Swagata Bandyopadhyay and
Abhigyan Dutta*
Department of Physics, The University of Burdwan, Burdwan-713104, India. E-mail: adutta@phys.buruniv.ac.in; Fax: +91 342 2634015; Tel: +91 342 2657800
First published on 27th July 2015
Room temperature phase stabilization of cubic Bi2O3 has been achieved by adding Dy2O3 as the dopant, using a low temperature citrate-auto-ignition method. The samples were sintered at different temperatures retaining the cubic fluorite structure. Rietveld refinement of the X-ray diffraction profiles has given detailed microstructural information of the prepared samples. The transmission electron micrographs confirmed the presence of atomic planes as obtained from X-ray diffraction. The UV-Vis spectra show a red shift of the absorption peak with the increase in sintering temperature. Impedance spectroscopy studies of the samples exhibited thermally activated non-Debye type relaxation process. In addition, studies of the electrical conductivity have suggested the negative temperature coefficient of resistance (NTCR) behavior of the samples. The comparable values of activation energies, obtained from different parameters, indicated that the ions follow the same type of mechanism for conduction as well as for relaxation. The temperature independence of the mechanisms has been confirmed from scaling of different spectra. The correlation between structural and electrical properties of the samples has been discussed and interpreted accordingly.
A controversy on the stability of δ-Bi2O3 phase at room temperature aroused because Watanabe10 claimed that, δ-Bi2O3 could not be stabilized by any oxide addition in long-term aspect. The stabilized bismuth oxides were treated as a quench-in phase or in other words, meta-stabilized phase. (A) Systematic study9 on the effects of different rare earth dopants on Bi2O3–Ln2O3 (Ln = Dy, Ho, Er, Tm and Yb) system showed that, the degradation rate of conductivity depends on the ionic radii and doping concentration. It was also observed that, two different mechanisms were responsible for degradation. One is phase transformation from fluorite to rhombohedral, which occurs most rapidly at about 650 °C and the other one is the reversible order–disorder transformation of the vacancies, which occurs at about 500 °C. It was shown that, the relative decay in conductivity is lowest for Dy2O3 stabilized cubic Bi2O3.9,11 A recent study12 on the energetic of doped Bi2O3 has also confirmed that, decrease in the volatility and susceptibility to reduction of bismuth oxide at high temperatures has been lowest for the Dy doping. Moreover, the dopant cations affect the nature of the vacancy arrangements in the oxygen sublattice and the polarizability of the cation affects the stability of the disordered oxygen sublattice.13,14 In a recent work15 on Dy doped cubic bismuth oxide system, it has been shown that, the oxygen ionic conductivity has been much dependent on crystallite size, prepared through different synthesis routes.
In the present work, Bi2O3 is successfully stabilized down to room temperature and δ-phase has been found to be stable for long time duration. This has been achieved by singly doping of Dy2O3, keeping the as prepared sample at lowest possible sintering temperature for the lowest time duration.
The main objectives of this work are two folds. Firstly, to prepare the room temperature phase stabilized Dy doped cubic δ-Bi2O3 and secondly, to investigate the effect of sintering temperature on various structural, microsructural and ionic transport properties of prepared samples to enlighten new properties of this interesting solid electrolyte material. To the best of our knowledge, the analysis of electrical relaxation phenomena of this material has not been done yet.
For doped Bi0.7Dy0.3O1.5−δ, stoichiometric amounts of Bi(NO3)3·5H2O and Dy2O3 (ALFA AESAR, 99.9%) were used as precursors. First Dy2O3 was dissolved in 50 ml conc. HNO3, diluted by 100 ml de-ionized water and the solution was stirred at 70 °C temperature using magnetic stirrer for 1 h, until the solution became clear. After that, proper amount of weighted Bi(NO3)3·5H2O was mixed in the nitrate solution and again stirred for 2 h. Then citric acid at 1:
3 molar ratios with the Bi cations was dissolved in 200 ml de-ionized water added to the solution. The solution was stirred for 4 h at temperature 85 °C. Due to constant evaporation the solution became yellowish and gel/foam was formed and auto-ignition process was completed within a minute and produced yellowish powder samples of doped compositions.
The agglomerated yellowish ash was crushed and grinded for 15 min for homogeneous mixing. The doped samples were divided into three parts and sintered at 600 °C, 700 °C, 800 °C respectively and pure Bi2O3 was sintered at 600 °C, each for 6 h duration with furnace cooling to room temperature. The samples are abbreviated as DSB1 (600 °C), DSB2 (700 °C) and DSB3 (800 °C) respectively.
