Next Article in Journal / Special Issue
Lead-Free NaNbO3-Based Ceramics for Electrostatic Energy Storage Capacitors
Previous Article in Journal
Effect of Acid Surface Treatments on the Shear Bond Strength of Metal Bracket to Zirconia Ceramics
Previous Article in Special Issue
Nanosized Tungsten Powder Synthesized Using the Nitridation–Decomposition Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Origin of Temperature Coefficient of Resonance Frequency in Rutile Ti1−xZrxO2 Microwave Ceramics

1
Department of Physics, Abdul Wali Khan University Mardan, Mardan 23200, Pakistan
2
College of Electronics and Information, Hangzhou Dianzi University, Hangzhou 310018, China
3
Key Laboratory for Special Functional Materials in Jilin Provincial Universities, Jilin Institute of Chemical Technology, Jilin 132022, China
4
School of Optical and Electronic Information, Huazhong University of Science and Technology, Wuhan 430074, China
5
School of Instrumentation Science and Engineering, Harbin Institute of Technology, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
Ceramics 2024, 7(2), 698-711; https://doi.org/10.3390/ceramics7020046
Submission received: 15 March 2024 / Revised: 2 May 2024 / Accepted: 20 May 2024 / Published: 23 May 2024
(This article belongs to the Special Issue Advances in Electronic Ceramics)

Abstract

:
In this study, we report the effect of Zr4+ doping on the optical energy gap and microwave dielectric properties of rutile TiO2. Rietveld analysis explicitly confirmed that Zr4+ occupies the octahedral site, forming a single-phase tetragonal structure below the solubility limit (x < 0.10). Notably, at x = 0.025, a significant enhancement in Q × fo was observed. This enhancement was attributed to the reduction in dielectric loss, associated with a decrease in oxygen vacancies and a lower concentration of Ti3+ paramagnetic centers. This conclusion was supported by Raman and electron paramagnetic resonance spectroscopy, respectively. The origin of high τf in rutile Ti1−xZrxO2 is explained on the basis of the octahedral distortion/tetragonality ratio, covalency, and bond strength.

1. Introduction

TiO2 is one of the most studied metal oxides, used in many conventional applications (white pigment in paintings, paper, plastics, etc.) and advanced applications (microwave (MW) dielectric resonators, photocatalysts, resistive random-access memory (RRAM), and high-energy density capacitors) [1,2,3,4]. Naturally, TiO2 exists in different polymorphs i.e., brookite, rutile, and anatase. Among these phases, the thermodynamically stable phase is the rutile, as brookite and anatase phase can transform irreversibly to the rutile phase in the temperature range of 700–920 °C [5]. In several studies, colossal permittivity was found for (Mg2+ + Ta5+) co-doped TiO2 [6], (Tb3+/4+ and Sb5+) co-doped TiO2 [7], (Zr and Ta) co-doped TiO2 [8], (Lu3+/Nb5+) co-doped TiO2 [9], and (Nb and Mn) co-doped TiO2 [10]. Cohn et al. [2] reported MW properties of TiO2 for dielectric resonators, which gained significant attention [11]. TiO2 possesses a high relative permittivity (εr) of ~100 and a suitable quality factor (Q × fo) of ~10,000 GHz, but its poor temperature coefficient of resonant frequency ( τ f ) of ~450 ppm/°C is not suitable for practical applications [12]. Several studies have proved that the dependency of the MW dielectric properties of rutile on the synthesis conditions, such as low oxygen partial pressure/high sintering temperature, results in a coring effect due to the reduction of Ti4+ to Ti3+ [13,14]. Noh et al. [15] sintered TiO2 using spark plasma sintering and conventional sintering methods. The ɛr and Q × fo of SPS-TiO2 re-oxidized specimens were found to be 112.6 and 26,000, respectively. These properties were comparable to those samples sintered at 1300 °C using a solid-state route (ɛr = 101.3, Q × fo = 41,600). Similarly, Weng et al. [16] sintered TiO2 at 1040 °C/2 h and reported ɛr > 100, Q × fo > 23,000 GHz, and τ f = 200 ppm/°C. During sintering (in the air or under a low partial pressure of oxygen) at high temperatures, the generation of oxygen vacancies causes a coring effect, which significantly deteriorates the MW dielectric properties of TiO2 [17]. Oxygen vacancies in dense rutile TiO2 can be reduced by sintering it in an oxygen atmosphere or using a proper dopant. More recently, a tetravalent Hf4+ was doped in TiO2 (HfxTi1−xO2), which exhibited εr = 17, Q × fo = 84,020 GHz, and τ f = −47 ppm/°C (at x = 0.05) and εr = 37, Q × fo = 53,600 GHz, and τ f = −28 ppm/°C (at x = 0.5) [18]. Mao et al. [19] studied the effect of Mn2+ and W6+ dopants in TiO2, which improved the εr (~105) and Q × fo (~39,000 GHz), but τ f (~464.4 ppm/°C) further deteriorated. Similarly, other rutile TiO2-based systems, such as Zn0.15Nb0.3Ti0.55O2 [20], Bi-doped TiO2 [21], and Ni0.5Ti0.5TaO4 [22] ceramics, exhibit a high τ f . Similarly, Souza et al. [23] synthesized TiO2 − ZrO2 (x = 0, 0.1, and 0.2) and reported εr = 88, Q × fo = 16,285 GHz for the pure TiO2, εr = 85.4, Q × fo = 9765 GHz for x = 0.1, and εr = 79.2, Q × fo = 8883 GHz for x = 0.2; however, they did not report the temperature coefficient of the resonance frequency of these ceramics. Therefore, the purpose of the present work was to study the effect of the Zr4+ substation on the MW dielectric properties of rutile TiO2, which exhibited a high τf value; however, the quality factor was improved due to a reduction in the coring effect. Furthermore, a structure–property relationship was developed for the unaffected τ f in Zr-doped TiO2.

