Josie E.
Auckett
*a,
Laura
Lopez-Odriozola
a,
Stewart J.
Clark
b and
Ivana Radosavljevic
Evans
*a
aDepartment of Chemistry, Durham University, Durham, DH1 3LE, UK. E-mail: ivana.radosavljevic@durham.ac.uk
bDepartment of Physics, Durham University, Durham, DH1 3LE, UK
First published on 18th January 2021
A number of metal oxides that crystallise in the scheelite structure type are known to be excellent oxide ion conductors. Here we report the synthesis of a series of materials with general formula LaNb1−xMoxO4+0.5x (x = 0, 0.08, 0.12, 0.16, 0.20) and excellent oxide-ionic conductivity for x ≥ 0.16 (7.0 × 10−3 S cm−1 at 800 °C). Bond valence energy landscape analysis showing possible facile oxide ion migration pathways give important insights into the local influence of defects on oxide-ionic conductivity in these phases. We also use variable-temperature powder X-ray diffraction data to present, for the first time for any scheelite-type material, a symmetry distortion mode refinement-based analysis of the phase transition between the scheelite and fergusonite structure types. This structural phase transition is known to have implications for both oxide-ionic conductive and ferroelastic properties. We demonstrate that one particular distortion mode, namely the Γ2+ displacive mode of the Nb atoms, is the most significant structural distortion leading to the symmetry-breaking phase transition from the tetragonal scheelite to the monoclinic fergusonite form of the material. Our diffraction data and ab initio lattice dynamics calculations provide evidence that the fergusonite–scheelite transition in these materials exhibits characteristics of a first-order transition.
Several promising oxide-ionic conducting (OIC) materials have been identified among the family of ABO4 oxides known as scheelites, including (Ca,AM)WO4−δ (AM = alkali metal),2,3 (Ca,La)MoO4+δ,4 (Pb,RE)WO4+δ (RE = rare earth element)5,6 and La(Nb,W)O4+δ.7 Of these, the PbWO4-derived materials have achieved the highest OIC performance (4 × 10−2 S cm−1 for Pb0.8La0.2WO4.1 at 800 °C),6 though their commercialisation is undesirable due to the toxic and environmental concerns associated with Pb-containing compounds.
The scheelite structure can be described as a derivative of the cubic fluorite structure, with rock-salt-type alternation of two cations of different size giving rise to a tetragonal network of edge-sharing AO8 and corner-sharing BO4 units. At room temperature, many scheelite-type phases adopt a monoclinic ferroelastic structural form known as fergusonite, which converts to the paraelastic scheelite phase upon heating.8–10 The symmetry relationship between the fergusonite (I2/a) and scheelite (I41/a) structures and the apparently continuous evolution of lattice parameters through the transition point have led most authors to consider the phase transition as being of second order.8–12 Differential scanning calorimetry (DSC) measurements of (Y,Dy,La)NbO4 (ref. 10) and studies of mode softening in the calculated phonon spectra of YTaO4 (ref. 11), as well as Raman spectroscopy experiments on pressure-induced scheelite–fergusonite transitions in YVO4 (ref. 13) and CaMoO4 (ref. 14), also support this view. However, a few authors have argued that the transition is in fact first-order with “quasi-continuous” character,15,16 citing fleeting phase coexistence observed near the phase transition in several studies of undoped and Ca-doped LaNbO4.10,12,15,17 Structural considerations, such as two very long (∼2.5 Å) Nb–O bonds in the fergusonite phase that increase to effectively non-bonding distance (∼3.1 Å) in the scheelite phase, have also been used to argue that the formation of the fergusonite phase is driven by optimisation of the Nb bonding environment,8 and that the transition is reconstructive.15
The fergusonite–scheelite phase transition is relevant to ionic conductivity in scheelite-based OIC materials because the conductivity of the two phases is predicted to differ, though discrepancies exist in the literature concerning which phase is more desirable. Packer et al.18 believed the tetragonal phase to be preferable for ionic conductivity, and Bi et al.19 reported that tetragonal (Ca,La)NbO4–δ exhibited lower activation energy for ionic diffusion than its monoclinic counterpart. However, oxygen tracer diffusion measurements showed a discontinuous decrease in the diffusion coefficient in CeNbO4+δ above the transition to scheelite near 850 °C.20 The common strategy of chemically doping scheelite-based OIC phases to enhance their conductivity often alters the temperature of the scheelite–fergusonite phase transition, which may further complicate understanding of the role of the dopant in influencing conductive performance.
