Gregory J.
Limburn
a,
Daniel W.
Davies
bc,
Neil
Langridge
a,
Zahida
Malik
a,
Benjamin A. D.
Williamson
d,
David O.
Scanlon
b and
Geoffrey
Hyett
*a
aSchool of Chemistry, University of Southampton, Southampton, SO17 1BJ, UK. E-mail: g.hyett@soton.ac.uk
bDepartment of Chemistry, University College London, 20 Gordon Street, London, WC1H 0AJ, UK
cResearch Computing Service, Information and Communication Technology, Imperial College London, London, SW7 2AZ, UK
dDepartment of Materials Science and Engineering, Norwegian University of Science and Technology (NTNU), Trondheim 7491, Norway
First published on 8th February 2022
Four novel compositions containing chalcogenide layers, adopting the Ba3M2O5M′2Ch2 layered structure have been identified: Ba3Sc2O5Cu2Se2, Ba3Y2O5Cu2S2, Ba3Sc2O5Ag2Se2 and Ba3In2O5Ag2Se2. A comprehensive comparison of experimental and computational results providing the crystallographic and electronic structure of the compounds under investigation has been conducted. Materials were synthesised at 800 °C under vacuum using a conventional ceramic synthesis route with combination of binary oxide and chalcogenide precursors. We report their structures determined by Rietveld refinement of X-ray powder diffraction patterns, and band gaps determined from optical measurements, which range from 1.44 eV to 3.04 eV. Through computational modelling we can also present detailed band structures and propose that, based on their predicted transport properties, Ba3Sc2O5Ag2Se2 has potential as a visible light photocatalyst and Ba3Sc2O5Cu2Se2 is of interest as a p-type transparent conductor. These four novel compounds are part of a larger series of sixteen compounds adopting the Ba3M2O5M′2Ch2 structure that we have considered, of which approximately half are stable and can be synthesized. Analysis of the compounds that cannot be synthesized from this group allows us to identify why compounds containing either M = La, or silver sulfide chalcogenide layers, cannot be formed in this structure type.
Specifically, in the 32522 structure, A3M2O5M′2X2, the ‘heavy’ anion layer [M′2X2]2− adopts the anti-litharge structure, while the oxide layer [A3M2O5]2+ adopts a triple layer fragment of the perovskite structure, but truncated such that the geometry of the M ion is pyramidal rather than octahedral. There is a competing and related structure type which can be formed using ions with the same formal charges, the 42622, with general formula A4M2O6M′2X2, sometimes expressed as the empirical formula AMO3M′X. In comparison to the 32522 structure, the 42622 has an additional rock-salt structured AO layer which shears and displaces the vertex sharing of the apical oxygen of the MO5 polyhedra. Schematic unit cells for the two structures are shown in Fig. 1. It has been previously identified that there is a delicate balance between the stability of the 42622 or 32522 structure for a given combination of ions.12 For example, Sr4Ga2O6Cu2S2 is stable, but the 32522 equivalent is not. Replacing gallium with scandium, it is found that both structures, Sr4Sc2O6Cu2S2 and Sr3Sc2O5Cu2S2, can be formed. With further substitution of strontium by barium it is found that only the 32522 structure is formed, Ba3Sc2O5Cu2S2. The overall trend is that moving from smaller ions (Sr2+, Ga3+) to larger ions (Ba2+, Sc3+) leads to a preference for the 32522 structure over the 42622 structure.
A survey of the literature indicates that there are a total of 10 compounds adopting the 32522 structure, where full structural details are available. The 42622 is slightly more numerous with 14 examples known. These mixed anion compounds have been found with a range of trivalent M ions in the [A4M2O6]2+ or [A3M2O5]2+ layers, including ions of chromium, gallium, vanadium, iron, manganese, scandium, and indium. The smaller Cr3+, Ga3+ and V3+ ions are only found with the 42622 structure, the larger Fe3+, Mn3+ and Sc3+ have been found in both the 32522 and 42622 types, while the largest, In3+, has only been found in the 32522 structure. A structure map plotting the average M–O bond length and the average A–O bond length is shown in Fig. 1, this indicates that for a given M–O bond length the 32522 structure has a larger A ion site that the 42622 structure. Therefore, the structure map indicates that novel compositions that might adopt the 32522 structure are more likely to be found in compositions containing larger A ions.
