M. Struzika,
M. Malys*a,
M. Krynskia,
M. Wojcika,
J. R. Dygasa,
W. Wrobela,
F. Kroka and
I. Abrahams*b
aFaculty of Physics, Warsaw University of Technology, ul. Koszykowa 75, 00-662 Warszawa, Poland. E-mail: mmalys@wp.pl; Tel: +48 22 234 8216
bMaterials Research Institute, Department of Chemistry and Biochemistry, School of Biological and Chemical Sciences, Queen Mary University of London, Mile End Road, London E1 4NS, UK. E-mail: i.abrahams@qmul.ac.uk; Tel: +44 207 882 3235
First published on 24th September 2015
Structure and electrical conductivity in the oxide ion conducting compound Bi14YO22.5 have been investigated by powder X-ray and neutron diffraction, a.c. impedance spectroscopy, measurements of transference number and ab initio molecular dynamics (MD) simulations. Phase behaviour was studied using variable temperature X-ray diffraction and differential thermal analysis. The structure at room temperature is of the βIII tetragonal-type, details of which are discussed, including its relationship to both β and δ phases of Bi2O3. Bi14YO22.5 undergoes a reversible phase transition to a cubic δ-Bi2O3 type phase at high temperatures. This phase exhibits very high electrical conductivity, with transference numbers indicating that this conductivity is almost purely ionic in nature. MD simulations confirm 3-dimensional oxide ion transport in the βIII-phase.
At relatively low levels of substitution of bismuth (ca. 0.05 mol%), β or γ (sillenite) phases are often observed. Above this level of substitution, a variety of fluorite based phases can be obtained, depending on the dopant and level of substitution. Recently, there has been interest in phases with a Bi:
M ratio of 14
:
1 (6.67 mol% substitution), such as Bi14ReO24.5 (ref. 8 and 9) and Bi14MO24 (where M = W, Mo, Cr).10 Bi14ReO24.5 exhibits exceptionally high conductivity at intermediate temperatures, in the order of 10−3 S cm−1 at 300 °C.8,9
The (Bi1−xYx)2O3 system has been reviewed by Sammes et al.5 At bismuth rich compositions, around the 14:
1 Bi
:
Y ratio, a tetragonal phase is observed, which forms part of a solid solution range extending from ca. 5.8 to ca. 8.1%.11 However, it has been argued by Watanabe12 that the observed tetragonal solid solution is in fact metastable. Using an alternative phase diagram, he argued that the lowest x-value room temperature stable phase in this system is the end member of a solid solution ranging from x = 0.215 to x = 0.24, with compositions at lower x-values possessing only mixtures of this phase and α-Bi2O3 as the stable phases.
We have previously studied δ-phase compositions around x = 0.25 in the (Bi1−xYx)2O3 system, which show high levels of conductivity13,14 and revealed details of the vacancy ordering, using total neutron scattering analysis.15 Ab initio modelling of the x = 0.25 composition reveals that yttrium acts as a trap for mobile oxide ions, effectively lowering ionic conductivity compared to the parent δ-Bi2O3.16 In the present study we examine a more bismuth rich composition in the (Bi1−xYx)2O3 system at the 14:
1 Bi
:
Y ratio (x = 0.067), which exhibits a tetragonal structure at room temperature and a cubic fluorite phase at high temperatures.12 Here details of the structure and electrical conductivity of this composition have been studied in detail using X-ray and neutron powder diffraction, in combination with a.c. impedance spectroscopy, differential thermal analysis (DTA) and measurements of transference number. These results are compared with those from ab initio molecular dynamics calculations.
The ionic and electronic contributions to the total conductivity were measured using a modified EMF method, with an external adjustable voltage source in the concentration cell O2 (pO2 = 1.01 × 105 Pa):Pt|oxide|Pt:O2 (pO2 = 0.2095 × 105 Pa), as described in detail elsewhere.18 Measurements were performed on cooling between ca. 780 °C and 430 °C at stabilized temperatures. Samples were prepared as sintered pellets of 14 mm diameter and ca. 2 mm thickness. The Pt electrode had a diameter of 10.7 mm.
Powder neutron diffraction data were collected on the Polaris diffractometer at the ISIS Facility, Rutherford Appleton Laboratory. Data collected on back-scattering and low-angle detectors were used in subsequent refinements, covering the respective time of flight ranges 1.0 to 20 and 0.5 to 20 ms. The sample was contained in a cylindrical 11 mm diameter vanadium can located in front of the back-scattering detectors. Data were collected at room temperature for ca. 200 μA h.