One such doped system, LaNb1−xWxO4+δ (LNWO), has attracted recent interest due to its combination of low-toxicity elements, good redox stability and compatibility with candidate electrode materials such as NiO/YSZ (yttria-stabilised zirconia) and (La,Sr)MnO3.21,22 The substitution of 8–16% W6+ for Nb5+ improves the total conductivity of LaNbO4 by two orders of magnitude (to ∼10−3 S cm−1 at 800 °C) and also suppresses the significant proton conductivity observed in undoped LaNbO4 in humid atmospheres.7,21 The fergusonite–scheelite transition temperature is decreased by W6+ doping, with the x = 0.16 sample appearing almost tetragonal at room temperature.23 The presence of interstitial oxide ions in LNWO also leads to modulated ordering schemes24,25 similar to those observed in Ce4+-doped CeNbO4+δ.26
Oxide ion mobility is facilitated in LNWO via an interstitialcy mechanism that probably incorporates both interstitial–framework and interstitial–interstitial jumps.25,27 However, the simulations of Toyoura et al.27 predict a detrimental oxide-ion trapping effect in the vicinity of the W6+ dopant ions. It was suggested by those authors that doping instead with Mo6+ might mitigate this effect and further enhance the oxide-ionic transport properties. The synthesis of LaNb1−xMoxO4+δ (LNMO) phases with 0 ≤ x ≤ 0.2 was reported by Cao et al.,28 who found that 20% Mo6+ doping yielded total conductivity of ∼2.6 × 10−2 S cm−1 in air at 900 °C – three orders of magnitude higher than undoped LaNbO4. Electrochemical measurements performed under different atmospheres confirmed that the observed conductivity of LaNb0.8Mo0.2O4.1 in air is dominated by the ionic component (ion transport number tion > 0.95). LNMO therefore represents a highly promising system with superior performance characteristics for SOFC electrolyte applications in comparison to LaNbO4 and LNWO.
Promising conductivity and stable chemical characteristics in air make the LNMO series attractive for further investigation. We present here the results of our detailed variable-temperature structural studies, which are supported by ab initio lattice dynamics calculations to shed light on the nature of the fergusonite–scheelite phase transition, along with new insights into the conductivity-enhancing mechanism of Mo6+ doping in LaNbO4 obtained from bond valence energy landscape calculations.
Variable-temperature XRD (VT-XRD) data were collected using a Bruker D8 diffractometer with Cu Kα radiation. Powder samples were dispersed on an amorphous silica disc sample holder with Si powder added. The sample disc was positioned on a ceramic pedestal inside a HTK1200 furnace with Kapton windows, and XRD data were collected in the range 10° ≤ 2θ ≤ 70° for each scan, with a scan time of 70 min at each temperature. Measurements were taken over the range 50–1000 °C in 50 °C steps on both heating and cooling with a nominal temperature ramp rate of 30 °C min−1 between steps. A correction was applied to the sample temperature recorded at each step on the basis of a pre-determined furnace calibration curve.29
Rietveld refinements were performed using the Topas Academic software program.30,31 Initial models of the expected LaNbO4 tetragonal and monoclinic structures were sourced from the Inorganic Crystallographic Structure Database (ICSD). In all samples, the mixed Nb/Mo site was modelled as containing only Nb due to the lack of X-ray contrast between Nb5+ and Mo6+.