In this work we report on our attempt to identify new materials adopting the 32522 structure with optical and transport properties which will make them viable for p-type conducting or photocatalysis applications. In order to prevent the intra-band trapping states that would be detrimental to both these applications, this requires that the M ion must be a d0 and d10 trivalent cation.13 Our search of this space will include the Sc3+ ion, and the less studied larger ions, In3+, Y3+ and La3+ all combined with the largest practicable alkaline earth ion, Ba2+, which based on the analysis of the structure map should favour the formation of the 32522 structure. In a systematic review we will report our attempts to combine these four oxide layers; [Ba3Sc2O5]2+, [Ba3In2O5]2+, [Ba3Y2O5]2+, and [Ba3La2O5]2+, with the four heavy anion layers; [Cu2S2]2−, [Cu2Se2]2−, [Ag2S2]2− and [Ag2Se2]2−. This provides a matrix of 16 compounds in which to study the structural and property trends, by XRD, spectroscopy and computational modelling. Four of these sixteen are known, although one has not had a detailed structure published previously. We find that no lanthanum containing compound can be formed, nor any compound with a [Ag2S2]2− heavy layer, and we will discuss the factors behind the instability of these compositions. However, we can report Ba3Sc2O5Cu2Se2, Ba3Sc2O5Ag2Se2, Ba3In2O5Ag2Se2, and Ba3Y2O5Cu2Se2 as novel compounds.
BaO + 2BaCh + M2O3 + M′2O → Ba3M2O5M′2Ch2 | (1) |
Preliminary photocatalytic studies were conducted on a sample of Ba3Sc2O2Ag2Se2, using the dichloroindophenol (DCIP) dye degradation test.17 A co-catalyst of cobalt oxide was deposited on the oxyselenide powder using a method modified from the literature.18 In brief, 25 mg of Ba3Sc2O2Ag2Se2 was placed in sample vial to which was added 10 mL of a 4.24 mol dm−3 Co(NO3)2·6H2O (Sigma Aldrich, 98%) solution in acetone. The acetone was allowed to evaporate overnight, and the cobalt nitrate impregnated powder sample was transferred to an alumina boat which was placed in a tube furnace. The sample was heated to 700 °C under a flow of ammonia, and annealed for 1 hour. The sample was cooled under ammonia flow, and then heated to 200 °C for 1 hour under air exposure. This process initially reduced the cobalt nitrate to metallic cobalt, and then partially oxidised to the desired cobalt oxide co-catalyst, with an approximate loading of 2% cobalt by mass.
The CoOx@Ba3Sc2O2Ag2Se2 powder was assessed for photocatalytic activity by taking a 10 mg sample in a sample vial and adding 5 mL of a 6.44 × 10−5 M DCIP dye and 2.28 × 10−3 M glycerol solution. The test was conducted by placing a crown stirrer in the vial to disperse the powder, and exposing the sample vial to a 5 sun solar simulator (LS0104 150 W Xenon lamp) equipped with a UV filter. At 60 min intervals, for 180 min, an aliquot was taken out of the vial, and then centrifuged to separate the dye from the powder. The visible light absorption spectrum of the dye solution was collected in the range of 300 nm to 800 nm using a PerkinElmer lambda 750S spectrometer. The solution and powder were recombined, and the aliquot returned to the sample vial prior to continued light exposure. Changes in dye concentration caused by photocatalytic degradation were determined using the Beer–Lambert law relationship between absorption and concentration of the dye. Control tests were conducted using dye and glycerol solution under solar simulator illumination without the catalyst present, and the solution with and without catalyst stirred in the dark.
Room temperature conductivity measurements were carried out on a sample of Ba3Sc2O2Cu2Se2. The compound was pressed into a pellet using a 5 mm die under 2 tons of pressure. The pellet was then annealed in an alumina crucible in a vacuum sealed silica tube at 800 °C for 12 hours. Wire contacts were attached to the annealed pellet (1.5 mm thick, 5 mm diameter) of Ba3Sc2O2Cu2Se2 using silver epoxy, and four point resistance measurements collected using a Keithley DMM6500 multimeter.
