ε′ value than the conventional oxide. Many CDOs have intensively been investigated, especially for CaCu3Ti4O12 (CCTO) and related compounds, expecting to replace traditional CDOs such as BaTiO3—based ceramics.
The dielectric and electrical properties of the perovskite CCTO ceramics have been extensively studied over the past two decades [4,5,6,7,8,9,10,11,12]. This is because CCTO ceramics showed high ε′ ~ 103–105 over wide ranges of temperature and frequency. Moreover, the ε′ of CCTO ceramics are relatively stable in the temperature range of 100–400 K compared to conventional BaTiO3–based ceramics used. Unfortunately, CCTO still presents too high dielectric loss (tanδ >> 0.05), which is not required for application in MLCCs [13,14,15]. Therefore, researchers have studied reducing the tanδ of CCTO ceramics by tuning the ceramic microstructure of CCTO and related ceramics according to their heterogeneous electrical structure. The special microstructure, which consists of semiconducting grains and highly resistive boundaries (GBs), can be produced using one–step processing method [5]. This heterogeneous microstructure is called to be an internal barrier layer capacitor (IBLC) structure. Accordingly, the resistivity and correlated tanδ can be improved by engineering the grains and GBs [16,17]. In addition, the presence of insulating GBs affects a nonlinear relationship between current density (J) and electrical field strength (E), which is a behavior required for developing varistor devices [4,9,18].
To improve the colossal dielectric and nonlinear JE properties, many effective methods have been proposed and studied, such as doping with suitable ions [19,20,21,22], tuning ceramic microstructure [9] and fabricated CCTO–matrix composites [10,18,23,24]. These methods have the same approach, which is to increase the total resistance of the insulating GBs (Rgb) for reducing tanδ. One of the most effective methods is to fabricate the CCTO–matrix composites using an appropriate ceramic filler such as CaTiO3 (CTO), Al2O3 or TiO2 [6,7,10,18,25]. For the CTO/CCTO/and TiO2/CCTO composites, although the tanδ can be significantly reduced to <0.05, their ε′ values are usually significantly decreased in the order of 103. For these two composite systems, the nonlinear electrical properties can also be significantly improved. The TiO2/CCTO composites can be easily prepared by designing TiO2–rich phase in the CCTO ceramics using the formula CaCu3Ti4+xO12+2x. Notably, the mean grain size was reduced, resulting in a significant increase in Rgb. This is the primary cause of the observed improvement of the colossal dielectric and nonlinear electrical properties of CCTO ceramics.
In addition to CCTO ceramics, the colossal dielectric properties of ACu3Ti4O12 oxides (A = Na1/2Bi1/2 [26], Na1/2Y1/2 [27,28,29,30], Na1/2La1/2 [31], Bi2/3 [32,33], Y2/3 [32,34], Cd [16,17,35], La2/3 [32], Sm2/3 [36], Na1/3Ca1/3Bi1/3 [37], Na1/3Cd1/3Y1/3 [38], Na1/3Sr1/3Y1/3 [39] and Na1/2Sm1/2 [40]) are very attractive, especially for the Na1/2Y1/2Cu3Ti4O12 (NYCTO) ceramics [27,28,29,30]. The NYCTO ceramics exhibited a high ε′ ~ 104 with low tanδ < 0.05 at 1 kHz compared to that of the CCTO ceramics [28,29,30]. Recently, the preparation, colossal dielectric permittivity and nonlinear electrical properties of NYCTO ceramics has been widely reported [27,28,29,30,39,41,42,43].
In this work, the TiO2—rich NYCTO ceramics were prepared by a conventional mixed—oxide method and investigated the dielectric properties. The crystal structure, phase composition and microstructural evolution of the sintered ceramics, as well as their oxidation states, were characterized. The primary cause of the enhanced colossal dielectric response was systematically elucidated. This study contributes an exciting concept for improving the colossal dielectric properties of the NYCTO ceramics by reducing their tanδ. We believe that this research work provides an effective route to improve the CDOs for future applications in MLCCs.

2. Experimental Details

TiO2—rich Na1/2Y1/2Cu3Ti4+xO12 (NYCTO+xTiO2) ceramics (where x = 0.0, 0.1 and 0.2) were prepared using solid-state reaction method (SSR). The starting materials were Na2CO3 (99.9%), Y2O3 (99.99%), CuO (99.9%) and TiO2 (99.9%), which were purchased from Sigma–Aldrich (St. Louis, MO, USA). Details for the preparation of NYCTO—based oxides were provided elsewhere [28,30,39]. The mixed powders for all compositions were calcined at 1000 °C for 10 h. Mixed powders (without calcination) were pressed into 9.5-mm-diameter pellets by uniaxial compression at ~100 MPa. Finally, the pellets were sintered at 1070 °C for 10 h in air. The sintered NYCTO + xTiO2 ceramics with x = 0.0, 0.1 and 0.2 were referred to as the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2, respectively.
The crystal structures of the sintered sample were characterized using X–ray diffraction (XRD, PANalytical, EMPYREAN, Shanghai, China), scanning electron microscopy (MiniSEM, SEC, SNE–4500 M), field emission scanning electron microscopy (FIB–FESEM,) with energy dispersive X–ray (EDX) spectroscopy and X–ray photoemission spectroscope (XPS, PHI5000 Versarobe II, ULVAC–PHI, Chigasaki, Japan). Comprehensive details were provided in our previous works [12,21,44,45,46].
For the nonlinear electrical and dielectric measurements, the surfaces of samples were polished. Next, the parallel and smooth surfaces were coated with silver paints and fired in the air at 600 °C for 30 min. The impedance and dielectric parameters of all sintered ceramics were measured with an impedance analyzer (KEYSIGHT E4990A, Santa Rosa, CA, USA). The dielectric properties were measured in the temperature range of −160 to 210 °C and the frequency range from 102–107 Hz. The nonlinear relationship between current density (J) and electrical field strength (E) was analyzed by using a high–voltage measurement unit (Keithley 247 model, Cleveland, OH, USA).