For the VT-XRD data, the sequences of XRD patterns collected during each heating–cooling cycle were analysed sequentially in a batch mode utilising macro functionality available in Topas, with parameters from each converged refinement copied as initial values for the next refinement in the sequence. Lattice parameters for Si at each experimental temperature were fixed at values obtained from a quadratic thermal expansion function fitted to the data in ref. 32. Basic peak shape parameters (pseudo-Voigt U, V, W and Y, and asymmetry due to axial divergence) were refined initially against the lowest-temperature LNMO00 pattern and subsequently fixed for all other temperatures and samples. An 8-term spherical harmonic function was then used to model peak shape variations as a function of temperature and composition. The monoclinic fergusonite structure model was converted to the non-standard setting I2/b (long axis c, for comparability with the standard scheelite setting I41/a) and used to fit the data collected over the whole temperature range, with the lattice parameters a, b, c and γ refined without constraints. Other refined parameters included scale factors for each phase, sample height, the Nb site z coordinate, and one isotropic atomic displacement parameter for each phase.
In addition to the sequential conventional Rietveld refinements against the VT-XRD data, the data collected on LaNbO4 (LNMO00) were also analysed using the symmetry-adapted distortion mode refinement approach.33,34 The Si internal standard was treated as part of the structural model in the conventional way. The monoclinic structure of LaNbO4 was described via eight freely refinable parameters, amplitudes of displacive distortion modes relative to the tetragonal scheelite-type parent structure, which were derived using ISODISTORT software.35
Sintered LNMO16 and LNMO20 discs not selected for impedance measurements were analysed by energy-dispersive X-ray spectroscopy (EDX) to confirm that Mo volatilisation had not significantly decreased the expected amount of Mo at the surface of the sample during the repeated sintering. These discs were sputter-coated with graphite (∼20 nm) and grounded using conductive silver paste to prevent surface charging under the electron beam. The EDX analysis was performed using a Hitachi SU70 scanning electron microscope equipped with an Oxford Instruments Aztec 3.3 microanalysis system, with an accelerating voltage of 15 kV. A Co metal standard was used to calibrate the beam intensity prior to recording EDX spectra.
Phonon calculations were performed using density functional perturbation theory.40 To address the hypothesis that the phase transition in LNMO was associated with temperature-dependent phonon softening, calculations were required that accurately described the lowest frequency modes, which are the ones with the largest relative error. Careful consideration of convergence parameters is required.41 We therefore tested phonon k-point convergence up to grid dimensions of 5 × 5 × 5, finding that a 3 × 3 × 3 grid sufficed. The importance of including the Γ point as part of this grid for accurate interpolation later should be noted. Additionally, in perturbative calculations, grids containing derivatives of quantities are required; given the nature of a plane-wave (Fourier) expansion, finer grids are required to accurately represent derivatives of potential and densities. We found that a grid 2.5 times finer than that for the wavefunction was required to describe the lowest-frequency modes with a small relative error. Finally, for phonon densities of states and dispersion curves, a Fourier interpolation onto a significantly finer grid (for density of states) and path in the Brillouin zone (for dispersion) was used.