Powder X-ray diffraction data were obtained at room temperature by X-ray diffractometer (BRUKER, Model D8 Advance-AXS) using CuKα radiation (λ = 1.5414 Å) in 2θ range from 20–80°, with step size 0.02° and scan rate 0.5 s per step. The microstructure of the pure and doped samples were also studied using high-resolution transmission electron microscope (HR-TEM) (JEOL; Model: JEM-2010) operated at 200 KV. For HR-TEM study, a pinch of powdered sample was dispersed ultrasonically in ethyl alcohol and a drop of colloidal solution was placed on a 300 mesh carbon coated Cu grid. The grid was dried over night under high power lamp.
Optical band gap of the pure and doped samples were determined from ultraviolet-visible (UV-Vis) absorption spectra, taken at room temperature using Shimadzu spectrophotometer (Model: 1800), in the wavelength range 200–800 nm. For UV-Vis study, powdered samples were dispersed ultrasonically in ethyl alcohol and ethyl alcohol was used as reference during the absorption measurement.
For electrical measurement, powdered samples were uniaxially pressed in a 10 mm diameter stainless steel die into cylindrical shaped pellets of average thickness about 0.17 cm. Acetone was used as binder material. The pellets were annealed in air at 200 °C for 2 h. The pellets were then polished to acquire smooth surface and high temperature conductive graphite pest (MERCK) was applied on both sides of the pellets to make the electrodes. Subsequently, the coated samples were annealed again at 200 °C for 30 min.
Electrical measurements were performed by two probe method using an impedance analyzer (HIOKI; Model: 3532-50) in the frequency range 42 Hz–5 MHz and in temperature range 200–440 °C with 20 °C step in air atmosphere inside a tube furnace.
For Rietveld analysis, MAUD 2.33 (ref. 16) software has been employed. It is designed to refine simultaneously both the structural (unit cell parameters and atomic positions and occupancies) and micro structural parameters (crystallite size and r.m.s. strain). Various structural parameters including lattice parameter (a) and micro structural parameters like particle size (D), r.m.s. microstrain (〈ε2〉1/2)17 have been refined by a standard least square method assuming the peak shapes as a pseudo-Voigt function.18,19 Instrumental broadening has been corrected by a specially processed standard Si sample. The background of the XRD patterns is fitted with a polynomial of degree 5. The Marquardt least-squares method has been employed to minimize the difference between experimental and simulated patterns.
This minimization is monitored using two reliability index parameters Rwp (weighted residual error factor) and Rexp (expected error factor) defined as:
![]() | (1) |
![]() | (2) |
The Rietveld refined patterns for pure Bi2O3 and Dy doped Bi2O3 samples are shown in Fig. 1(b) and (c) respectively. The monoclinic and cubic phase has been simulated by considering the COD file no. 9012546 and 9009850 respectively. Pure Bi2O3 contains single monoclinic phase with lattice parameter a = 5.8479 Å, b = 8.1660 Å, c = 7.5094 Å and α = γ = 90°, β = 112.987° and cell volume = 330.13 (Å)3. This monoclinic phase is converted into single FCC cubic fluorite phase with cell volume ∼165 (Å)3. This decrease in cell volume with doping is due to the fact that ionic radius of Dy3+ (1.027 Å) is lesser than the ionic radius of Bi3+ (1.17 Å).
The Rietveld analysis has exhibited that, no additional phase corresponding to Dy has been observed in the XRD profiles for the doped samples, which ensures the complete dissolution of the dopant into the host matrix. The sharp diffraction peaks in intensity as well as the narrow spectral width indicate the high quality of crystallinity of the samples. The variation of crystallite size (D) and r.m.s. microstrain (〈ε2〉1/2) with sintering temperature of the doped DSB samples are shown in Fig. 1(d). From this figure, it is clear that, crystallite size (D) increases where as microstrain decreases with increase in sintering temperature.
The increase in lattice parameter and crystallite size with increase in sintering temperature indicates the grain growth of the samples with sintering temperature. This is because of the fact that, the thermal energy generates a force that reduces the total area of grain boundary. Due to the reduction of total area of grain boundary, the actual number of grains per unit volume decreases and grain size increases. This grain growth can further be explained by the following equation,20
Dn − D0n = Kt | (3) |
![]() | (4) |
Structural models of the pure and doped Bi2O3 samples have been generated using the atomic coordinates obtained from Rietveld analysis and the models are presented in Fig. 2(a) and (b) respectively. In the crystal structure of pure Bi2O3, as shown in Fig. 2(a), all the Bismuth atoms (Bi) are situated at the 4e site with the atomic co-ordinate Bi1 (0.524, 0.1831, 0.3613), Bi2 (0.0409, 0.0425, 0.7762) and all oxygen atoms are situated at the 4e sites with the atomic co-ordinate O1 (0.78, 0.30, 0.71), O2 (0.242, 0.044, 0.134), O3 (0.271, 0.024, 0.513). All the atomic positions were fixed during the whole refinement process. When this pure Bi2O3 is doped with Dy atom, the monoclinic structure is converted into FCC cubic structure, which is shown in Fig. 2(b). Here all Bi and Dy atoms are situated at the 4a site and oxygen atoms (O1 and O2) are situated at the 8c and 32f sites respectively. The Wyckoff position and fractional occupancy of the atoms for the sample DSB3 are listed in Table 1. Values of different structural and microstructural parameters like lattice parameter (a), cell volume (V), crystallite size (D), r.m.s. microstrain (〈ε2〉1/2), Rwp, Rexp, GoF are listed in Table 2. The Rietveld refinement also shows that r.m.s. microstrain decreases with increase of sintering temperature which indicates the better stability of the samples with rise in sintering temperature.