2. Experimental Procedures

The Ti1−xZrxO2 (x = 0–0.1) ceramics were prepared through a conventional solid-state route using reagent-grade TiO2 (99.9%) and ZrO2 (99.9%) precursors. The precursors were dried at 700 °C for 6 h to remove moisture (if any). The precursors were weighed and then mixed/milled using a mortar and pestle for 1 h in acetone. The mixed powders were calcined at 1000 °C (x = 0) and 1200 °C (x > 0) for 4 h at a heating/cooling rate of 5 °C/min in air. The calcined powders were re-milled and then pressed uniaxially into 10 mm-in-diameter cylindrical pellets. The green pellets were sintered in the temperature range of 1300−1400 °C for 4 h at a heating/cooling rate of 5 °C/min in air.
The phase analysis of the samples was carried out using an X-ray diffractometer (Bruker D8 Advance, Germany) with Cu-Kα radiation. For Rietveld analysis, the GSAS + EXPGUI package was used [24,25]. For background fitting, the Shifted Chebyshev polynomial was used with 12 terms. To increase the level of accuracy, the damping factor was kept at 9. To study the vibrational modes, the Raman spectra of all the samples were collected at room temperature using a Micro Ramboss spectrometer with a laser source of 532 nm and a power of 80 mW, with a frequency of 150–800 cm−1. Electron paramagnetic resonance (EPR) spectroscopy was carried out at room temperature using a Bruker BioSpin GMBH apparatus, Germany, with an X-band frequency of 9.843 GHz. The gyromagnetic factor (g) value was calculated by using the Equation (1),
g = h v β H
where h is the Planck constant ( h = 6.626 × 10−34 Js), v is the MW frequency, β is the Bohr magneton ( β = 5.788 × 10−15 eVT−1), and H is the strength of the magnetic field. The microstructure and elemental distribution of the samples were examined via scanning electron microscopy with energy-dispersive X-ray analysis (SEM/EDX, Thermo Fisher Scientific, Waltham, MA, USA). The diffuse reflectance spectrum of the sample was obtained using a Jasco-V750 UV–vis spectrometer. The MW dielectric properties of the samples were measured using a network analyzer (Keysight, E5071C ENA, Santa Rosa, CA, USA) with the TE01δ shielded cavity method [26]. The temperature coefficient of resonant frequency ( τ f ) was calculated with the following formula:
τ f = f 1 f 2 f 1 ( T 1 T 2 )
where f 1 and f 2 are the resonant frequencies at temperature T 1 and T 2 , respectively.