An energy envelope around the fergusonite–scheelite phase transition was found by calculating energy as the unit cell was stepped linearly from one structure to the other. Atomic coordinates and lattice parameters were stepped simultaneously as follows: If x0 and y0 represent the initial and final parameter values and n is the number of chosen increments on the transition path, the path was defined by
Bond valence energy landscape (BVEL) maps with a grid spacing of 0.2 Å were calculated for O2− ions in LNMO using the 3DBVSMAPPER 2.02 software program.42 In order to investigate the local impact of dopant and interstitial defects on BVEL pathways in Mo6+-doped LaNbO4, 3 × 3 × 2 supercells of the tetragonal scheelite and monoclinic fergusonite structures were constructed in which two Nb5+ ions (separated from each other by > 10 Å in every direction) were replaced with Mo6+ ions. In order to conserve charge balance, one interstitial O2− ion was inserted at coordinates corresponding to a stable interstitial site identified by the DFT calculations of Toyoura et al.,27 resulting in the overall stoichiometry LaNb0.972Mo0.028O4.014 for the simulated material. Lattice parameters for these supercells were obtained by interpolating between values refined for LNMO00 and LNMO08 against XRD data at 323 and 705 °C for the monoclinic and tetragonal structures, respectively. Prior to the calculation of BVEL maps, the constructed superstructures were geometry-optimised without symmetry constraints using Castep with free energy and ionic force convergence criteria of 2 × 10−5 eV per atom and 5 × 10−2 eV Å−1, respectively.
All structures and density maps were visualised using VESTA.43
Unconstrained lattice parameters obtained for all samples by refining the fergusonite structure against VT-XRD data up to 950 °C are summarised in Fig. 1. Average linear thermal expansion coefficients for the Mo6+-doped samples are in the range 10.4–15.5 × 10−6 K−1, suggesting good thermomechanical compatibility with common electrode materials such as Ni-YSZ (10.3–14.1 × 10−6 K−1)45 and lanthanum strontium manganite (LSM, 11.2 × 10−6 K−1)46 (ESI Table S1†). It can be seen from Fig. 1(a and c) that the monoclinic structure of LNMO near room temperature becomes progressively less distorted with increasing Mo6+ content, with only the LNMO20 sample appearing to maintain the tetragonal structure at the lowest measured temperature (37 °C). However, the room-temperature XRD patterns in Fig. S1† show that pairs of peaks split by the monoclinic distortion appear merged in both LNMO16 and LNMO20, consistent with the findings of Cao et al.28 who also used visual inspection of XRD patterns to assign the tetragonal phase to their 15% and 20% Mo-doped samples.
Careful examination of the VT-XRD patterns revealed fleeting coexistence of the scheelite and fergusonite phases near the phase transition in several samples, as evidenced by the simultaneous appearance of the intense scheelite (112) reflection and the distinct (112) and (12) reflections of the fergusonite phase (Fig. 2). Although these “triple” peaks were clearly identifiable by eye, the small number of observable reflections from the minor phase were not intense or well-resolved enough in most cases to support simultaneous refinement of both phases during the sequential Rietveld analysis. Coexistence of the fergusonite and scheelite phases has been observed previously in studies of undoped and Ca-doped LaNbO4,10,12,15,17 and although some early authors speculatively attributed it to compositional or thermal inhomogeneity in their samples,12,17 its repeated observation across multiple studies suggests strongly that phase coexistence is an inherent feature of the fergusonite–scheelite transition in LaNbO4 and its derivatives.
Further examination of the data collected during repeated heating and cooling cycles also revealed a hysteretic behaviour whereby diffraction patterns nominally recorded at the same temperature near the phase transition sometimes showed different phases to be present, depending on the thermal history of the sample. This was particularly true for LNMO20, for which the refined cell appeared metrically tetragonal at the beginning of the heating cycle recorded in Fig. 1 but showed a measurable monoclinic distortion below 200 °C upon cooling (ESI Fig. S2†). It appears likely that the hysteresis effect is more pronounced in LNMO20 because the high-temperature phase is more readily trapped below its thermodynamically favourable range by slow transition kinetics when the transition temperature is low. The phase coexistence and hysteresis observed in LNMO are both suggestive of an energy barrier associated the fergusonite–scheelite transition, supporting the view that the transition is of first order in this material.