Structure refinement was carried out by Rietveld whole profile fitting using the GSAS suite of programs.19 For the room temperature analysis, a combined X-ray and neutron approach was adopted. For the high temperature phase, a simple cubic fluorite model in space group Fmm was used for refinements, as previously described.13 For the low temperature structure, initial refinements were based on the structure of β-Bi2O3 (ref. 20 and 21) in space group P
21c (no. 114 (ref. 22)). Difference Fourier maps were used to locate interstitial oxide ion scattering density. It was found that the structure exhibited considerable positional disorder with respect to the structure of β-Bi2O3. The systematic absences for space group P
21c are in fact a subset of those for P42/nmc, and therefore a satisfactory refinement can be obtained in the lower symmetry space group, even when the true symmetry is higher. However, close inspection of the diffraction data reveal that the systematic absences are consistent with the higher symmetry space group, showing absence of Bragg peaks of the type hk0, when h + k = 2n + 1 (Fig. 1). We have previously reported the structure of βIII-Bi1.85Zr0.15O3.075 (ref. 23) in space group P42/nmc (no. 137, origin choice 2 (ref. 22)). This model is also consistent with that proposed for Bi7Y0.5O25,24 but exhibits greater positional disorder. Refinement of the βIII-model resulted in an excellent fit to the data for Bi14YO22.5 at room temperature and the structure presented here corresponds to this model. The extent of Y substitution in the present system is low and therefore Bi and Y atom parameters were tied to a single site. Anisotropic displacement parameters were refined for all atoms. A large U33 parameter on O(3) indicated that the site was split and in the final refinements this was placed on an 8g site. Oxide ion site occupancies were initially allowed to vary. The occupancies for O(1) and O(3) were found to be correlated with the amount of yttrium in the system and in the final refinements these were fixed according to stoichiometry. Crystal and refinement parameters for Bi14YO22.5 at room temperature are presented in Table 1.
![]() | ||
Fig. 1 Detail of fit to neutron low-angle data in space group P![]() |
a For definition of R-factors see ref. 19. | |
---|---|
Chemical formula | Bi14YO22.5 |
Formula weight | 3374.62 g mol−1 |
Crystal system | Tetragonal |
Space group | P42/nmc |
Unit cell dimension | a = 7.75622(3) Å, c = 5.63625(3) Å |
Volume | 339.071(3) Å3 |
Z | 0.533 |
Density (calculated) | 8.814 g cm−3 |
R-factorsa | (a) Neutron backscattering Rwp = 0.0111, Rp = 0.0230, Rex = 0.0062, RF2 = 0.0351 |
(b) Neutron low angle Rwp = 0.0281, Rp = 0.0325, Rex = 0.0220, RF2 = 0.0840 | |
(c) X-ray Rwp = 0.0405, Rp = 0.0251, Rex = 0.0170, RF2 = 0.2044 | |
(d) Overall Rwp = 0.0174, Rp = 0.0252, χ2 = 3.151 | |
Total no. of variables | 140 |
No. of profile points used | 3949 (neutron backscattering) |
5139 (neutron low angle) | |
2842 (X-ray) | |
No. of reflections | 1237 (neutron back-scattering) |
1023 (neutron low-angle) | |
232 (X-ray) |
To capture the oxide ion jump trajectories, high frequency oscillations of the oxide ions were excluded from calculations using a low pass Chebyshev filter. The cutoff frequency for the Chebyshev filter was determined based on the Fourier analysis of the ionic trajectories. For the frequencies above 5 × 1012 Hz, only thermal vibrations are observed. Energy distribution landscapes were derived from ionic density maps assuming a Boltzmann distribution.