3. Results and Discussions

Figure 1a shows the room-temperature XRD patterns of Ti1−xZrxO2 (x = 0–0.1) ceramics sintered at their optimum sintering temperatures. All diffraction peaks of the samples matched JCPDS #89−4920 for rutile TiO2, having a tetragonal crystal structure with a space group P42/mnm. Figure 1b shows the Rietveld fitted and experimental data of Ti1−xZrxO2 (x = 0.025). Fitted plots for other samples are given in the Supplementary Information (Figure S1). The refined lattice parameters, cell volume, R factors (goodness of fitting or reliability factor), tetragonality (c/a), bond lengths (d), and theoretical density are given in Table 1. Due to the substitutionally doped Zr4+, the lattice parameters and cell volume have been increased and the diffraction peaks were shifted to the lower 2θ values with an increase in x [27]. The shift in the diffraction peaks is attributed to the larger ionic radius of Zr4+ (rVI = 0.72 Å) than Ti4+ (rVI = 0.605 Å) [28,29]. A secondary phase (ZrO2) was observed at x = 0.1, which suggests that the solubility limit of Zr4+ in TiO2 is less than 0.10 [30,31]. The tetragonality ratio has been increased from 0.6445 to 0.6476 with the increase in the Zr concentration.
Raman spectroscopy was carried out to confirm the rutile phase of Ti1–xZrxO2 (x = 0–0.1) ceramic samples. Rutile has 15 optical modes (i.e., 1A1g + 1A2g + 1A2u +1B1g + 1B2g + 2B1u + 1Eg + 3Eu) in which the Eg, B1g, B2g, and A1g are Raman active [32,33]. Figure 2 shows the bands that are indicative of the rutile symmetry, which were observed at about 144 cm−1 (B1g), 242 cm−1 (two phonon process), 441 cm−1 (Eg), and 606 cm−1 (A1g) for all the samples. Furthermore, when the Zr concentrations increased, the Eg peak moved to a lower wavenumber, showing that the doping Zr ions enter the host lattice, which expands [33]. The 144 cm−1 (B1g) peak is associated with the bending vibration of the O—Ti—O bond [27]. The peak at 242 cm−1 is the complex vibration peak due to the multi-phonon scattering (second-order effect), associated with the O—Ti—O bond’s complex vibration on rutile TiO2 [34]. Generally, the second-order effect was believed to be due to internal stress/strain or might be because of the Zr having a different ionic radius than that of the host [35]. The shift in the phonon peak towards the lower frequencies shows the reduction in the anharmonic behavior of the crystals, which shows an increase in crystal stability and enhancement of optical, dielectric, and thermal transport properties [36]. The Eg mode (441 cm−1) represents the vibrational motion of oxygen atoms along the c-axis, whereas the A1g mode (606 cm−1) is related to the Ti−O stretching vibration [27]. When Ti4+ ions in the lattice were replaced by Zr4+, the O2− ions could be more closely bound, because the Zr4+ ions have a higher electro-positivity than Ti4+. As a result of this, the vibrational motion of oxygen atoms along the c-axis is hindered, resulting in a decrease in the vibrational energy of oxygen atoms that corresponds to the Eg mode (441 cm−1), as well as a decrease in frequency. On the other hand, the replacement of Zr4+ ions has no effect on the A1g mode (606 cm−1), matching the Ti−O stretching vibration energy. Therefore, the peak at 606 cm−1 is not shifted with an increase in the Zr4+ concentration [27].
TiO2 belongs to a class of reducible oxides as it readily loses oxygen upon sintering at high temperatures in air with the formation of excess electrons and oxygen vacancies. The reduction in TiO2 can be expressed using Kroger–Vink notations, in terms of either the formation of Ti4+ interstitials, oxygen vacancies, Ti3+ interstitials, or Ti3+ species and oxygen vacancies in octahedral sites given by the Equations (3)–(6), respectively [14].
T i T i × + 2 O o × T i i + 4 e + O 2
2 O o × 2 V o + 4 e + O 2
T i T i × + 2 O o × T i i + 3 e + O 2
2 T i T i × + O o × 2 T i T i   + V o + 1 2 O 2
The excess electrons formed due to oxygen vacancies are usually stabilized at cations by forming Ti3+ as illustrated in Equation (5). This fact has been firmly confirmed, based on accurate theoretical calculations, that the electrons generated as a result of oxygen vacancies are preferentially stabilized on d orbitals of Ti [37] and were also shown by Chester [38] and Zimmermann [39]. For the detection of paramagnetic species, we have used EPR spectroscopy to investigate the bulk and surface defects in the rutile TiO2. Ti1−xZrxO2 (x = 0–0.1) samples were examined via conventional continuous wave-EPR (CW-EPR) by homogeneously changing the magnetic field in a specific range until a resonance condition was achieved, which entails the transition between spin states. Figure 3 displays the room temperature EPR spectra of rutile Ti1−xZrxO2 (x = 0–0.1), where the signal belongs to the paramagnetic center Ti3+ formed due the generation of excess electrons due to oxygen vacancies, consistent with the previous result [40]. The line of the EPR signal for rutile TiO2 recorded at room temperature tends to broaden and eventually disappear. This is because of the relaxation time of Ti3+, which causes the weakening of the EPR signal’s intensity at a temperature higher than 77 K [40]. With an increase in the Zr4+ concentration, the intensity of the signal decreases, which means the concentration of the paramagnetic center, implying that Ti3+ has been reduced in the sample. The corresponding g-value (1.97) belongs to the inner Ti3+ paramagnetic centers [41,42,43].
The relative density of rutile Ti1−xZrxO2 (x = 0–0.1) with an increasing Zr concentration was measured. At an optimum sintering temperature, a relative density of ~93% was achieved. Figure 4a–e shows the SEM micrographs of sintered, thermally etched, and polished rutile Ti1−xZrxO2 (x = 0–0.1) samples, illustrating their morphologies and grain-size distributions. The SEM images show a well-developed grain morphology and an almost dense microstructure with some pores or voids. The grain size increased from 5 to 19 µm as the Zr concentration increased from 0 to 0.1 (Figure 4f). This is because Zr promotes the grain-growth [44]. The composition (x = 0, sintered at 1300 °C) exhibited a uniform distribution of grain size, while the doped samples sintered at high temperatures (1375−1400 °C) displayed a bimodal distribution, which is an indication of the Ostwald-ripening mechanism for the grain-growth [45]. During sintering, grain growth is dominated by the grain boundary and volume diffusion, which results in larger grains [11]. The pores/voids of the doped samples have been decreased as compared to the pure rutile TiO2, but residual porosity has been observed in all the compositions. To further confirm the homogeneity in the samples, elemental mappings for Ti1−xZrxO2 (x = 0.025) are shown in Figure 4g–j, which are also supported by the EDX spectra, shown in Figure S2. At x = 0.10, the contrast of some small grains was different, consistent with the elemental mapping (Figure 5), which is in agreement with the XRD data (Figure 1), showing the formation of a secondary phase (ZrO2) [30,31]. This confirmed the presence of Ti and O in the TiO2 sample, while for the doped samples, Ti, O, and Zr were homogeneously distributed. However, for samples with x = 0.10, there are Zr-rich regions that show the formation of the secondary phase (Figure 5), consistent with the XRD and SEM.
Diffuse reflectance spectroscopy (DRS) studies were performed in the region 200–800 nm to evaluate the optical bandgap and the effect of Zr4+ doping on the bandgap of rutile TiO2. Besides studying the possible transitions between conduction and the valence band, we can also find any other transition due to impurities with DRS as well. Figure 6a shows the DRS spectrum of Ti1−xZrxO2 (x = 0–0.1) samples, indicating a sharp decrease in reflectance at about 310 nm for the un-doped rutile sample, caused by strong absorption. After increasing the concentration of Zr4+, the absorption edge shifted to a higher wavelength for x ≤ 0.075, showing a decrease in the optical band gap energy. This absorption may be attributed to the charge transfer between O2p and M3d (M = Ti, Zr) [46]. The optical band gap ( E g ) was calculated by solving the Kubelka-Munk equation (Equation (7)).
[ F ( R ) h υ ] = α ( h υ E g ) n
where F ( R ) is the Kubelka–Munk function, hv is energy, E g is band gap energy, and α is the proportionality constant. The power n describes the type of electronic transition and has a constant value, i.e., n = 0.5 for direct allowed and n = 2 for indirect allowed E g . Furthermore, n = 3/2 and n = 3 are used for direct and indirect forbidden energy gaps, respectively [47,48]. The key parameters affecting the bandgap of materials are the crystallinity, processing technique, cell parameters, and average particle size [49]. The E g values were calculated from the Tauc plot (Figure 6b). E g for the x = 0 sample was estimated to be 2.92 eV, which decreased to 2.62 eV with the Zr concentration from x = 0 to 0.075 because Zr doping will modify the band structure and hence band gap. It has been speculated that the position of the conduction band has decreased while the position of the valence band has remained unchanged [46,50,51]. With a further increase in the Zr concentration (x = 0.1), the bandgap increased to 2.66 eV due to the presence of the ZrO2 phase, which has a higher band gap value, i.e., 4.6−5.25 eV [52,53].