Consideration of the full VT-XRD data set allowed the LNMO phase diagram in Fig. 3 to be constructed. In all cases, phases were assigned up to the highest temperature at which they were observed in any run for a given sample upon either heating or cooling. The temperature of the fergusonite–scheelite phase transition, interpreted as occurring when the refined unit cell becomes metrically tetragonal (a = b and γ = 90°), is observed to decrease with increasing Mo6+ content. The loss of the structure modulation at high temperatures was determined visually by disappearance of the weak unindexed peak at 2θ ≈ 26.5°. This peak was especially weak in the LNMO20 phase, leading to some difficulty in determining the precise temperature of its disappearance. By comparison to the similar phase diagram published previously for LNWO,24 the fergusonite–scheelite transition occurs in LNMO at lower temperatures for equivalent dopant concentrations, while the modulated tetragonal phase persists to higher temperatures in LNMO and is also observed at all measured doping levels, compared to only ≥12% W6+ in LNWO.
Refining eight distortion modes allowed in the monoclinic LaNbO4 structure against the laboratory XRD pattern collected at room temperature gave an excellent fit with Rwp = 3.66% and RBragg = 2.80% (ESI Fig. S3†). The mode amplitudes obtained (ESI Fig. S4†) suggest that one of them, namely the a2 Γ2+ displacive mode of the Nb atoms, is the most significant structural distortion leading to the symmetry-breaking phase transition from the tetragonal scheelite to the monoclinic fergusonite form of LaNbO4. This is supported by the results of the Rietveld fit in which only this mode amplitude is refined and the remaining seven are fixed to zero, which gives only a small increase in the agreement indices (Rwp = 4.20% and RBragg = 3.28%, ESI Fig. S5†). The second- and third-most significant distortion modes are displacements of La and O atoms (Γ2+ modes labelled a1 and a7, with amplitudes of ∼25% and 50% that of a2, respectively). The atomic displacements that these modes represent are illustrated in Fig. 4.
The Γ2+ modes labelled a1 and a2 move the La and Nb atoms, respectively, along the scheelite c axis in such a way that the two types cations in the same layer in the structure (i.e. with the same z fractional coordinate) move in opposite directions, and in opposite directions to their counterparts in the layers above and below. This is represented by the blue and green arrows in Fig. 4. The Γ2+ mode labelled a7 causes in-plane displacements of the O atoms bonded to Nb (red arrows in Fig. 4), which result in a twisting of the NbO4 tetrahedra about the scheelite c axis. The main overall effect of these atomic displacements is the off-centring and partial flattening of the NbO4 tetrahedra. The bond lengths obtained from the distortion mode refinements against the room temperature XRD data suggest that Nb is best described as having a (4 + 2) coordination number, with three pairs of Nb–O bonds at 1.871(7), 1.924(8) and 2.662(7) Å.
VT-XRD data collected on LaNbO4 were also analysed using the distortion mode refinement approach. In these refinements, the amplitudes of the three Γ2+ modes representing displacements of the oxygen atoms were set to zero at temperatures above the phase transition (as they must be by definition), since the laboratory XRD data are not very sensitive to them, while the remaining five modes were freely refined. The unit cell parameters obtained agree closely with those from the conventional refinements, shown in Fig. 1. Freely refined Γ2+ distortion mode amplitudes representing the displacements of the Nb and La atoms in the monoclinic fergusonite structure relative to their positions in the high-symmetry tetragonal scheelite structure are shown in Fig. 5(a). Both decrease in magnitude with increasing temperature and reach zero at the phase transition temperature, as expected.