![]() | ||
Fig. 2 (a) Crystal structure of pure Bi2O3 sample possessing monoclinic structure. (b) Crystal structure of Dy doped stabilized δ-Bi2O3 sample possessing face centred cubic structure. |
Atom | Wyckoff position | Atomic coordinate (x,x,x) | Fractional occupancy |
---|---|---|---|
Bi/Dy | 4a | 0 | 0.6932/0.2968891 |
O1 | 8c | 0.25 | 0.18615 |
O2 | 32f | 0.3617 | 0.29373 |
Sample Name | Lattice parameter a (Å) | Cell volume V (Å)3 | Crystallite size D (nm) | r.m.s. microstrain 〈ε2〉1/2 (× 10−4) | Rwp (%) | Rexp (%) | GoF |
---|---|---|---|---|---|---|---|
DSB1 | 5.4837 | 164.8988 | 40.53 | 17.10 | 16.5 | 12.64 | 1.30 |
DSB2 | 5.4852 | 165.0327 | 88.44 | 6.28 | 25.7 | 15.23 | 1.68 |
DSB3 | 5.4868 | 165.1833 | 121.40 | 2.73 | 20.4 | 15.25 | 1.33 |
The orientation of the lattice plane is confirmed by using the simulated lattice pattern in Fig. 3(b) with orientation along (200) direction. In the inset of Fig. 3(b), the exact lattice spacing (dTEM) ∼0.296 nm is confirmed by the difference of the peak heights. This lattice spacing (dTEM) value is comparable with that obtained from Rietveld analysis (dXRD) for (200) plane of the same sample. The proof of crystal symmetry is given by using the selected area electron diffraction pattern as shown in Fig. 3(c). In this figure, the bright and intense dots can be indexed in a cubic cell and absence of intermediate small dots also confirms the phase purity even in the nano scale region. From this reciprocal lattice reconstruction of DSB3, it can be concluded that, the structure is not associated with any superstructure. The pattern in Fig. 3(c) has been found to be tilted at 2.6° with positive x-axis. So the arrow heads are found to be tilted. Another high resolution TEM image of the sample is also shown in Fig. 3(d). To confirm the contributing lattice planes, we have also shown the fast Fourier transform (FFT) pattern in the left upper panel of Fig. 3(e) which is taken from the selected part of Fig. 3(d) as indicated by the red box. It can be confirmed from the distinct circular rings in the FFT image that, the contributing lattice planes are (200) and (220). The simulated lattice pattern signifies the different orientations of these two planes in the right upper and lower panel of Fig. 3(e). Further confirmation of cationic position is given by using the atomic model as shown in Fig. 3(f). Comparing with Fig. 3(d) and (f), it can be concluded that, Bi/Dy is localized at the same Wyckoff site and no distinction is possible due to their common site sharing.
(αEP) = A(EP − Eg)q | (5) |
Sample name | Activation energy for dc conductivity (Eaσ) (eV) | Activation energy for relaxation (Eaτ) (eV) | Activation energy for most probable relaxation time τm (Eτm) (eV) | Optical band gap (Eg) (eV) |
---|---|---|---|---|
Bi2O3 | 1.16 | — | 1.10 | 2.71 |
DSB1 | 1.17 | 1.16 | 1.17 | 2.46 |
DSB2 | 1.16 | 1.14 | 1.15 | 2.21 |
DSB3 | 1.15 | 1.11 | 1.13 | 2.07 |
In addition, the wavelength, at which the strong absorption occurs, decreases for the samples from Bi2O3 to DSB3. It exhibits the red shift of the spectra with the increasing sintering temperature. This red shift can be partially explained by the shrinkage of energy band gap with an increase in particle size.22 The increased particle size due to grain growth at higher sintering temperature results in larger number of atoms in the particle causing the greater splitting of energy level resulting in the shrinkage of optical band gap. This is also consistent with the XRD results.