Table 2 shows the MW dielectric properties of the rutile Ti1–xZrxO2 (x = 0−0.1) ceramics. A high εr for rutile Ti1–xZrxO2 (96–103) was, observed which may be due to their characteristic extended apical bonds as compared to the equatorial bonds causing the rattling of cations in a rutile structure [20]. εr increased from 96 (x = 0) to 103 (x = 0.075) with an increase in the dopant (Zr) concentration and then decreased to 99 for sample x = 0.10. The enhancement of the εr could be related to the Zr substitution at the Ti lattice site [54]. This increase may also be attributed to the complex process of grain growth and densification of the ceramic [55,56]. Another crucial factor contributing to the enhancement of εr is the ionic polarizability. The ionic polarizability of Zr4+ (3.25 Å3) is higher than that of Ti4+ (2.93 Å3), which may be the phenomenon behind the increase in the εr [57]. Therefore, theoretical dielectric polarizability and ionic polarizabilities ( α theo ) were calculated using Shannon’s additivity rule [57] as follows:
α theo = ( 1 x ) α T i 4 + + x α Z r 4 + + 2 α O 2
where α T i 4 + , α Z r 4 + , and α O 2 are the ionic polarizabilities of Ti, Zr, and O, respectively. The increasing trend of the α theo is in good agreement with the increasing trend of εr; therefore, it can be concluded that in rutile Ti1−xZrxO2, the change in εr arises from the ionic polarizability difference, as shown in Table 2. The non-linear behavior of εr for sample x = 0.1 is because of the presence of ZrO2 as a secondary phase, which has an εr value of 15–40, which depends upon the frequency and crystal structure [58], leading to the decrease in the εr of the x = 0.1 sample. In addition, Ravichandran et al. [59] established a relationship between the εr and optical bandgap energy, i.e., the εr is inversely proportional to the E g [60], which is consistent with our study. With a decrease in the E g value, the εr values increased.
Generally, the microstructures, densities, grain size, crystal lattice, packing fraction, secondary phases, and valency of Ti in the Ti-based ceramics affect the Q × fo [61,62,63]. The Q × fo initially increased from 9500 (x = 0) to 32,366 GHz (at x = 0.025) and then decreased to 18,798 GHz (at x = 0.1). The relative density of the samples showed a decreasing trend with an increase in the Zr concentration, which is consistent with the previous study [23]. At 1300 °C, TiO2 has a density of ~93% but shows oxygen deficiency due to a low diffusion coefficient, making it unable to allow the oxygenation of the ceramic. However, oxygen diffused only to the surface of the ceramic because the core appears dark due to the presence of the reduced Ti3+ species, i.e., the bulk non-stoichiometry of oxygen is less, but it may be sufficiently large in the core to provoke the formation of condensation/shear structures of oxygen vacancies, just like that of Magnelli phases [17]. The reduction in TiO2 increases the dielectric loss, leading to the deterioration of the Q × fo. In comparison, the doping of Zr4+ in TiO2 has overcome the problem and enhanced the Q × fo value. This phenomenon can be explained using the shift in the Eg band towards the lower wavenumber (Figure 2) with an increasing Zr concentration, which is attributed to the decrease in the oxygen vacancies in the ceramics [32]. Similarly, this fact is also confirmed using EPR spectroscopy, which showed a decrease in the concentration of the Ti3+ paramagnetic center. The Q × fo of ceramics, having a rutile structure, also strongly depends on the c/a ratio [19,64], which agrees with the present findings (Table 1). The slight decrease in the Q × fo from 32,366 (x = 0.025) to 18,798 GHz (x = 0.1) may be due to an increase in the number of voids (Figure 4). The Q × fo and lattice vibrations of the rutile TiO2 can also be influenced by the packing fraction (PF). The PF shows the density of atomic stacking and can be calculated using Equation (9) [65]:
% P F = v o l u m e   o f   t h e   p a c k e d   i o n s v o l u m e   o f   t h e   u n i t   c e l l × Z
where Z shows the formula units per unit cell. As the Q × fo and PF are directly proportional, the decrease in Q × fo may also be attributed to the decrease in PF with an increase in the Zr concentration, as given in Table 2. At x = 0.10, the presence of the ZrO2 phase may also be responsible for the decrease in the Q × fo [23].
τ f slightly increased with an increase in the Zr concentration. The mechanism responsible for the increase in the τ f is mostly the increase in εr because the materials with high εr have high τ f and vice versa [66]. This may also be attributed to the substitution of Zr4+, a larger cation, for Ti4+, a smaller cation, which distorts the octahedra, increasing the τ f of the samples. Figure 7b shows the dependence of the τ f on the unit cell volume of Ti1−xZrxO2 (x = 0–0.1) ceramics. τ f is the combined effect of thermal expansion ( α L ), as well as the temperature coefficient of permittivity ( τ ε ), as shown in the Equation (10) [67]:
τ f = 1 2 τ ε α L
The α L value for ceramics is constant and in the range of 10 ppm/°C; therefore, the magnitude and sign of the τ f depend on τ ε . According to the Clausius–Mossotti relation, τ ε is given by Equations (11) and (12) [67].
τ ε = 1 ε ε T = ( ε 1 ) ( ε + 2 ) ε ( A + B + C )
A = 1 3 V V T p , B = 1 3 α m α m V T V T p , C = 1 3 α m α m T V
Here, α m represents polarizability, while Havinga and Bosman [67,68] relate the terms A and B to the expansion of volume, but their effects cancel out each other due to their equal magnitude and opposite sign. As shown by Lee [67], the term C is the restoring force acting upon the ions and correlates with the potential well’s shape, which shows its high dependence on the lattice parameters and structure. The τ ε proportionally decreased with the increase in the volume of the unit cell, as with an increase in x, which in turn increased the τ f .
The τ f is also related to the structural characteristics of the oxygen octahedra, such as the bond strength ( s ), bond length, and degree of covalency of constituents [65,69]. These characteristics can be calculated based on the degree of covalency, which is related to covalency ( f c ) and s using the following equations [70,71].
s = R R 1 N
f c = a R R 1 N M
d e g r e e   o f   c o v a l e n c y = a R R 1 N M R R 1 N × 100 = a s M s × 100 = f c s × 100
where R 1 , N , a , and M are constants. The values of R 1 and N depend on the cation site, while a and M depend on the number of electrons. The values of these constants are already reported in Refs. [70,71,72]. R is the average bond length obtained from the Rietveld refinement results. The degree of the Ti/Zr−O bond covalency increases from 38.441 to 38.756% with an increase in the Zr concentration as given in Table 3. Therefore, an increase in τ f values may be associated with the decrease in s, which is in good agreement with Q × fo, i.e., a decreasing trend from x = 0.025–0.1.
The relationship among the relative permittivity, structure parameter, and τ f value of rutile Ti1−xZrxO2 (x = 0–0.1) is also important. εr and τ f , for rutile tetragonal ceramic, mainly dependent on the tetragonality (c/a ratio) [73] or octahedral distortion [74]. The distortion of each TiO6 octahedra is affected by the c/a ratio and oxygen positional parameter ( μ ). The relationship between μ and c/a ratio is in the given (Equation (16)) [73].
μ = 2 4 2 1 d e d a 2 c a 2 + 2 1 2 4 1 d e d a 2
where d e and d a are the lengths of four equatorial and two apical Ti—O bonds, respectively, and can be calculated from a , c , and μ [64]. In special cases, when μ = μ * = 1 4 1 + 1 2 c a 2 , all the d e and d a bonds are equal. A perfect octahedron in a tetragonal structure would have an ideal c/a ≈ 0.586 and μ ideal = 1/2 (c/a)ideal, giving all twelve O—O bonds equal lengths [64]. The degree of octahedral distortion can be indicated by the c/a ratio up to some extent [64]. Figure 7a shows the dependence of τf and εr on the c/a ratio of Ti1−xZrxO2 (x = 0–0.1) ceramics. It is observed that the c/a ratio of Ti1−xZrxO2 increased with an increase in the Zr concentration showing an increase in εr and τf. The c/a ratio for the Ti1−xZrxO2 (x = 0–0.1) ceramic was higher than the ideal c/a ratio (~0.586) and increased with an increase in x, which resulted in a high distortion of the TiO6 octahedra. When the temperature increased, the rise in thermal energy was supposed to be absorbed completely in recovering the octahedral distortion, rather than in restoring the direct dependence of the polarizability on temperature [74], which means that the tetragonality ratio would change towards the ideal c/a ratio [64]. When the c/a ratio is closer to the (c/a)ideal (0.586), then there will be a slight change in c/a upon the increase in temperature from 25 to 85 °C. Therefore, an increase in the τf value for the Ti1−xZrxO2 solid solution is due to an increase in the c/a ratio.