The Nb–O bond lengths extracted from distortion mode refinements are shown in Fig. 5(b). The bond length trends illustrate the change of the Nb (4 + 2) environment in fergusonite to a four-fold coordination in the scheelite structure (with the next two O nearest neighbours at ∼3.1 Å at 1000 °C). It is worth noting that the long Nb–O distance at 531 °C determined from our distortion mode refinements against XRD data, 2.97(1) Å, is in excellent agreement with 2.971(3) Å reported by David8 from neutron diffraction data at 530 °C, and the trend across the observed temperature range is in agreement with that reported by Arulnesan et al. from synchrotron XRD data.15 The evolution of the longest Nb–O distance with temperature corresponds to its contribution to the Nb bond valence sum decreasing from about 7% at room temperature to only 2% after the phase transition (Fig. 5(c)). Anomalous variations above the level of noise observed for some parameters in Fig. 5(b) and (c) in the temperature region 300–450 °C are attributed to an artefact of the sequential refinement as it approaches the transition temperature, where the small magnitude of the vanishing monoclinic distortion makes reliable determination of the distortion parameters more difficult.
Our phonon calculations confirmed the stability of tetragonal and monoclinic LaNbO4 in their respective temperature stability ranges above and below ∼500 °C, respectively (Fig. 6(a) and S6†). However, the dispersion curves calculated for tetragonal LaNbO4 at temperatures as low as 37 °C showed no imaginary frequencies to indicate instability with respect to phonon mode softening (Fig. 6(b) and S7†). This result indicates that the fergusonite–scheelite transition in LaNbO4 is not a phonon-driven second-order transition but is more likely to be first-order, as the dynamic stability of the scheelite phase below its experimentally-realised temperature range implies that its formation at lower temperatures is inhibited by the presence of an energy barrier.52 Transition path energy calculations were therefore performed to probe the possible existence of this barrier. In contrast to the results of similar calculations presented for YTaO4 by Feng et al.,11 which showed the Landau free energy decreasing monotonically from the scheelite phase to the fergusonite phase, our calculations found a shallow energy maximum between the two phases corresponding to an upper bound of 0.207 eV per unit cell for the transition energy.
![]() | ||
Fig. 7 Total conductivity of selected LNMO samples measured in air. Data collected upon heating and cooling are denoted by filled and open symbols, respectively. |
Data published previously for LNMO20 showed that its total conductivity is almost purely ionic with a very small (1–8%) contribution from proton conduction under oxidising conditions, though an order of magnitude increase in total conductivity under a 5% H2/N2 atmosphere was attributed to the onset of significant electronic conductivity caused by reduction of Mo6+.28 The conductivity profiles for our LNMO00 and LNMO20 samples are in excellent agreement with those published elsewhere.7,28 Doping with Mo6+ increases the total conductivity of LaNbO4 by at least 2 orders of magnitude at 600 °C and 3 orders of magnitude at 900 °C, with the conductivity of LNMO16 at 800 °C (7.0 × 10−3 S cm−1) being five times higher than that reported21 for the analogous W6+-doped phase. This conductivity is lower than that of La2Mo2O9, (ref. 53) Na0.5Bi0.5TiO3 (ref. 54), La1.54Sr0.46Ga3O7.27 (ref. 55) and the best Bi2O3-based oxide ion conductors56 and slightly lower than yttria-stabilised zirconia at the same temperature, and higher than the best performers in the brownmillerite57,58 structural family.
A decrease of the gradient in the Arrhenius plots, indicating a significant decrease in activation energy (Ea), is observed above ∼600 °C in LNMO00 and above ∼820 °C in LNMO16 and LNMO20. Activation energies determined from linear fitting to data measured above and below these temperatures are given in Table 1. A decrease in activation energy above the fergusonite–scheelite phase transition has been reported for LaNbO4 previously.19 Changes in total conductivity activation energies in scheelite-type phases are typically attributed to phase transitions or the onset of contributions from different charge carriers.21 However, both LNMO16 and LNMO20 adopt the tetragonal scheelite structure over the whole temperature range represented by the data in Fig. 7, as their corresponding structural phase transition temperatures lie below the range covered by our impedance spectroscopy measurements. One possible explanation for the change observed in LNMO can be found in Fig. 3, which shows that ∼800 °C marks the approximate temperature of the loss of modulated order in LNMO16 (as noted previously, the temperature of the same transition in LNMO20 may occur at a higher temperature than suggested by Fig. 3 due to the weakness of the superstructure reflection used to determine its presence). Although changes in activation energy upon the loss of modulated order have not been reported for LNWO, the phase diagram published by Li et al.24 indicates that such changes would only be expected in LNWO12 at ∼650 °C and in LNWO16 above 850 °C, and we note that these temperatures lie at or beyond the limits of temperature ranges spanned by conductivity data presented for these samples in the literature.21,25 Subtle changes to the local structure of LNMO, which would influence conductivity locally but cannot be determined from Bragg diffraction alone, present another possible explanation.