3. Results and Discussion

The XRD patterns of the sintered NYCTO + xTiO2 (x = 0, 0.1 and 0.2) ceramics are illustrated in Figure 1a, showing the single phase of NYCTO in all ceramics with a perovskite-structure (JCPDS 75–2188). The crystal structure of NYCTO is demonstrated in Figure 1b. The XRD peak corresponding to TiO2 phase cannot be detected in the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics, which may be due to a small amount of an excess TiO2 molar ratio that was lower than the resolution limit of the XRD technique. Accordingly, the lattice parameters (a) can be calculated and found to be 7.383, 7.383 and 7.384 Å for the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics, respectively. The a values are comparable to those reported in the literature [27,28,29,30]. The excess TiO2 composition in NYCTO ceramics does not affect the lattice parameter. This result indicates that the TiO2—rich phase is segregated from the primary phase of NYCTO, which may exist as the TiO2—rich boundary. The XRD result shows that the NYCTO + xTiO2 has successfully been fabricated using the SSR method.
Even though the excessive TiO2 phase was not detected in the XRD patterns for all ceramics, the variation in compositions of the CuO and/or TiO2 ratios in an ACu3Ti4O12 compound usually affects the dielectric and electrical properties [6,10,24,45,47,48,49]. Thus, we first investigated the dielectric properties of the NYCTO + xTiO2 ceramics at around room temperature (30 °C). The relationship between the ε′ and frequency of the NYCTO + xTiO2 ceramics is shown in Figure 2a. The ε′ value of the NYCTO ceramic is huge (2.07 × 104 at 103 Hz) with a quite low tanδ ~ 0.115, which is similar to that reported in the previous works [27,28,30]. However, the ε′ of the NYCTO ceramic is largely dependent on the frequency in a low–frequency range, which is usually owing to the dominant effect of non–Ohmic sample–electrode interface [8,15,44,50]. At 106 Hz, the ε′ begins to decline due to the primary dielectric relaxation mechanism [12,19]. Interestingly, the ε′ values of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics are more stable with frequency than that of the NYCTO ceramic. The TiO2—rich phase can improve the frequency dependence of the ε′ of the NYCTO + xTiO2 ceramics. The ε′ values of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics at 30 °C and 103 Hz are 1.39 × 104 and 7.15 × 103, respectively. Even though the dielectric response in the NYCTO + 0.1TiO2 ceramic was decreased due to the excessive TiO2, its ε′ value was still larger than 104 over the measured frequency range. The decrease in the ε′ value of the NYCTO + xTiO2 ceramics is similar to that observed in the CaCu3Ti4+xO12 [6].
Figure 2b illustrates the tanδ at 30 °C for the NYCTO + xTiO2 ceramics over the frequency range of 102–106 Hz. In the frequency range of 102–105 Hz, the tanδ values of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics are much lower than the NYCTO ceramic. Furthermore, in this frequency range, the tanδ values of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics are lower than 0.1. The rapid increase in tanδ of all the ceramics is attributed to the primary dielectric relaxation, i.e., Maxwell—Wagner polarization relaxation [12,19]. The tanδ values of the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics 103 Hz are 0.115, 0.020 and 0.016, respectively. Notably, the tanδ value of the NYCTO ceramics can be significantly reduced by increasing the excessive TiO2 molar ratio. This result indicates the influential role of TiO2-rich on the significantly improved dielectric properties of the NYCTO ceramics.
The temperature dependence of the dielectric properties of the NYCTO + xTiO2 ceramics is illustrated in Figure 3a,b. In the temperature range from −125 to 110 °C, the ε′ values of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics are more stable with temperature than that of the NYCTO ceramic. At the temperature below −120 °C, the ε′ rapidly decreased, corresponding to the rapid increase in tanδ. This is the Maxwell—Wagner polarization relaxation in the NYCTO + xTiO2 ceramics. Considering in the high-temperature range, the ε′ and tanδ significantly increase, which is associated with the DC conduction of free charge carriers associated with oxygen vacancies [11,33,51,52].
It is widely accepted that the colossal dielectric properties of the ACu3Ti4O12 oxide groups is the result from the electrical heterogeneous in the microstructure [5,11,15,30,33,48]. The electrically heterogeneous microstructure can be confirmed using impedance spectroscopy [5,53]. Furthermore, the heterogeneous electrical microstructure can also be used to explain the nonlinear electrical properties of the ACu3Ti4O12 ceramics [21,23,24]. Generally, the semicircular arc due to the electrical response of the semiconducting grains of the ACu3Ti4O12 compounds can be observed in the impedance complex plane (Z*) plots at low temperatures [5,54]. Thus, to confirm the formation of semiconducting grains, Z* plots of all the NYCTO + xTiO2 ceramics are demonstrated at −100 °C, as shown in Figure 3c. The semicircular arcs of all the ceramics can be observed, while only parts of relatively large semicircular arcs can be observed. These two parts can be assigned as the electrical responses of the semiconducting grain and insulating GB, respectively [5,54]. The grain resistance (Rg), which can be calculated from the diameter of the semicircular arc of the NYCTO + xTiO2 ceramics, was increased by increasing the TiO2 molar ratio. In general, we expect that the TiO2—rich should not affect the electrical properties of the semiconducting grains but should only affect the insulating boundaries due to the segregation of the TiO2—rich phase. In this current study, the excessive TiO2 molar ratio can affect the electrical properties inside the semiconducting grains, which will be discussed in the last section. Nevertheless, according to the impedance spectroscopy, the variation in the colossal dielectric properties and dielectric behavior of the NYCTO + xTiO2 ceramics should be described in all aspects based on the IBLC model.
Figure 4a displays the Z* plots of all the ceramics at 30 °C. Only parts of the relatively large arcs can be observed with a nonzero intercept (inset of Figure 4a), which is similar to that reported in the previous works [5,6,12,19,21,44,54]. The Rg value at 30 °C, which can be calculated from the nonzero intercept, increases with increasing the excessive TiO2 molar ratio. Even though an entire arc cannot be observed, the Rgb values of all the NYCTO + xTiO2 ceramics can be estimated. As shown in Figure 4b, the tanδ of the NYCTO + xTiO2 ceramics is inversely proportional to the Rgb value. According to the IBLC structure [20,44], the low—frequency tanδ value is correlated to the total resistance, which is governed by the Rgb value. The low—frequency tanδ can be reduced by increasing Rgb. Thus, the correlation of the tanδ and Rgb values follows the IBLC model.
In addition to the variation in tanδ, the IBLC model should be used to reasonably describe the overall dielectric properties of the NYCTO + xTiO2 ceramics. Therefore, the microstructure of the sintered ceramics was studied. Figure 5 shows the SEM images of the polished surface of the NYCTO + xTiO2 ceramics and their grain size distributions. All the ceramics reveal the grain and GB structure with a highly dense microstructure. The mean grain size of the NYCTO + xTiO2 ceramics was extremely reduced by increasing x from 0 to 0.2. This result is similar to that reported in the previous reports for TiO2—rich CCTO ceramics [6]. The mean grain sizes of the NYCTO + xTiO2 ceramics with x = 0, 0.1 and 0.2 are 32.80 ± 20.44, 4.05 ± 2.08 and 3.18 ± 1.35 μm, respectively. The excessive TiO2 could intercept the grain growth rate of the NYCTO ceramics due to the pinning effect of excessive TiO2—rich phase particles during the sintering process [55]. We also found that the segregation of the Cu—rich phase is slightly observed along the grain boundaries, as remarked in the square area of Figure 5b.
The observed decrease in ε′ (Figure 2a) value of the NYCTO + xTiO2 ceramics should be caused by the decrease in the mean grain size, following a simple series layer model of the IBLC structure [20,44,56],