4. Summary

In this study, Ti1−xZrxO2 ceramics were processed and their crystal structure, defect chemistry, microstructure, optical bandgap energy, and MW dielectric properties were investigated. Rutile TiO2 crystallized into a tetragonal structure (P42/mnm). At x = 0.1, a secondary phase (ZrO2) was observed, which shows that the solubility limit of Zr is below 0.10. The SEM micrographs showed an increase in the grain size from 5 µm to 19 µm with an increasing Zr concentration. The electron paramagnetic resonance spectroscopy showed the existence of Ti3+ species in the rutile sample, which reduces with the increase in the Zr4+ concentration. The Tauc plots showed a decreasing trend in the bandgap energies of Ti1−xZrxO2 (x = 0–0.075), i.e., 2.92−2.62 eV. The relative permittivity slightly increased, due to the extended apical bonds, as compared to the equatorial bonds causing the rattling of cations in the rutile structure. The quality factor initially increased with an increase in the Zr concentration (i.e., x = 0.025), which was explained based on a decrease in the coring effect and hence a decrease in oxygen vacancies. Above x = 0.025, the quality factor started decreasing due to a decrease in the packing fraction. Similarly, the temperature coefficient of the resonance frequency increased with an increase in x, due to an increase in the tetragonality (c/a ratio), which is indirectly associated with the distortion of the octahedra. Similarly, the degree of covalency also decreased, which is also a reason for the increase in τ f .