Structural graphics were prepared using ORTEP-3 for Windows36 and VESTA.37
Atom | Wyc. | x | y | z | Occ. |
---|---|---|---|---|---|
Bi/Y | 8g | 0.25 | 0.00572(5) | 0.48660(6) | 0.933/0.067 |
O(1) | 16h | 0.5428(4) | 0.0590(4) | 0.2797(5) | 0.467 |
O(2) | 4d | 0.25 | 0.25 | 0.6496(1) | 1 |
O(3) | 4c | 0.25 | 0.75 | 0.3110(9) | 0.133 |
Anisotropic thermal parameters (Å2) | ||||||
---|---|---|---|---|---|---|
Atom | U11 | U22 | U33 | U12 | U13 | U23 |
Bi/Y | 0.0308(2) | 0.0102(2) | 0.0204(2) | 0 | 0 | −0.0067(1) |
O(1) | 0.0176(5) | 0.0405(11) | 0.0157(11) | 0.0052(5) | −0.0030(6) | 0.0166(7) |
O(2) | 0.0346(4) | 0.0124(4) | 0.0204(4) | 0 | 0 | 0 |
O(3) | 0.043(5) | 0.076(7) | 0.016(4) | 0 | 0 | 0 |
Bi/Y–O(1) | 2.138(4) × 2 | O(1)–Bi/Y–O(1) | 97.49(9) |
Bi/Y–O(1)′ | 2.250(4) × 2 | O(1)–Bi/Y–O(1)′ | 90.19(3) |
Bi/Y–O(2) | 2.1057(5) | O(1)–Bi/Y–O(2) | 86.70(7) |
Bi/Y–O(3) | 2.217(2) | O(1)–Bi/Y–O(3) | 93.72(10) |
Bi/Y⋯O(1)′′ | 2.842(2) × 2 | O(1)′–Bi/Y–O(1)′ | 82.33(9) |
Bi/Y⋯O(1)′′′ | 2.586(2) × 2 | O(1)′–Bi/Y–O(2) | 80.22(7) |
Bi/Y⋯O(2)′ | 2.6829(6) | O(1)′–Bi/Y–O(3) | 100.26(12) |
Bi/Y⋯O(3)′ | 2.598(3) | O(2)–Bi/Y–O(3) | 179.354(1) |
The structure of βIII-Bi14YO22.5 may be thought of as lying between those of β-Bi2O3 and δ-Bi2O3 (Fig. 4). The structure of β-Bi2O3,21 consists of a near cubic close packed array of bismuth cations, with O2− anions distributed in 3/4 of the tetrahedral cavities. The vacant tetrahedral cavities are ordered to yield channels perpendicular to the c-axis. The main difference between the structures of the β and δ phases of Bi2O3 lies in the vacancy ordering. In the latter case, vacancies are disordered over all the tetrahedral cavities on the crystallographic scale. Additionally, the cations in β-Bi2O3 are arranged in a slightly corrugated fashion above and below the x–y plane, while in the structure of δ-Bi2O3, the cations lie within the corresponding planes, as clearly evident in Fig. 3. In the case of the βIII-phase, the cation distribution shows the ideal planarity of the δ-phase, but the anion distribution more closely resembles that in the β-phase. The O(1) and O(2) sites in βIII-Bi14YO22.5 lie close to the corresponding sites in β-Bi2O3. However, the O(3) site in the βIII-phase corresponds to a site in the vacant channels of the β-phase and hence, in this respect, the anion distribution of the βIII-phase approaches that of the δ-phase. The O(1) site shows positional disorder compared to the corresponding position in β-Bi2O3. Since the shortest O(1)⋯O(1) contact distance is only 0.379 Å, a maximum occupancy of 0.5 is possible for this site. The initial refinements indicated that the occupancy of the O(1) site was significantly lower than this maximum value, with the total number of oxide ions on this site approximately equal to the amount of bismuth in the structure.
In βIII-Bi14YO22.5, the bismuth coordination is asymmetric, reflecting stereochemical activity of the Bi 6s2 lone pair of electrons, with three short Bi–O contacts of 2.11 to 2.25 Å and longer Bi–O contacts of 2.59 to 2.84 Å (the number of contacts takes into account the disorder on the O(1) site, which prevents simultaneous occupancy of neighbouring sites). These longer contacts may be considered to be non-bonding interactions, resulting in an essentially trigonal pyramidal coordination for Bi (Fig. 5a). The weighted average O–Bi–O angle within the trigonal pyramid is 85.7° for the configuration shown in Fig. 5a, which is again consistent with stereochemical activity of the Bi 6s2 lone pair of electrons. Based on geometrical considerations only, the alignment of the Bi 6s2 lone pair is such that it points towards the nominally empty channel (nominally empty, since O(3) is located in the channel), as previously described in the βIII-Bi1.85Zr0.15O3.075 system.23 The O(3) site occupancy initially refined close to double that of the yttrium occupancy of the Bi/Y 8g site and taking into account the multiplicities of these two sites this gave an approximate 1:
1 ratio for Y
:
O(3). The absence of atoms in a position corresponding to O(3) in the undoped system, β-Bi2O3 and the 1
:
1 Y
:
O(3) ratio, suggest that oxide ions on the O(3) site are exclusively associated with the Y3+ cation and allow it to adopt a more characteristic distorted octahedral coordination geometry involving O(1), O(2) and O(3) ions (Fig. 5b). The average Y–O bond length is 2.31 Å, which is slightly longer than the average value of 2.27 Å observed in Y2O3.39 Bond valence sums calculated using the parameters of Brese and O'Keeffe40 for Bi3+ and Y3+ were 3.08 and 3.03, respectively, for the proposed coordination geometries in Fig. 5.