ε = ε g b G t g b
where G is the mean grain size, tgb is the thickness of the GB and εgb is the dielectric constant of the GBs. Furthermore, it is also suggested that, but does not clearly prove, the decrease in the ε′ might be due to the increase in tgb due to the TiO2–rich phase. The EDS and EDS–SEM mapping techniques were used to further characterize the microstructure and elemental distribution. As revealed in Figure 6a and its inset, all elements comprising the NYCTO + 0.1TiO2 ceramic are observed in the EDS spectra, confirming the existence of Na, Y, Cu, Ti and O elements. As demonstrated in Figure 6b–g, the Na, Y, Cu and O elements disperse well throughout the microstructure. It is observed that the relatively higher brightness of Ti mapping element along the GBs compared to that of the grains can be observed, confirming the segregation of TiO2—rich boundary. Thus, this is one of the most important factors contributing to the decrease in the ε′ values of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics. According to the microstructure analyses, the density of the insulating GB layers in the NYCTO + xTiO2 ceramics was significantly increased by increasing TiO2—rich phase owing to the decreased mean grain size [9,20]. Obviously, The significantly increased Rgb (Figure 4) is also attributed to the enhancement of insulating GB density. Furthermore, the increased Rgb is also due to the increase in tgb. The XEX spectra detected at the grain and GB areas are shown in Figure 7. It was found that the percentage ratios of Ti(wt%)/Cu(wt%) at the grain and GB were found to be 0.808 and 0.834, respectively.
According to our previous work [28,29], it was found that the NYCTO ceramics exhibited the nonlinear JE characteristics with nonlinear coefficients (α) of 5.7–6.6. Furthermore, it was reported that the nonlinear properties of CCTO ceramics could be enhanced by increasing the excessive TiO2 molar ratio [6,7]. The α value of the CaCu3Ti4+xO12+2x ceramic with x = 0.15 was increased to 7.9. As illustrated in Figure 8, all the NYCTO + xTiO2 ceramics exhibit the JE characteristics. The α and electric breakdown (Eb) of the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics are 4.96 and 724.50 V/cm, 4.71 and 910.59 V/cm, and 13.36 and 7948.04 V/cm, respectively. The Eb increased significantly with increasing TiO2—rich phase, similar to that observed in the CaCu3Ti4+xO12+2x ceramics [6]. Nevertheless, the enhanced α value of 13.36 is larger than that of the CaCu3Ti4+xO12+2x ceramics. The segregation of TiO2—rich boundary and the increased GB density are the key factors, giving rise to the significantly improved nonlinear JE properties. The α value is often related to Eb value [4].
The nonlinear electrical behavior of ACu3Ti4O12 oxides is widely believed to be originated by the formation of the Schottky barrier at the GBs [4,5,7,18]. The increased Eb value is associated with the increase in Rgb due to the significantly increased GB density and GB thickness, which are classified as the geometric factors of the GBs [9]. Furthermore, the intrinsic factor of the GBs, i.e., the Schottky barrier height (Φb), can usually have a remarkable effect on the Rgb and Eb values [4,5,7,9]. The Φb is closely related to the conduction activation energy at the GBs (Egb) [7,13,14]. To calculate the Egb value, Rgb values at different temperatures were calculated. Figure 9a and its inset show Z* plots and nonzero intercept of the NYCTO ceramic at various temperatures. The Rg and Rgb values can be obtained and found to decrease with increasing temperature. Thus, Egb can be calculated by using the Arrhenius law [5,9]:
R g b = R 0 e x p ( E g b k B T ) ,
where R0, kB and T are the per-exponential constant term, Boltzmann constant and absolute temperature, respectively. Figure 9b depicts the relationship between Rgb and 1000/T of the NYCTO + xTiO2 ceramics. The Egb values can be calculated from the slopes of the Rgb and 1000/T plots, which are linearly fitted by using the Arrhenius law. The Egb values of the NYCTO + xTiO2 ceramics are 0.547, 0.583 and 0.714 eV, respectively. Therefore, Φb of the NYCTO + xTiO2 ceramics can be increased by increasing the TiO2—rich phase. The improved nonlinear JE properties and electrical properties of the GBs are also caused by the enhanced Φb, just as observed in the TiO2—rich CCTO ceramics [7]. The increased Φb values of the NYCTO + xTiO2 ceramics are likely attributed to the suppressed oxygen vacancies and/or oxygen enrichment at the GBs due to the segregation of TiO2—rich boundary [7,18].
In addition to the electrical properties of the GBs, the electrical properties of the semiconducting grains must also be characterized. The conduction activation energy of the grains (Eg) can be calculated from the temperature dependence of Rg. The Rg values at a low–temperature range can be easily calculated using the admittance spectroscopy (Y*) [28,29,57], as the following equation:
Y * = ( R g b 1 ) ( 1 ω 2 τ g τ g b + i ω τ g b ) 1 + i ω τ ,
where τgb = RgbCgb, τg = RgCg and τ = RgCgb and Cg and Cgb are the grain and GB capacitance values, respectively. According to the impedance spectroscopy, it was found that Rgb >> Rg and Cgb >> Cg. From Equation (3), Rg can be obtained from the relation R g = 1 / 2 Y m a x , where Y m a x is the maximum value at Y peak. As shown in Figure 10a–c, Y m a x appears in the temperature range from −60 to 0 °C. Consequently, Eg can be calculated by using the Arrhenius law, σ g = σ 0 e x p ( E g / k b T ) , where σg is the grain conductivity ( σ g 1 / R g ), and σ0 is a per–exponential constant term. The Eg can be calculated by the linear fitting data, as demonstrated in the insets of Figure 10a–c. The Eg values are 0.112, 0.118 and 0.126 eV for the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics, respectively. The Eg slightly increased with increasing the excessive TiO2, corresponding to the increase in Rg (Figure 3c). The difference between Eg and Egb clearly indicates the formation of IBLC microstructure, consisting of the semiconducting grains and insulating GBs.
The XPS technique was further used to characterize the electrical properties of the grains. Figure 11a–c displays the XPS spectra of the Cu 2p3/2 for the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics, respectively. According to the fitted curves, the XPS peak of the Cu 2p3/2 can be divided into two peaks at relatively low and high binding energies, corresponding to the Cu+ and Cu2+, respectively. Note that only Ti4+ can be detected in the XPS spectra, as shown in Figure 11d–f. Thus, the conduction in the semiconducting grains of all the ceramics is attributed to the electron hopping between Cu+ ↔ Cu2+. The Cu+/Cu2+ ratios of the NYCTO, NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics are 0.069 ± 0.027, 0.053 ± 0.021 and 0.049 ± 0.020, respectively. The Cu+/Cu2+ ratios decreased with increasing TiO2—rich phase. Generally, CCTO and related ACu3Ti4O12 ceramics lose small amount of oxygen during sintering, giving rise to oxygen vacancies and associated free electrons [5,9]. Accordingly, a small amount of Cu+ and/or Ti4+ can be detected. For the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics, the diffusion of oxygen vacancies during the sintering process may be inhibited by the segregation of TiO2—rich phase along the GBs. Thus, the oxygen loss and related oxygen vacancies in the TiO2—rich NYCTO ceramics were reduced, leading to the decrease in Cu+ ions. The increased Rg (Figure 3c) can be proved to be caused by the decreased Cu+/Cu2+ ratios. In addition to the grain size effect, the observed decrease in the ε′ of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics can be described based on the IBLC microstructure. Under al applied electric field, charge carriers inside the semiconducting grains are moved to trap at the insulating GB due to a high potential barrier height, inducing the interfacial polarization and hence high ε′ value. The intensity of the interfacial polarization of the NYCTO + 0.1TiO2 and NYCTO + 0.2TiO2 ceramics should be lower than that of the NYCTO ceramic because of a lower concentration of free carriers inside the grains, which is considered by a larger Rg value.