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ceramics7020046/s1, Figure S1: Rietveld refined and experimental pattern of rutile Ti1-xZrxO2 (a) x = 0, (b) x = 0.05, (c) x = 0.075, (d) x = 0.1; Figure S2: Energy dispersive X-Ray analysis of Ti1-xZrxO2 (a) x = 0, (b) x = 0.025, (c) x = 0.05, (d) x = 0.075, (e) x = 0.1.

Author Contributions

Conceptualization, R.M.; methodology, I.K. and A.K.; formal analysis, D.H., W.L. and D.W.; investigation, I.K. and A.K; resources, R.M., D.H., K.S., W.L. and D.W.; writing−original draft preparation, I.K.; writing−review and editing, K.S., M.M., D.W., W.L., R.M. and I.K. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support extended by the Directorate General of Science and Technology, Khyber Pakhtunkhwa, Pakistan through Project No. DGST/BSTIET/MS/2022/8773.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article and Supplementary Information.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Balunarayanan, R.; Aiswarya, S.; Sreekala, C.; Menon, S.K. Nano Cylindrical Dielectric Resonator Antenna Using Titanium dioxide for Wi-Fi Applications. Proceddings of the 2021 Fourth International Conference on Electrical, Computer and Communication Technologies (ICECCT), Coimbatore, India, 15–17 September 2021; IEEE: Piscataway, NJ, USA, 2021; pp. 1–6. [Google Scholar]
  2. Cohn, S.B. Microwave bandpass filters containing high-Q dielectric resonators. IEEE Trans. Microw. Theory Tech. 1968, 16, 218–227. [Google Scholar] [CrossRef]
  3. Kanehara, K.; Hoshina, T.; Takeda, H.; Tsurumi, T. Terahertz permittivity of rutile TiO2 single crystal measured by anisotropic far-infrared ellipsometry. J. Ceram. Soc. Jpn. 2015, 123, 303–306. [Google Scholar] [CrossRef]
  4. Mett, R.R.; Sidabras, J.W.; Anderson, J.R.; Klug, C.S.; Hyde, J.S. Rutile dielectric loop-gap resonator for X-band EPR spectroscopy of small aqueous samples. J. Magn. Reson. 2019, 307, 106585. [Google Scholar] [CrossRef] [PubMed]
  5. Hanaor, D.A.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef]
  6. 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]
  7. 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]
  8. 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. 2022, 48, 6016–6023. [Google Scholar] [CrossRef]
  9. Thanamoon, N.; Chanlek, N.; Srepusharawoot, P.; Swatsitang, E.; Thongbai, P. Microstructural Evolution and High-Performance Giant Dielectric Properties of Lu3+/Nb5+ Co-Doped TiO2 Ceramics. Molecules 2021, 26, 7041. [Google Scholar] [CrossRef]
  10. Yang, C.; Wei, X.; Hao, J. Disappearance and recovery of colossal permittivity in (Nb + Mn) co-doped TiO2. Ceram. Int. 2018, 44, 12395–12400. [Google Scholar] [CrossRef]
  11. Hsu, T.-H.; Huang, C.-L. Microwave dielectric properties of ultra-low-temperature-sintered TiO2 as a τ f compensator. Appl. Phys. A 2023, 129, 20. [Google Scholar] [CrossRef]
  12. 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]
  13. Grabstanowicz, L.R.; Gao, S.; Li, T.; Rickard, R.M.; Rajh, T.; Liu, D.-J.; Xu, T. Facile oxidative conversion of TiH2 to high-concentration Ti3+-self-doped rutile TiO2 with visible-light photoactivity. Inorg. Chem. 2013, 52, 3884–3890. [Google Scholar] [CrossRef] [PubMed]
  14. Templeton, A.; Wang, X.; Penn, S.J.; Webb, S.J.; Cohen, L.F.; Alford, N.M. Microwave dielectric loss of titanium oxide. J. Am. Ceram. Soc. 2000, 83, 95–100. [Google Scholar] [CrossRef]
  15. Noh, J.H.; Jung, H.S.; Lee, J.-K.; Kim, J.-R.; Hong, K.S. Microwave dielectric properties of nanocrystalline TiO2 prepared using spark plasma sintering. J. Eur. Ceram. Soc. 2007, 27, 2937–2940. [Google Scholar] [CrossRef]
  16. Weng, Z.; Wu, C.; Xiong, Z.; Feng, Y.; AminiRastabi, H.; Song, C.; Xue, H. Low temperature sintering and microwave dielectric properties of TiO2 ceramics. J. Eur. Ceram. Soc. 2017, 37, 4667–4672. [Google Scholar] [CrossRef]
  17. Pullar, R.C.; Penn, S.J.; Wang, X.; Reaney, I.M.; Alford, N.M. Dielectric loss caused by oxygen vacancies in titania ceramics. J. Eur. Ceram. Soc. 2009, 29, 419–424. [Google Scholar] [CrossRef]
  18. Ding, Y.H.; Liu, L.; Yang, Z.J.; Li, L.; Chen, X.M. Structure and microwave dielectric characteristics of Hf1–xTixO2 ceramics. J. Am. Ceram. Soc. 2022, 105, 1127–1135. [Google Scholar] [CrossRef]
  19. Mao, S.; Yang, J.; Gong, M.; Ao, L.; Fang, Z.; Kashif, K.; Qu, C.; Zhang, X.; Yang, H.; Xiong, Z.; et al. The improved microwave dielectric characteristics of TiO2 ceramics produced by Mn2+ and W6+ co-substitution. J. Mater. Sci. Mater. Electron. 2022, 33, 27041–27052. [Google Scholar] [CrossRef]
  20. Kim, E.S.; Kang, D.H. Relationships between crystal structure and microwave dielectric properties of (Zn1/3B2/35+)xTi1−xO2 (B5+ = Nb, Ta) ceramics. Ceram. Int. 2008, 34, 883–888. [Google Scholar] [CrossRef]
  21. Fukuda, K.; Awai, R.K. Microwave characteristics of TiO2-Bi2O3 dielectric resonator. Jpn. J. Appl. Phys. 1993, 32, 4584. [Google Scholar] [CrossRef]
  22. Wu, X.; Jing, Y.; Li, Y.; Su, H. Novel Tri-rutile Ni0.5Ti0.5TaO4 Microwave Dielectric Ceramics: Crystal Structure Chemistry, Raman Vibration Mode, and Chemical Bond Characteristic In-Depth Studies. J. Phys. Chem. C 2022, 126, 14680–14692. [Google Scholar] [CrossRef]
  23. Souza, J.V.C.; Castro, P.J.; Nono, M.d.C.d.A.; Mineiro, S.L. Microstructure, crystalline phase, and dielectric property analyses of TiO2 composition with ZrO2 addition. Mater. Sci. Forum. 2010, 660, 641–645. [Google Scholar] [CrossRef]
  24. Larson, A.C.; Von Dreele, R.B. General Structure Analysis System; Report LAUR 86-748; Los Alamos National Laboratory: Los Alamos, NM, USA, 2004.
  25. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  26. Krupka, J. Frequency domain complex permittivity measurements at microwave frequencies. Meas. Sci. Technol. 2006, 17, R55. [Google Scholar] [CrossRef]
  27. Wang, J.; Yu, Y.; Li, S.; Guo, L.; Wang, E.; Cao, Y. Doping behavior of Zr4+ ions in Zr4+-doped TiO2 nanoparticles. J. Phys. Chem. C 2013, 117, 27120–27126. [Google Scholar] [CrossRef]
  28. Chang, S.-M.; Doong, R.-A. Characterization of Zr-doped TiO2 nanocrystals prepared by a nonhydrolytic sol−gel method at high temperatures. J. Phys. Chem. B. 2006, 110, 20808–20814. [Google Scholar] [CrossRef] [PubMed]
  29. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  30. Gnatyuk, Y.; Smirnova, N.; Korduban, O.; Eremenko, A. Effect of zirconium incorporation on the stabilization of TiO2 mesoporous structure. Surf. Interface Anal. 2010, 42, 1276–1280. [Google Scholar] [CrossRef]
  31. Yu, J.C.; Lin, J.; Kwok, R.W. Ti1–xZrxO2 Solid Solutions for the Photocatalytic Degradation of Acetone in Air. J. Phys. Chem. B. 1998, 102, 5094–5098. [Google Scholar] [CrossRef]