T c (°C) | E a below Tc (eV) | E a above Tc (eV) | |
---|---|---|---|
LNMO00 | 600 | 1.09(3) | 0.85(4) |
LNMO16 | 830 | 1.458(4) | 0.74(2) |
LNMO20 | 830 | 1.422(10) | 0.77(3) |
Sections of BVEL maps calculated for the LNMO03 supercell at 705 °C are presented in Fig. 8 and 9. As found in previous studies of LaNbO4, the low-energy O2− pathways involve the lattice ions (interstitialcy mechanism), and their connectivity permits net ionic transport along all three principal crystal axes. The calculated activation energy barrier (defined as the difference between the minimum energy value in the pathway and the energy value at the saddle point enabling continuous connectivity) is 0.82 eV, in agreement with the experimentally-determined value for undoped LaNbO4 (Table 1). However, Fig. 8 reveals an absence of low-energy pathways in the vicinity of a Mo6+ dopant cation, consistent with the higher size/charge ratio of these ions in comparison to Nb5+. “Trapping” of oxide ions near W6+ was suggested as a possible impediment to oxide ion mobility in LNWO,27 and our results suggest that similar “trapping” behaviour occurs in LNMO to some extent. Conversely, the relaxation of La3+ and Nb5+ cations near the O2− interstitial defect in the DFT-optimised structure appears to create more energetically-favourable O2− pathways in this region of the structure, especially within the a–b plane (Fig. 9).
Given that the interstitial concentration in synthesised samples is typically much higher than 2.8%, and that local distortions in these samples are likely to be influenced by multiple defects simultaneously, it is difficult to extrapolate the effect of interstitial doping on conduction pathways and their activation barriers under application-relevant conditions. Nevertheless, our results indicate strongly that the significant enhancement of ionic conductivity observed in LaNbO4 upon doping with M6+ cations is attributable to the presence of interstitial anions which accompany the dopant cations, rather than to the dopant cations themselves. This anion-induced enhancement probably occurs by two mechanisms: by increasing the total number of mobile charge carriers in the material, and also by distorting the local structure to increase the energetic favourability of the conduction pathway. These two benefits offset the local inhibitory effect of the M6+ dopant cations, despite the dopant ions being twice as numerous as interstitials. We note that both Mo6+ and W6+ dopants introduce the same number of interstitial O2− ions, yet the conductivity enhancement is observed to be greater in Mo6+-doped LaNbO4; this is perhaps related to a stronger trapping effect near the W6+ dopant ions (which our BVEL analysis does not quantify) or to the greater tendency for Mo6+ to tolerate a variety of coordination geometries and low coordination numbers than W6+,65 which has been identified as a factor contributing to the superior oxide-ionic conductivity of Mo-based analogues of other conductive oxides.66
A further consequence of the above finding is that the enhancement of oxide-ionic conductivity in Mo6+-doped LaNbO4 is likely to reach an upper limit when the concentration of oxide-trapping dopant cations becomes high enough to disrupt a significant proportion of conduction pathways in the material. The close similarity of the conductivity data recorded for LNMO16 and LNMO20 (Fig. 7) may indicate that such a limit is reached near these levels of Mo6+ doping.