4. Conclusions

In conclusion, we have successfully synthesized the TiO2—rich NYCTO ceramics prepared using the SSR method. The effects of Ti—excess on the microstructure, colossal dielectric properties and nonlinear JE characteristics were studied. Only the main phase of the NYCTO structure was detected in the XRD patterns, which might be due to the presence of an amorphous phase of TiO2 along the GBs. Significantly reduced grain with highly dense microstructure was observed in the TiO2—rich NYCTO ceramics, which was due to the pinning effect of the TiO2—rich phase particles. The reduced grain sizes, which can cause an increase in the GB density, resulted in the significant enhancement of Rgb, and hence reduced tanδ. The colossal ε′ values of ~ 0.7–1.4 × 104 was achieved in the TiO2—rich NYCTO ceramics. The TiO2—rich NYCTO ceramics also showed the enhanced nonlinear JE properties due to the improved GB properties. The Rg value was also increased owing to the decreased Cu+/Cu2+ ratio, confirming by the XPS result. The overall colossal dielectric permittivity and nonlinear electrical properties were well described using the Maxwell–Wagner polarization relaxation model based on the formation of the Schottky barrier at the grain boundary.

Author Contributions

Conceptualization, P.T.; methodology, P.S. (Pariwat Saengvong); formal analysis, P.S.(Pariwat Saengvong) and P.T.; investigation, P.S. (Pariwat Saengvong), N.C., B.P., A.P., V.H., S.K., P.S. (Pornjuk Srepusharawoot) and P.T.; writing—original draft preparation, P.S. (Pariwat Saengvong) and P.T.; writing—review and editing, P.T.; visualization, P.S. (Pariwat Saengvong) and P.T. All authors have read and agreed to the published version of the manuscript.

Funding

This project is funded by National Research Council of Thailand (NRCT) (Grant No. N41A640084). This research was also funded by the Research and Graduate Studies, Khon Kaen University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Acknowledgments