  32. Swamy, V.; Muddle, B.C.; Dai, Q. Size-dependent modifications of the Raman spectrum of rutile TiO2. Appl. Phys. Lett. 2006, 89, 163118. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Harris, C.X.; Wallenmeyer, P.; Murowchick, J.; Chen, X. Asymmetric lattice vibrational characteristics of rutile TiO2 as revealed by laser power dependent Raman spectroscopy. J. Phys. Chem. C 2013, 117, 24015–24022. [Google Scholar] [CrossRef]
  34. Huang, W.; Zhang, Y.; Lu, J.; Chen, J.; Gao, L.; Omran, M.; Chen, G. Microwave drying method investigation for the process and kinetics of drying characteristics of high-grade rutile TiO2. Ceram. Int. 2023, 49, 15618–15628. [Google Scholar] [CrossRef]
  35. 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]
  36. Yu, C. Dielectric Materials for High Power Energy Storage; Queen Mary University of London: London, UK, 2017. [Google Scholar]
  37. Di Valentin, C.; Pacchioni, G.; Selloni, A. Reduced and n-type doped TiO2: Nature of Ti3+ species. J. Phys. Chem. C 2009, 113, 20543–20552. [Google Scholar] [CrossRef]
  38. Chester, P. Cross-Doping Agents for Rutile Masers. J. Appl. Phys. 1961, 32, 866–868. [Google Scholar] [CrossRef]
  39. Kiwi, J.; Suss, J.; Szapiro, S. EPR spectra of niobium-doped TiO2 and imiplications for water photocleavage processes. Chem. Phy. Lett. 1984, 106, 135–138. [Google Scholar] [CrossRef]
  40. Chiesa, M.; Paganini, M.C.; Livraghi, S.; Giamello, E. Charge trapping in TiO2 polymorphs as seen by Electron Paramagnetic Resonance spectroscopy. Phys. Chem. Chem. Phys. 2013, 15, 9435–9447. [Google Scholar] [CrossRef] [PubMed]
  41. Fresno, F.; Hernández-Alonso, M.D.; Tudela, D.; Coronado, J.M.; Soria, J. Photocatalytic degradation of toluene over doped and coupled (Ti, M) O2 (M = Sn or Zr) nanocrystalline oxides: Influence of the heteroatom distribution on deactivation. Appl. Catal. B 2008, 84, 598–606. [Google Scholar] [CrossRef]
  42. Kumar, C.P.; Gopal, N.O.; Wang, T.C.; Wong, M.-S.; Ke, S.C. EPR investigation of TiO2 nanoparticles with temperature-dependent properties. J. Phys. Chem. B 2006, 110, 5223–5229. [Google Scholar] [CrossRef]
  43. Livraghi, S.; Maurelli, S.; Paganini, M.C.; Chiesa, M.; Giamello, E. Probing the local environment of Ti3+ ions in TiO2 (rutile) by 17O HYSCORE. Angew. Chem. Int. Ed. 2011, 50, 8038–8040. [Google Scholar] [CrossRef]
  44. Zhong, M.; Li, J.; Shao, J.; Cao, Y.; Li, K.; Zhao, W. An investigation into the enhanced permittivity properties of Zr co-doped (Ga0.5Nb0.5)0.03Ti0.97O2 ceramics. Ceram. Int. 2019, 45, 14983–14990. [Google Scholar] [CrossRef]
  45. Chao, S.; Petrovsky, V.; Dogan, F. Effects of sintering temperature on the microstructure and dielectric properties of titanium dioxide ceramics. J. Mater. Sci. 2010, 45, 6685–6693. [Google Scholar] [CrossRef]
  46. Zhang, J.; Li, L.; Liu, D.; Zhang, J.; Hao, Y.; Zhang, W. Multi-layer and open three-dimensionally ordered macroporous TiO2–ZrO2 composite: Diversified design and the comparison of multiple mode photocatalytic performance. Mater. Des. 2015, 86, 818–828. [Google Scholar] [CrossRef]
  47. Khan, R.U.; Khan, I.; Ali, B.; Muhammad, R.; Samad, A.; Shah, A.; Song, K.; Wang, D. Structural, dielectric, optical, and electrochemical performance of Li4Mo5O17 for ULTCC applications. Mater. Res. Bull. 2023, 160, 112142. [Google Scholar] [CrossRef]
  48. Lee, S.; Woodford, W.H.; Randall, C.A. Band gap energy of perovskite structured ABO3 compounds. In Proceedings of the 2008 17th IEEE International Symposium on the Applications of Ferroelectrics, Santa Fe, NM, USA, 24–27 February 2008; Volume 1, p. 1. [Google Scholar]
  49. Yadav, P.; Sinha, E. Structural, photophysical and microwave dielectric properties of α-ZnMoO4 phosphor. J. Alloys Compd. 2019, 795, 446–452. [Google Scholar] [CrossRef]
  50. Du, F.; Yu, S. Preparation and Characterization of Zr–N-Codoped TiO2 Nano-Photocatalyst and Its Activity Enhanced-Mechanism. J. Nanosci. Nanotechnol. 2014, 14, 6965–6969. [Google Scholar] [CrossRef] [PubMed]
  51. Duan, B.; Zhou, Y.; Huang, C.; Huang, Q.; Chen, Y.; Xu, H.; Shen, S. Impact of Zr-doped TiO2 photocatalyst on formaldehyde degradation by Na addition. Ind. Eng. Chem. Res. 2018, 57, 14044–14051. [Google Scholar] [CrossRef]
  52. Hadi, E.H.; Sabur, D.A.; Chiad, S.S.; Fadhil, N. Physical properties of nanostructured li-doped ZrO2 thin films. J. Green Eng. 2020, 10, 8390–8400. [Google Scholar]
  53. Tomar, L.J.; Chakrabarty, B. Synthesis, structural and optical properties of TiO2-ZrO2 nanocomposite by hydrothermal method. Adv. Mater. Lett. 2013, 4, 64–67. [Google Scholar] [CrossRef]
  54. Juma, A.; Acik, I.O.; Oluwabi, A.; Mere, A.; Mikli, V.; Danilson, M.; Krunks, M. Zirconium doped TiO2 thin films deposited by chemical spray pyrolysis. Appl. Surf. Sci. 2016, 387, 539–545. [Google Scholar] [CrossRef]
  55. Padmamalini, N.; Ambujam, K. Structural and dielectric properties of ZrO2–TiO2–V2O5 nanocomposite prepared by CO-precipitation calcination method. Mater. Sci. Semicond. 2016, 41, 246–251. [Google Scholar] [CrossRef]
  56. Zhao, E.; Hao, J.; Xue, X.; Si, M.; Guo, J.; Wang, H. Rutile TiO2 microwave dielectric ceramics prepared via cold sintering assisted two step sintering. J. Eur. Ceram. Soc. 2021, 41, 3459–3465. [Google Scholar] [CrossRef]
  57. Shannon, R.D. Dielectric polarizabilities of ions in oxides and fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  58. Oh, Y.; Bharambe, V.; Mummareddy, B.; Martin, J.; McKnight, J.; Abraham, M.A.; Walker, J.M.; Rogers, K.; Conner, B.; Cortes, P.; et al. Microwave dielectric properties of zirconia fabricated using NanoParticle Jetting™. Addit. Manuf. 2019, 27, 586–594. [Google Scholar] [CrossRef]
  59. Ravichandran, R.; Wang, A.X.; Wager, J.F. Solid state dielectric screening versus band gap trends and implications. Opt. Mater. 2016, 60, 181–187. [Google Scholar] [CrossRef]
  60. Hervé, P.; Vandamme, L. General relation between refractive index and energy gap in semiconductors. Infrared Phy. Techn. 1994, 35, 609–615. [Google Scholar] [CrossRef]
  61. Fang, Z.-X.; Tang, B.; Li, E.; Zhang, S.-R. High-Q microwave dielectric properties in the Na0.5Sm0.5TiO3+Cr2O3 ceramics by one synthetic process. J. Alloys Compd. 2017, 705, 456–461. [Google Scholar] [CrossRef]
  62. Fang, Z.; Yang, H.; Yang, H.; Xiong, Z.; Zhang, X.; Zhao, P.; Tang, B. Ilmenite-type MgTiO3 ceramics by complex (Mn1/2W1/2)4+ cation co-substitution producing improved microwave characteristics. Ceram. Int. 2021, 47, 21388–21397. [Google Scholar] [CrossRef]
  63. Yu, S.; Tang, B.; Zhang, S.; Zhou, X. The effect of Mn addition on phase development, microstructure and microwave dielectric properties of ZrTi2O6–ZnNb2O6 ceramics. Mater. Lett. 2012, 80, 124–126. [Google Scholar] [CrossRef]