Z*(ω) = Z′(ω) − jZ′′(ω) | (6) |
ε*(ω) = ε′(ω) − jε′′(ω) | (7) |
M*(ω) = M′(ω) + jM′′(ω) | (8) |
![]() | (9) |
![]() | (10) |
Fig. 5(c) shows the complex impedance plots for all the samples at a fixed measuring temperature. This figure shows that, both bulk and gain boundary resistance decreases and total conductivity increases with increase of sintering temperature. The reciprocal temperature dependence of the total conductivity for all the samples are shown in Fig. 5(d) which obeys the Arrhenius equation given by,
![]() | (11) |
It has been found that, the values of bulk and grain boundary resistance is much higher for the pure sample in comparison with the doped samples because, stabilization of the monoclinic Bi2O3 structure into cubic fluorite δ-Bi2O3 structure increases the number of oxygen vacancies to a great extent. In case of doped samples, total conductivity increases with increase in sintering temperature because, crystallite size increases which results in the decrease in grain boundary and consequently its contribution to the total conductivity. The increase of conductivity with sintering temperature can also be understood by measuring a term, called grain boundary blocking factor (α) which is expressed as,24
![]() | (12) |
The different fitting parameters (ai, Ri) obtained from complex impedance spectra and calculated values of capacitance (Ci) and conductivity for the bulk (σb) and grain boundary (σgb) contribution at 300 °C for the samples are listed in Table 4. For all the samples, the value of frequency exponent (ai) for both bulk and grain contribution is less than unity (ai < 1), indicating the presence of non-Debye type relaxation for all the compositions.
Sample name | Grain | Grain boundary | ||||||
---|---|---|---|---|---|---|---|---|
ab | Cb (pF) | Rb (Ω) (× 104) | σb (Ω−1 cm−1) (× 10−5) | agb | Cgb (nF) | Rgb (Ω) (× 104) | σgb (Ω−1 cm−1) (× 10−6) | |
Bi2O3 | 0.98 | 35.36 | 20.49 | 0.11 | 0.97 | 0.15 | 24.64 | 0.87 |
DSB1 | 0.99 | 156.13 | 1.96 | 1.10 | 0.99 | 2.14 | 7.54 | 2.86 |
DSB2 | 0.98 | 158.95 | 1.86 | 1.16 | 0.82 | 31.50 | 6.29 | 3.43 |
DSB3 | 0.99 | 172 | 1.59 | 1.36 | 0.98 | 82.50 | 3.04 | 7.08 |
Fig. 6(a) shows the frequency variation of the imaginary part of the complex impedance (Z′′) at different temperatures for a particular sample (DSB3). Z′′ value shows a peak maxima (Z′′max) at a particular frequency (relaxation frequency ωmax) and this peak shifts towards higher frequencies with increasing temperature indicating the presence of relaxation in the system.25 The peak frequency ωmax is characterized by composition and structural arrangements of atoms in the sample.
Fig. 6(b) shows the scaling behavior of Z′′ spectra. Here the frequency axis is scaled by relaxation frequency ωmax and Z′′ is scaled by Z′′max. The spectra for all the compositions obey the time–temperature superposition principle (TTSP) i.e. all the Z′′ spectra at different temperatures superimposed on a single master curve. This behavior simply indicates that, the change in temperature only changes the number of charge carriers without changing the conduction mechanism.26
Using the value of ωmax, the relaxation time has been calculated using the equation,
ωmaxτ = 1 | (13) |
Fig. 7 shows the variation of τ with reciprocal of absolute temperature. It has been observed that, the value of τ decreases with increase of measuring temperature for all the samples and its temperature dependence nature follows the Arrhenius relation given by,
![]() | (14) |
The most probable relaxation time τm can be calculated from the peak frequency ωmax using the relation ωmaxτm = 1. The inset of Fig. 8(a) shows the reciprocal temperature dependence of ωmax for all the compositions which also obeys the Arrhenius relation.
The values of the activation energy for the most probable relaxation time (Eτm) obtained from the slope of the curves are given in Table 3 which again show the same type of variation with the increase in sintering temperature. All the three types of activation energies (Eaσ, Eaτ, Eτm) are found to be comparable with each other which indicates that the ions have to overcome the same barrier while conducting as well as when relaxing.
The scaling of the modulus isotherms is shown in Fig. 8(b) where the frequency axis is scaled by ωmax and M′′(ω) is scaled by M′′max. The perfect overlap of all the curves on a single master curve at all temperatures indicates that the dynamical mechanisms are temperature independent.
This journal is © The Royal Society of Chemistry 2015 |