The geometrical position of the bismuth lone pairs is critical in determining the nature of the ionic conduction mechanism in this compound. In order to examine details of the nature of the interactions between bismuth and the surrounding oxygen atoms, ab initio calculations of the valence electron distribution in βIII-Bi14YO22.5 were carried out. Fig. 6a shows an iso-surface plot of valence electron density of a representative bismuth atom from these calculations. The figure clearly shows significant polarization of electron density between bismuth and the three closest oxygen atoms, indicative of a significant degree of covalency in these interactions. The next nearest oxygen atoms show no obvious polarization of electron density in the direction of the bismuth atom, consistent with these interactions being much weaker in bonding character. The three bonded atoms are arranged in a trigonal pyramidal coordination with bismuth, with the lone pair on bismuth pointing out of the pyramid apex. The results therefore agree with the proposed directionality of the bismuth lone pair based on the geometrical considerations above. Fig. 7 shows the valence electron density around a empty channel in βIII-Bi14YO22.5 and confirms that the lone pairs on bismuth point into the channel as previously proposed in the structure of βIII-Bi1.85Zr0.15O3.075.23 An iso-surface plot of valence electron density around a representative yttrium atom is shown in Fig. 6b. The figure reveals six oxygen atoms in the yttrium coordination sphere, confirming the coordination derived from the crystallographic analysis. Interestingly, only one of the oxygen atoms shows significant polarization of electron density in the direction of the yttrium atom. This suggests that the majority of yttrium–oxygen contacts are ionic in nature.
![]() | (1) |
![]() | ||
Fig. 10 Thermal evolution of equivalent fluorite volume, Vf, in Bi14YO22.5 on heating (open symbols) and cooling (filled symbols). |
Fig. 11 shows the Arrhenius plot of total conductivity for the title compound on heating and cooling. Two linear regions are observed, one at low temperatures and a second of lower activation energy at high temperatures, with a significant jump in conductivity between the low and high temperature regions. There is thermal hysteresis associated with the transition, which occurs at around 650 °C on heating and ca. 550 °C on cooling, in good agreement with the results from DTA and X-ray measurements. The characteristic values of conductivity at 300 °C (σ300) and at 800 °C (σ800) are 1.06(1) × 10−5 S cm−1 and 1.83(7) S cm−1, respectively. The corresponding activation energies for the low (ΔELT) and high (ΔEHT) temperature regions are 0.96(1) eV and 0.64(4) eV, respectively. These values confirm very high conductivity for the cubic fluorite phase.
![]() | ||
Fig. 11 Arrhenius plot of total electrical conductivity for Bi14YO22.5 on successive heating (open symbols) and cooling (closed symbols) runs. |
Fig. 12 shows the thermal variation of ionic transference number for Bi14YO22.5 on cooling. The plot reflects the cubic to tetragonal phase transition seen in this system, with a transition between ca. 650 and 600 °C. The transition occurs at around the same temperature as seen in heating measurements using other techniques and reflects the fact that very long equilibration times are used in the transference number measurements. Both tetragonal and cubic phases show predominantly ionic conductivity over the measured temperature range, with the tetragonal phase exhibiting a transference number of around 0.8 and the cubic phase around 0.95 for oxide ion conduction.
We have previously proposed an oxide ion transport mechanism for the βIII-phase in the bismuth zirconate system,23 involving oxide ion hopping along the c-axis channels, facilitated by the positioning of the bismuth lone electron pairs. It is possible to use the results from the DFT calculations to derive a 3-dimensional map of energy distribution, which allows for the identification of energetically favourable pathways for oxide ion migration. An energy distribution landscape was derived from the ionic density map assuming a Boltzmann distribution of ion positions:16,42
![]() | (2) |
Energy distribution landscapes for vacant and filled channels in the structure of βIII-Bi14YO22.5 are shown in Fig. 13. The energy barriers at the saddle points for conduction along the channels are lower in the empty channels compared to those in the filled channels, with average values of 0.39 and 0.44 eV, respectively. Traces of oxide ion movements over the total simulation time (Fig. 14) show that oxide ion jumps occur not only through the c-axis channels (in the z-direction), but also between channels in the x–y plane. These results suggest that ion conduction in βIII-Bi14YO22.5 is in fact 3-dimensional in character and is similar to that in δ-Bi2O3. The level of ionic conductivity in the βIII-phase is relatively low compared to that in the δ-phase observed at high temperatures. This can be attributed to the reduced mobility of oxide ions at lower temperatures. Interestingly, at higher levels of substitution (x = 0.25) in the (Bi1−xYx)2O3 system, where the δ-phase is readily preserved to room temperature, low temperature values of ionic conductivity are quite similar (σ300 = 3.7 × 10−5 S cm−1 in Bi3YO6) to those in the present study, but a higher value of the low temperature activation energy (ΔELT = 1.19 eV)15 is observed in the δ-phase.
![]() | ||
Fig. 13 Energy distribution landscapes in βIII-Bi14YO22.5, showing sections through (a) an empty channel and (b) a filled channel. Scales are given in eV. |
![]() | ||
Fig. 14 Individual oxide ion trajectories (marked by different colours) projected onto (a) x–y and (b) x–z planes in βIII-Bi14YO22.5. |
This journal is © The Royal Society of Chemistry 2015 |