  64. Pang, L.X.; Wang, H.; Zhou, D.; Yao, X. Sintering behavior, structures, and microwave dielectric properties of (LixNb3x)Ti1−4xO2. J. Am. Ceram. Soc. 2008, 91, 2947–2951. [Google Scholar] [CrossRef]
  65. Kim, E.S.; Chun, B.S.; Freer, R.; Cernik, R.J. Effects of packing fraction and bond valence on microwave dielectric properties of A2+ B6+ O4 (A2+: Ca, Pb, Ba; B6+: Mo, W) ceramics. J. Eur. Ceram. Soc. 2010, 30, 1731–1736. [Google Scholar] [CrossRef]
  66. Reaney, I.M.; Iddles, D. Microwave dielectric ceramics for resonators and filters in mobile phone networks. J. Am. Ceram. Soc. 2006, 89, 2063–2072. [Google Scholar] [CrossRef]
  67. Lee, H.-J.; Hong, K.-S.; Kim, S.-J.; Kim, I.-T. Dielectric Properties of MNb2O6 Compounds (Where M = Ca, Mn, Co, Ni, or Zn). Mater. Res. Bull. 1997, 32, 847–855. [Google Scholar] [CrossRef]
  68. Bosman, A.J.; Havinga, E.E. Temperature dependence of dielectric constants of cubic ionic compounds. Phys. Rev. 1963, 129, 1593. [Google Scholar] [CrossRef]
  69. Yang, H.; Zhang, S.; Yang, H.; Chen, Y.; Yuan, Y.; Li, E. Effects of ZrO2 substitution on crystal structure and microwave dielectric properties of Zn0.15Nb0.3(Ti1–xZrx)0.55O2 ceramics. Ceram. Int. 2018, 44, 22710–22717. [Google Scholar] [CrossRef]
  70. Brown, I.t.; Shannon, R. Empirical bond-strength–bond-length curves for oxides. Acta Crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1973, 29, 266–282. [Google Scholar] [CrossRef]
  71. Brown, I.t.; Wu, K.K. Empirical parameters for calculating cation–oxygen bond valences. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1976, 32, 1957–1959. [Google Scholar] [CrossRef]
  72. Ding, Y.H.; Wang, X.; Guo, R.Z.; Li, L.; Chen, X.M. Improvement of τf in HfTiO4 microwave dielectric ceramics with Zr-and Sn-substitution. J. Am. Ceram. Soc. 2024, 107, 2337–2347. [Google Scholar] [CrossRef]
  73. Choi, J.W.; Van Dover, R. Correlation between temperature coefficient of resonant frequency and tetragonality ratio. J. Am. Ceram. Soc. 2006, 89, 1144–1146. [Google Scholar] [CrossRef]
  74. Kim, E.S.; Kang, D.H.; Kim, S.J. Effect of crystal structure on microwave dielectric properties of (Ni1/3B2/3)1–xTixO2 (B = Nb and Ta). Jpn. J. Appl. Phys. 2007, 46, 7101. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of rutile Ti1−xZrxO2 (x = 0−0.1), and (b) fitted XRD patterns for x = 0.025.
Figure 1. (a) XRD patterns of rutile Ti1−xZrxO2 (x = 0−0.1), and (b) fitted XRD patterns for x = 0.025.
Ceramics 07 00046 g001
Figure 2. Raman spectra of rutile Ti1−xZrxO2 (x = 0–0.1).
Figure 2. Raman spectra of rutile Ti1−xZrxO2 (x = 0–0.1).
Ceramics 07 00046 g002
Figure 3. The room-temperature EPR spectra of rutile Ti1−xZrxO2 (x = 0–0.1).
Figure 3. The room-temperature EPR spectra of rutile Ti1−xZrxO2 (x = 0–0.1).
Ceramics 07 00046 g003
Figure 4. SEM micrographs of thermally etched and polished Ti1−xZrxO2 at (a) x = 0, (b) x = 0.025, (c) x = 0.05, (d) x = 0.075, (e) x = 0.1; (f) grain size distribution as a function of the Zr concentration; (g) elemental distribution and EDX mapping of (h) Ti, (i) O, and (j) Zr for a sample with x = 0.025.
Figure 4. SEM micrographs of thermally etched and polished Ti1−xZrxO2 at (a) x = 0, (b) x = 0.025, (c) x = 0.05, (d) x = 0.075, (e) x = 0.1; (f) grain size distribution as a function of the Zr concentration; (g) elemental distribution and EDX mapping of (h) Ti, (i) O, and (j) Zr for a sample with x = 0.025.
Ceramics 07 00046 g004
Figure 5. EDX mapping of (a) Ti1−xZrxO2 (x = 0.1), (b) Ti, (c) O, and (d) Zr.
Figure 5. EDX mapping of (a) Ti1−xZrxO2 (x = 0.1), (b) Ti, (c) O, and (d) Zr.
Ceramics 07 00046 g005
Figure 6. (a) Reflection spectra and (b) Tauc plots of Ti1−xZrxO2 (x = 0–0.1).
Figure 6. (a) Reflection spectra and (b) Tauc plots of Ti1−xZrxO2 (x = 0–0.1).
Ceramics 07 00046 g006
Figure 7. (a) Relationship between the c/a ratio and τf value and/or εr and (b) the variation in τf with the unit cell volume of Ti1−xZrxO2 (x = 0–0.1).
Figure 7. (a) Relationship between the c/a ratio and τf value and/or εr and (b) the variation in τf with the unit cell volume of Ti1−xZrxO2 (x = 0–0.1).
Ceramics 07 00046 g007
Table 1. Rietveld refined lattice parameters, atomic positions, and refinement factors of Ti1–xZrxO2 (x = 0–0.1) ceramics.
Table 1. Rietveld refined lattice parameters, atomic positions, and refinement factors of Ti1–xZrxO2 (x = 0–0.1) ceramics.
x Values 00.0250.050.0750.1
a = b (Å) 4.5890 (7)4.5923 (2)4.6001 (5)4.6061 (5)4.6137 (9)
c (Å) 2.9579 (8)2.9635 (7)2.9732 (6)2.9811 (1)2.9882 (8)
α = β = γ (°) 9090909090
c/a 0.64450.64530.64630.64710.6476
Vcell (Å)3 62.2962.4562.9263.2663.61
Rwp (%) 11.3614.1212.0611.9812.97
Rp (%) 8.8210.919.179.2110.09
χ2 1.492.141.661.711.96
Atomic positions
(x, y, z)
Ti(0, 0, 0)(0, 0, 0)(0, 0, 0)(0, 0, 0)(0, 0, 0)
O(0.3025,
0.3025,
0)
(0.3053,
0.3053,
0)
(0.3026,
0.3026,
0)
(0.3050,
0.3050,
0)
(0.2956,
0.2956,
0)
Zr (0, 0, 0)(0, 0, 0)(0, 0, 0)(0, 0, 0)
Table 2. MW dielectric properties and optical bandgap of Ti1−xZrxO2.
Table 2. MW dielectric properties and optical bandgap of Ti1−xZrxO2.
x ST (°C)Relative Density (%)εrαtheo3)Q × fo (GHz)τf (ppm/°C)Packing Fraction (%)Bandgap Energy (eV)
0130093966.95950041770.62.92
0.025137591996.8132,36042470.42.84
0.051375891016.9629,71043270.02.81
0.0751400891036.9721,05042969.62.62
0.11400--996.9818,80045369.32.68
Table 3. Bond strength and covalency of Ti1−xZrxO2 (x = 0–0.1).
Table 3. Bond strength and covalency of Ti1−xZrxO2 (x = 0–0.1).
xBond TypeR (Å)R1NsfcCovalency (%)Degree of Covalency
(%)
0Ti−O1.96011.8065.20.65320.251138.44138.441
0.025Ti−O1.96331.8065.20.64770.247738.25538.496
Zr−O1.95060.96000.459647.874
0.05Ti−O1.96641.8065.20.64240.244638.07738.554
Zr−O1.95060.95090.452847.616
0.075Ti−O1.96991.8065.20.63650.241137.87638.585
Zr−O1.95060.94080.445347.327
0.1Ti−O1.97101.8065.20.63460.240037.81438.756
Zr−O1.95060.93770.442947.237
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, I.; Khan, A.; Muhammad, R.; Mao, M.; Han, D.; Song, K.; Lei, W.; Wang, D. Origin of Temperature Coefficient of Resonance Frequency in Rutile Ti1−xZrxO2 Microwave Ceramics. Ceramics 2024, 7, 698-711. https://doi.org/10.3390/ceramics7020046

AMA Style

Khan I, Khan A, Muhammad R, Mao M, Han D, Song K, Lei W, Wang D. Origin of Temperature Coefficient of Resonance Frequency in Rutile Ti1−xZrxO2 Microwave Ceramics. Ceramics. 2024; 7(2):698-711. https://doi.org/10.3390/ceramics7020046

Chicago/Turabian Style

Khan, Izaz, Aneela Khan, Raz Muhammad, Minmin Mao, Dandan Han, Kaixin Song, Wen Lei, and Dawei Wang. 2024. "Origin of Temperature Coefficient of Resonance Frequency in Rutile Ti1−xZrxO2 Microwave Ceramics" Ceramics 7, no. 2: 698-711. https://doi.org/10.3390/ceramics7020046

Article Metrics

Back to TopTop