This project is funded by National Research Council of Thailand (NRCT): (N41A640084). This research was also funded by the Research and Graduate Studies, Khon Kaen University. P. Saengvong expresses his gratitude to the Science Achievement Scholarship of Thailand (SAST) for a Ph.D. scholarship in Physics.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Wang, Y.; Jie, W.; Yang, C.; Wei, X.; Hao, J. Colossal Permittivity Materials as Superior Dielectrics for Diverse Applications. Adv. Funct. Mater. 2019, 29, 1808118. [Google Scholar] [CrossRef]
  2. Zhu, J.; Wu, D.; Liang, P.; Zhou, X.; Peng, Z.; Chao, X.; Yang, Z. Ag+/W6+ co-doped TiO2 ceramic with colossal permittivity and low loss. J. Alloys. Compd. 2021, 856, 157350. [Google Scholar] [CrossRef]
  3. Zhou, X.; Liang, P.; Zhu, J.; Peng, Z.; Chao, X.; Yang, Z. Enhanced dielectric performance of (Ag1/4Nb3/4)0.01Ti0.99O2 ceramic prepared by a wet-chemistry method. Ceram. Int. 2020, 46, 11921–11925. [Google Scholar] [CrossRef]
  4. Chung, S.-Y.; Kim, I.-D.; Kang, S.-J.L. Strong nonlinear current–voltage behaviour in perovskite-derivative calcium copper titanate. Nat. Mater. 2004, 3, 774–778. [Google Scholar] [CrossRef] [PubMed]
  5. Adams, T.; Sinclair, D.; West, A. Characterization of grain boundary impedances in fine-and coarse-grained CaCu3Ti4O12 ceramics. Phys. Rev. B 2006, 73, 094124. [Google Scholar] [CrossRef]
  6. Lin, Y.-H.; Cai, J.; Li, M.; Nan, C.-W.; He, J. High dielectric and nonlinear electrical behaviors in TiO2-rich CaCu3Ti4O12 ceramics. Appl. Phys. Lett. 2006, 88, 172902. [Google Scholar] [CrossRef]
  7. Lin, Y.-H.; Cai, J.; Li, M.; Nan, C.-W.; He, J. Grain boundary behavior in varistor-capacitor TiO2-rich CaCu3Ti4O12 ceramics. J. Appl. Phys. 2008, 103, 074111. [Google Scholar] [CrossRef]
  8. Li, M.; Shen, Z.; Nygren, M.; Feteira, A.; Sinclair, D.C.; West, A.R. Origin(s) of the apparent high permittivity in CaCu3Ti4O12 ceramics: Clarification on the contributions from internal barrier layer capacitor and sample-electrode contact effects. J. Appl. Phys. 2009, 106, 104106. [Google Scholar] [CrossRef]
  9. Thongbai, P.; Yamwong, T.; Maensiri, S.; Amornkitbamrung, V.; Chindaprasirt, P. Improved Dielectric and Nonlinear Electrical Properties of Fine-Grained CaCu3Ti4O12 Ceramics Prepared by a Glycine-Nitrate Process. J. Am. Ceram. Soc. 2014, 97, 1785–1790. [Google Scholar] [CrossRef]
  10. Maleki Shahraki, M.; Daeijavad, H.; Emami, A.H.; Abdollahi, M.; Karimi, A. An engineering design based on nano/micro-sized composite for CaTiO3/CaCu3Ti4O12 materials and its dielectric and non-Ohmic properties. Ceram. Int. 2019, 45, 21676–21683. [Google Scholar] [CrossRef]
  11. Kozlinskei, L.L.; de Andrade Paes, A.T.; Grzebielucka, E.C.; Borges, C.P.F.; de Andrade, A.V.C.; de Souza, E.C.F.; Antunes, S.R.M. Processing influence in the CaCu3Ti4O12 electrical properties. Appl. Phys. A 2020, 126, 447. [Google Scholar] [CrossRef]
  12. Jumpatam, J.; Putasaeng, B.; Chanlek, N.; Manyam, J.; Srepusharawoot, P.; Krongsuk, S.; Thongbai, P. Influence of Sn and F dopants on giant dielectric response and Schottky potential barrier at grain boundaries of CCTO ceramics. Ceram. Int. 2021, 47, 27908–27915. [Google Scholar] [CrossRef]
  13. 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]
  14. Jumpatam, J.; Putasaeng, B.; Yamwong, T.; Thongbai, P.; Maensiri, S. Enhancement of giant dielectric response in Ga-doped CaCu3Ti4O12 ceramics. Ceram. Int. 2013, 39, 1057–1064. [Google Scholar] [CrossRef]
  15. Lin, H.; Xu, W.; Zhang, H.; Chen, C.; Zhou, Y.; Yi, Z. Origin of high dielectric performance in fine grain-sized CaCu3Ti4O12 materials. J. Eur. Ceram. Soc. 2020, 40, 1957–1966. [Google Scholar] [CrossRef]
  16. Peng, Z.; Wang, J.; Zhou, X.; Zhu, J.; Lei, X.; Liang, P.; Chao, X.; Yang, Z. Grain engineering inducing high energy storage in CdCu3Ti4O12 ceramics. Ceram. Int. 2020, 46, 14425–14430. [Google Scholar] [CrossRef]
  17. Peng, Z.; Wu, D.; Liang, P.; Zhou, X.; Wang, J.; Zhu, J.; Chao, X.; Yang, Z. Grain boundary engineering that induces ultrahigh permittivity and decreased dielectric loss in CdCu3Ti4O12 ceramics. J. Am. Ceram. Soc. 2020, 103, 1230–1240. [Google Scholar] [CrossRef]
  18. Ramírez, M.A.; Bueno, P.R.; Varela, J.A.; Longo, E. Non-Ohmic and dielectric properties of a Ca2Cu2Ti4O12 polycrystalline system. Appl. Phys. Lett. 2006, 89, 212102. [Google Scholar] [CrossRef]
  19. Boonlakhorn, J.; Manyam, J.; Srepusharawoot, P.; Krongsuk, S.; Thongbai, P. Effects of Charge Compensation on Colossal Permittivity and Electrical Properties of Grain Boundary of CaCu3Ti4O12 Ceramics Substituted by Al3+ and Ta5+/Nb5+. Molecules 2021, 26, 3294. [Google Scholar] [CrossRef]