For the monoclinic form of LNMO03 at 323 °C, BVEL maps show qualitatively similar behaviour of the conduction pathways near the dopant and oxide interstitial defects as in tetragonal LNMO03 discussed above. The calculated activation energy for ionic diffusion in the monoclinic phase is 0.90 eV, this value being higher than that determined for the high-temperature scheelite form but lower than the value obtained from the total conductivity Arrhenius plot for undoped LaNbO4 below 500 °C (Fig. 7 and Table 1). However, the BVEL surface drawn for monoclinic LNMO03 at the +0.90 eV isolevel reveals continuously connected diffusion pathways along the b axis only (Fig. 10), in contrast to the tetragonal scheelite phase, which exhibits full three-dimensional connectivity at or very near the 0.82 eV barrier in its BVEL map. Three-dimensional connectivity in the monoclinic LNMO03 map is evident only above the +1.31 eV level, suggesting a much higher degree of anisotropy in the ionic conduction pathways of the fergusonite structure. In a polycrystalline sample with random grain orientations, the experimentally-determined activation energy might be expected to lie between the values calculated for the one-dimensional and three-dimensional diffusion barriers; this appears to be consistent with the values reported in Table 1. These findings suggest an oxide-ionic conductive advantage for the scheelite over the fergusonite structure in samples without targeted engineering of grain orientation. Importantly, if suppression of the fergusonite–scheelite phase transition to lower temperatures leads to more facile oxide-ionic conductivity, the stabilisation of the scheelite phase of LaNbO4 by M6+ doping may represent an additional mechanism by which ionic conductivity is enhanced in this material by the inclusion of high-valence dopants.
![]() | ||
Fig. 10 BVEL maps calculated for the fergusonite phase of LNMO03 (yellow isolevel drawn at +0.90 eV), illustrating the contrasting connectivity of low-energy pathways in the a and b directions. The colours of spheres representing atomic species are as in Fig. 8 and 9. |
Symmetry distortion mode analysis of the XRD data was used successfully to describe the fergusonite–scheelite transition for the first time, yielding lattice parameter and interatomic distance trends in close agreement with conventional refinements against variable-temperature diffraction data presented here and elsewhere. The most significant distortion mode contributing to the phase transition involves displacement of Nb from the centre of the tetrahedron, accompanied by a change in coordination from 4 in the high-symmetry scheelite phase to (4 + 2) in the low-symmetry fergusonite phase. This change is further evidenced by a decrease in the contribution of the long O bonds to the calculated BVS of Nb before and after the transition.
Mo6+ doping levels of 16–20% greatly enhance the oxide-ionic conductivity of LaNbO4, confirming an earlier report.28 Comparison of BVEL maps calculated for a lightly-doped LNMO material above and below the fergusonite–scheelite phase transition indicate that the diffusion pathways for oxide ions are predominantly one-dimensional in the fergusonite phase but three-dimensional in the scheelite phase, so that the suppression of the fergusonite phase by Mo6+ doping should be advantageous for moderate-temperature ionic conductivity in the doped materials. In addition, considerable local enhancement of conductivity pathways is attributed to the presence of the charge-balancing O2− interstitials that accompany the dopant ions. In addition to increasing the number of charge carriers, these interstitials introduce local distortions to the cation lattice that appear to facilitate more favourable conduction pathways for the O2− ions. The competition between this effect and an inhibitory local “trapping” effect near the Mo6+ ions is expected to limit the enhancement of oxide-ionic conductivity due to this effect in LNMO above a certain doping threshold, where the more numerous Mo6+ defects would come to dominate the local behaviour of the material. Our conductivity data suggest that this limit is reached near the 16–20% Mo6+ concentration.
Footnote |
† Electronic supplementary information (ESI) available: Additional figures depicting X-ray diffraction data and refined structure parameters; calculated phonon dispersion curves; elemental analysis by EDX spectroscopy. See DOI: 10.1039/d0ta07453e |
This journal is © The Royal Society of Chemistry 2021 |