  20. Boonlakhorn, J.; Chanlek, N.; Manyam, J.; Srepusharawoot, P.; Thongbai, P. Simultaneous two-step enhanced permittivity and reduced loss tangent in Mg/Ge-Doped CaCu3Ti4O12 ceramics. J. Alloys. Compd. 2021, 877, 160322. [Google Scholar] [CrossRef]
  21. Jumpatam, J.; Putasaeng, B.; Chanlek, N.; Boonlakhorn, J.; Thongbai, P.; Phromviyo, N.; Chindaprasirt, P. Significantly improving the giant dielectric properties of CaCu3Ti4O12 ceramics by co-doping with Sr2+ and F- ions. Mater. Res. Bull. 2021, 111043. [Google Scholar] [CrossRef]
  22. Boonlakhorn, J.; Putasaeng, B.; Kidkhunthod, P.; Manyam, J.; Krongsuk, S.; Srepusharawoot, P.; Thongbai, P. First-principles calculations and experimental study of enhanced nonlinear and dielectric properties of Sn4+-doped CaCu2.95Mg0.05Ti4O12 ceramics. J. Eur. Ceram. Soc. 2021, 41, 5176–5183. [Google Scholar] [CrossRef]
  23. Cortés, J.A.; Moreno, H.; Orrego, S.; Bezzon, V.D.N.; Ramírez, M.A. Dielectric and non-ohmic analysis of Sr2+ influences on CaCu3Ti4O12-based ceramic composites. Mater. Res. Bull. 2021, 134, 111071. [Google Scholar] [CrossRef]
  24. Cotrim, G.; Cortés, J.A.; Moreno, H.; Freitas, S.M.; Rezende, M.V.S.; Hein, L.R.O.; Ramírez, M.A. Tunable capacitor-varistor response of CaCu3Ti4O12/CaTiO3 ceramic composites with SnO2 addition. Mater. Charact. 2020, 170, 110699. [Google Scholar] [CrossRef]
  25. Jumpatam, J.; Chanlek, N.; Thongbai, P. Giant dielectric response, electrical properties and nonlinear current-voltage characteristic of Al2O3-CaCu3Ti4O12 nanocomposites. Appl. Surf. Sci. 2019, 476, 623–631. [Google Scholar] [CrossRef]
  26. Tuichai, W.; Danwittayakul, S.; Yamwong, T.; Thongbai, P. Synthesis, dielectric properties, and influences oxygen vacancies have on electrical properties of Na1/2Bi1/2Cu3Ti4O12 ceramics prepared by a urea combustion method. J. Sol.-Gel Sci. Technol. 2015, 76, 630–636. [Google Scholar] [CrossRef]
  27. Somphan, W.; Sangwong, N.; Yamwong, T.; Thongbai, P. Giant dielectric and electrical properties of sodium yttrium copper titanate: Na1/2Y1/2Cu3Ti4O12. J. Mater. Sci. Mater. Electron. 2012, 23, 1229–1234. [Google Scholar] [CrossRef]
  28. Jumpatam, J.; Mooltang, A.; Putasaeng, B.; Kidkhunthod, P.; Chanlek, N.; Thongbai, P.; Maensiri, S. Effects of Mg2+ doping ions on giant dielectric properties and electrical responses of Na1/2Y1/2Cu3Ti4O12 ceramics. Ceram. Int. 2016, 42, 16287–16295. [Google Scholar] [CrossRef]
  29. Jumpatam, J.; Somphan, W.; Boonlakhorn, J.; Putasaeng, B.; Kidkhunthod, P.; Thongbai, P.; Maensiri, S. Non-Ohmic Properties and Electrical Responses of Grains and Grain Boundaries of Na1/2Y1/2Cu3Ti4O12 Ceramics. J. Am. Ceram. Soc. 2017, 100, 157–166. [Google Scholar] [CrossRef]
  30. Liang, P.; Li, Y.; Zhao, Y.; Wei, L.; Yang, Z. Origin of giant permittivity and high-temperature dielectric anomaly behavior in Na0.5Y0.5Cu3Ti4O12 ceramics. J. Appl. Phys. 2013, 113, 224102. [Google Scholar] [CrossRef]
  31. Liu, Z.; Yang, Z. Structure and electric properties of NaxLa(2−x)/3Cu3Ti4O12 ceramics prepared by sol–gel method. J. Mater. Sci. Mater. Electron. 2018, 29, 9326–9338. [Google Scholar] [CrossRef]
  32. Liu, J.; Duan, C.-g.; Mei, W.N.; Smith, R.W.; Hardy, J.R. Dielectric properties and Maxwell-Wagner relaxation of compounds ACu3Ti4O12 (A=Ca,Bi2∕3,Y2∕3,La2∕3). J. Appl. Phys. 2005, 98, 093703. [Google Scholar] [CrossRef]
  33. Tang, Z.; Wu, K.; Li, J.; Huang, S. Optimized dual-function varistor-capacitor ceramics of core-shell structured xBi2/3Cu3Ti4O12/(1−x)CaCu3Ti4O12 composites. J. Eur. Ceram. Soc. 2020, 40, 3437–3444. [Google Scholar] [CrossRef]
  34. Wang, X.; Liang, P.; Peng, Z.; Peng, H.; Xiang, Y.; Chao, X.; Yang, Z. Significantly enhanced breakdown electric field in Zn-doped Y2/3Cu3Ti4O12 ceramics. J. Alloys. Compd. 2019, 778, 391–397. [Google Scholar] [CrossRef]
  35. Peng, Z.; Wang, J.; Lei, X.; Zhu, J.; Xu, S.; Liang, P.; Wei, L.; Wu, D.; Wang, J.; Chao, X.; et al. Colossal dielectric response in CdAlxCu3-xTi4O12 perovskite ceramics. Mater. Chem. Phys. 2021, 258, 123940. [Google Scholar] [CrossRef]
  36. Thomas, A.K.; George, M.; Abraham, K.; Sajan, D. Giant dielectric constant, dielectric relaxations, and tunable properties of Sm2/3Cu3Ti4O12 ceramics. Int. J. Appl. Ceram. Technol. 2021, 18, 499–510. [Google Scholar] [CrossRef]
  37. Kum-onsa, P.; Thongbai, P.; Putasaeng, B.; Yamwong, T.; Maensiri, S. Na1/3Ca1/3Bi1/3Cu3Ti4O12: A new giant dielectric perovskite ceramic in ACu3Ti4O12 compounds. J. Eur. Ceram. Soc. 2015, 35, 1441–1447. [Google Scholar] [CrossRef]
  38. Peng, Z.; Zhou, X.; Wang, J.; Zhu, J.; Liang, P.; Chao, X.; Yang, Z. Origin of colossal permittivity and low dielectric loss in Na1/3Cd1/3Y1/3Cu3Ti4O12 ceramics. Ceram. Int. 2020, 46, 11154–11159. [Google Scholar] [CrossRef]
  39. Saengvong, P.; Boonlakhorn, J.; Chanlek, N.; Putasaeng, B.; Thongbai, P. Giant dielectric permittivity with low loss tangent and excellent non−Ohmic properties of the (Na+, Sr2+, Y3+)Cu3Ti4O12 ceramic system. Ceram. Int. 2020, 46, 9780–9785. [Google Scholar] [CrossRef]
  40. Somphan, W.; Thongbai, P.; Yamwong, T.; Maensiri, S. High Schottky barrier at grain boundaries observed in Na1/2Sm1/2Cu3Ti4O12 ceramics. Mater. Res. Bull. 2013, 48, 4087–4092. [Google Scholar] [CrossRef]
  41. Kum-onsa, P.; Phromviyo, N.; Thongbai, P. Suppressing loss tangent with significantly enhanced dielectric permittivity of poly (vinylidene fluoride) by filling with Au–Na1/2Y1/2Cu3Ti4O12 hybrid particles. RSC 2020, 10, 40442–40449. [Google Scholar] [CrossRef]
  42. Kum-onsa, P.; Thongbai, P. Improved Dielectric Properties of Poly (vinylidene fluoride) Composites Incorporating Na1/2Y1/2Cu3Ti4O12 Particles. Mater. Today Commun. 2020, 25, 101654. [Google Scholar] [CrossRef]
  43. Ahmad, M.; Mahfoz Kotb, H. Giant dielectric properties of fine-grained Na1/2Y1/2Cu3Ti4O12 ceramics prepared by mechanosynthesis and spark plasma sintering. J. Mater. Sci. Mater. Electron. 2015, 26, 8939–8948. [Google Scholar] [CrossRef]
  44. Boonlakhorn, J.; Chanlek, N.; Manyam, J.; Srepusharawoot, P.; Krongsuk, S.; Thongbai, P. Enhanced giant dielectric properties and improved nonlinear electrical response in acceptor-donor (Al3+, Ta5+)-substituted CaCu3Ti4O12 ceramics. J. Adv. Ceram. 2021, 10, 1–13. [Google Scholar]
  45. Jumpatam, J.; Putasaeng, B.; Chanlek, N.; Thongbai, P. Influences of Sr2+ Doping on Microstructure, Giant Dielectric Behavior, and Non-Ohmic Properties of CaCu3Ti4O12/CaTiO3 Ceramic Composites. Molecules 2021, 26, 1994. [Google Scholar] [CrossRef]
  46. Boonlakhorn, J.; Srepusharawoot, P.; Thongbai, P. Distinct roles between complex defect clusters and insulating grain boundary on dielectric loss behaviors of (In3+/Ta5+) co-doped CaCu3Ti4O12 ceramics. Results Phys. 2020, 16, 102886. [Google Scholar] [CrossRef]
  47. Zhao, J.; Chen, M.; Tan, Q. Embedding nanostructure and colossal permittivity of TiO2-covered CCTO perovskite materials by a hydrothermal route. J. Alloys. Compd. 2021, 885, 160948. [Google Scholar] [CrossRef]
  48. Zhao, J.; Zhao, H.; Zhu, Z. Influence of sintering conditions and CuO loss on dielectric properties of CaCu3Ti4O12 ceramics. Mater. Res. Bull. 2019, 113, 97–101. [Google Scholar] [CrossRef]
  49. Gelfuso, M.V.; Uribe, J.O.M.; Thomazini, D. Deficient or excessive CuO-TiO2 phase influence on dielectric properties of CaCu3Ti4O12 ceramics. Int. J. Appl. Ceram. Technol. 2019, 16, 868–882. [Google Scholar] [CrossRef]
  50. Mao, P.; Wang, J.; Xiao, P.; Zhang, L.; Kang, F.; Gong, H. Colossal dielectric response and relaxation behavior in novel system of Zr4+ and Nb5+ co-substituted CaCu3Ti4O12 ceramics. Ceram. Int. 2021, 47, 111–120. [Google Scholar] [CrossRef]
  51. Peng, Z.; Wang, J.; Liang, P.; Zhu, J.; Zhou, X.; Chao, X.; Yang, Z. A new perovskite-related ceramic with colossal permittivity and low dielectric loss. J. Eur. Ceram. Soc. 2020, 40, 4010–4015. [Google Scholar] [CrossRef]
  52. Peng, Z.; Liang, P.; Wang, J.; Zhou, X.; Zhu, J.; Chao, X.; Yang, Z. Interfacial effect inducing thermal stability and dielectric response in CdCu3Ti4O12 ceramics. Solid State Ion. 2020, 348, 115290. [Google Scholar] [CrossRef]
  53. Moulson, A.J.; Herbert, J.M. Electroceramics: Materials, Properties, Applications, 2nd ed.; Wiley: West Sussex, New York, 2003; p xiv; 557p. [Google Scholar]
  54. 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]
  55. Rahaman, M.N. Ceramic Processing and Sintering, 2nd ed.; M. Dekker: New York, NY, USA, 2003; p xiii; 875p. [Google Scholar]
  56. Tuichai, W.; Danwittayakul, S.; Chanlek, N.; Takesada, M.; Pengpad, A.; Srepusharawoot, P.; Thongbai, P. High-Performance Giant Dielectric Properties of Cr3+/Ta5+ Co-Doped TiO2 Ceramics. ACS Omega 2021, 6, 1901–1910. [Google Scholar] [CrossRef] [PubMed]
  57. Sangwong, N.; Somphan, W.; Thongbai, P.; Yamwong, T.; Meansiri, S. Electrical responses and dielectric relaxations in giant permittivity NaCu3Ti3TaO12 ceramics. Appl. Phys. A 2012, 108, 385–392. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of NYCTO + xTiO2 ceramics with x = 0.0, 0.1 and 0.2. (b) NYCTO structure.
Figure 1. (a) XRD patterns of NYCTO + xTiO2 ceramics with x = 0.0, 0.1 and 0.2. (b) NYCTO structure.
Molecules 26 06043 g001
Figure 2. Dielectric properties at room temperature as a function of frequency for NYCTO + xTiO2 ceramics with different doping content: (a) dielectric permittivity (ε′) and (b) loss tangent (tanδ).
Figure 2. Dielectric properties at room temperature as a function of frequency for NYCTO + xTiO2 ceramics with different doping content: (a) dielectric permittivity (ε′) and (b) loss tangent (tanδ).
Molecules 26 06043 g002
Figure 3. (a,b) Dielectric permittivity (ε′) and loss tangent (tanδ) of NYCTO + xTiO2 ceramics as a function of temperature (−150–200 °C). (c) Impedance complex plane plots (Z*) at −100 °C, showing the electrical response on semiconducting grains.
Figure 3. (a,b) Dielectric permittivity (ε′) and loss tangent (tanδ) of NYCTO + xTiO2 ceramics as a function of temperature (−150–200 °C). (c) Impedance complex plane plots (Z*) at −100 °C, showing the electrical response on semiconducting grains.
Molecules 26 06043 g003
Figure 4. (a) Impedance complex plane (Z*) plots of NYCTO + xTiO2 ceramics at 30 °C; inset shows the enlarged view near the origin, showing a nonzero intercept. (b) Relationship of loss tangent (tanδ) at 1 kHz and 30 °C and grain boundary resistance (Rgb) at 30 °C.
Figure 4. (a) Impedance complex plane (Z*) plots of NYCTO + xTiO2 ceramics at 30 °C; inset shows the enlarged view near the origin, showing a nonzero intercept. (b) Relationship of loss tangent (tanδ) at 1 kHz and 30 °C and grain boundary resistance (Rgb) at 30 °C.
Molecules 26 06043 g004
Figure 5. SEM images of NYCTO + xTiO2 ceramics with x = (a) 0.0, (b) 0.1 and (c) 0.2 and grain size distributions of NYCTO + xTiO2 ceramics with x = (d) 0.0, (e) 0.1 and (f) 0.2.
Figure 5. SEM images of NYCTO + xTiO2 ceramics with x = (a) 0.0, (b) 0.1 and (c) 0.2 and grain size distributions of NYCTO + xTiO2 ceramics with x = (d) 0.0, (e) 0.1 and (f) 0.2.
Molecules 26 06043 g005
Figure 6. (a) EDS spectrum of NYCTO + xTiO2 ceramic with x = 0.1; inset shows EDS testing area. (b) SEM image and corresponding SEM—EDS mapping images of (c) Na, (d) Y, (e) Cu, (f) Ti and (g) O for NYCTO + xTiO2 ceramic with x = 0.1.
Figure 6. (a) EDS spectrum of NYCTO + xTiO2 ceramic with x = 0.1; inset shows EDS testing area. (b) SEM image and corresponding SEM—EDS mapping images of (c) Na, (d) Y, (e) Cu, (f) Ti and (g) O for NYCTO + xTiO2 ceramic with x = 0.1.
Molecules 26 06043 g006
Figure 7. EDX spectra of NYCTO + xTiO2 ceramic with x = 0.2 detected at (a) grain and (b) GB; insets show the detected points in the microstructure.
Figure 7. EDX spectra of NYCTO + xTiO2 ceramic with x = 0.2 detected at (a) grain and (b) GB; insets show the detected points in the microstructure.
Molecules 26 06043 g007
Figure 8. Nonlinear current density (J)—Electric field (E) at room temperature for NYCTO + xTiO2 ceramics with x = 0.0, 0.1 and 0.2.
Figure 8. Nonlinear current density (J)—Electric field (E) at room temperature for NYCTO + xTiO2 ceramics with x = 0.0, 0.1 and 0.2.
Molecules 26 06043 g008
Figure 9. (a) Impedance complex plane plots (Z*) of NYCTO ceramic at various temperature (140–210 °C). (b) Arrhenius plot for grain boundary conductivity (Rgb).
Figure 9. (a) Impedance complex plane plots (Z*) of NYCTO ceramic at various temperature (140–210 °C). (b) Arrhenius plot for grain boundary conductivity (Rgb).
Molecules 26 06043 g009
Figure 10. Imaginary part of admittance (Y″) as a function of frequency at different temperatures (−60–0 °C) for NYCTO + xTiO2 ceramics with x = (a) 0.0, (b) 0.1 and (c) 0.2; their insets show the Arrhenius plots for the grain conductivity (σg).
Figure 10. Imaginary part of admittance (Y″) as a function of frequency at different temperatures (−60–0 °C) for NYCTO + xTiO2 ceramics with x = (a) 0.0, (b) 0.1 and (c) 0.2; their insets show the Arrhenius plots for the grain conductivity (σg).
Molecules 26 06043 g010
Figure 11. (ac) XPS spectra of Cu 2p3/2 for all sintered NYCTO + xTiO2 ceramics. (df) XPS spectra of Ti 2p for all sintered NYCTO + xTiO2 ceramics.
Figure 11. (ac) XPS spectra of Cu 2p3/2 for all sintered NYCTO + xTiO2 ceramics. (df) XPS spectra of Ti 2p for all sintered NYCTO + xTiO2 ceramics.
Molecules 26 06043 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saengvong, P.; Chanlek, N.; Putasaeng, B.; Pengpad, A.; Harnchana, V.; Krongsuk, S.; Srepusharawoot, P.; Thongbai, P. Significantly Improved Colossal Dielectric Properties and Maxwell—Wagner Relaxation of TiO2—Rich Na1/2Y1/2Cu3Ti4+xO12 Ceramics. Molecules 2021, 26, 6043. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26196043

AMA Style

Saengvong P, Chanlek N, Putasaeng B, Pengpad A, Harnchana V, Krongsuk S, Srepusharawoot P, Thongbai P. Significantly Improved Colossal Dielectric Properties and Maxwell—Wagner Relaxation of TiO2—Rich Na1/2Y1/2Cu3Ti4+xO12 Ceramics. Molecules. 2021; 26(19):6043. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26196043

Chicago/Turabian Style

Saengvong, Pariwat, Narong Chanlek, Bundit Putasaeng, Atip Pengpad, Viyada Harnchana, Sriprajak Krongsuk, Pornjuk Srepusharawoot, and Prasit Thongbai. 2021. "Significantly Improved Colossal Dielectric Properties and Maxwell—Wagner Relaxation of TiO2—Rich Na1/2Y1/2Cu3Ti4+xO12 Ceramics" Molecules 26, no. 19: 6043. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26196043

Article Metrics

Back to TopTop