Competing phases for all Ba3M2O5M′2Ch2 compounds were identified using the Materials Project API,24 limiting the search to those compounds that appear in the ICSD and resulting in a total of 172 compounds as listed in Table S1 (ESI†). These competing phases were relaxed using the same calculation parameters described above, with a denser k-point mesh of at least 100 Å3 used for metallic phases. All PBEsol calculations were carried out using the Fireworks python package25 and final values for energy above the convex hull of the compositional phase diagram (Ehull) were calculated using the Pymatgen python package.26 Bond distances and angles were extracted automatically using the CrystalNN algorithm.27
The screened hybrid exchange correlation functional, HSE06,28,29 has been shown in previous work to accurately calculate the electronic structure of this family of compounds,3 and to calculate accurate band gaps across a wide range of materials,30 so was used here to calculate band structures of the stable compounds. A plane-wave cutoff of 550 eV was used and a Γ-centred mesh with a k-point density of at least 70 Å3 was used to sample the Brillouin zone.
Band structures and optical absorption plots were produced using the Sumo python package.31 The optical absorption spectra were calculated using the real and imaginary parts of the dielectric constant calculated through a Kramers–Kronig transformation and a summation over the unoccupied bands, respectively, using a method by Adolph and co-workers.32 Within this formalism, intra-band and indirect absorptions are not accounted for and only direct valence to conduction band transitions are considered.33 Charge carrier effective masses were calculated using the curvature and electronic band extrema:
![]() | (2) |
Using this approach good fits to the powder diffraction patterns were found for the samples where the observed Bragg peaks could be assigned to the expected target phase, with wRp values of less than 4.66% for these four samples. The Rietveld refinement fits to the data are shown in Fig. 2. Ba3Sc2O5Cu2Se2 was found with 99.0% purity by mass, with the remaining Bragg peaks assigned to small amounts of BaSe and BaCO3. The sample of Ba3Sc2O5Ag2Se2 was 97.9% pure with only BaCO3 as an impunity; Ba3In2O5Ag2Se2 was found alongside 2.4% BaCO3 and 0.6% In2O3; finally, Ba3Y2O5Cu2Se2 had no identifiable crystalline impurity. A summary of the key refinement fit and lattice parameters for these four samples can be seen in Table 1.
Ba3Sc2O5Cu2Se2 | Ba3Sc2O5Ag2Se2 | Ba3In2O5Ag2Se2 | Ba3Y2O5Cu2Se2 | |
---|---|---|---|---|
Lattice parameter a/Å | 4.18127(7) | 4.21678(11) | 4.25908(8) | 4.3898(3) |
Lattice parameter c/Å | 27.7132(6) | 28.6608(10) | 28.8780(6) | 27.474(2) |
Volume/Å3 | 484.51(2) | 509.63(4) | 523.84(3) | 529.42(9) |
Data points | 4443 | 4443 | 4443 | 4443 |
Reflections (main phase) | 106 | 112 | 112 | 116 |
Parameters | 37 | 34 | 35 | 46 |
Purity | 99.0% | 97.9% | 97.0% | 100% |
wRp | 3.52 | 3.92 | 4.43 | 4.66 |
RF2 | 2.20 | 3.76 | 2.25 | 4.90 |
χ 2 | 1.80 | 1.64 | 1.68 | 1.41 |
Colour | Pale brown | Yellow-green | Black | Brown |
Ba1 (0.5, 0.5, z) | 0 | 0 | 0 | 0 |
Ba2 (0.5, 0.5, z) | 0.1424(1) | 0.1367(1) | 0.1384(1) | 0.1462(4) |
M; (0, 0, z) | Sc; z = 0.0718(3) | Sc; z = 0.0695(3) | In; z = 0.0710(1) | Y; z = 0.0766(6) |
O1 (0.5, 0, z) | 0.0814(5) | 0.0783(5) | 0.0804(4) | 0.089(2) |
O2 (0, 0, z) | 0 | 0 | 0 | 0 |
M′; (0.5, 0, z) | Cu; z = 0.25 | Ag; z = 0.25 | Ag; z = 0.25 | Cu; z = 0.25 |
Se (0, 0, z) | 0.198(1) | 0.1885(2) | 0.1892(1) | 0.2002(7) |
In addition to the four novel Ba3M2O5M′2Ch2 compounds discussed above, there are a further four that have been previously reported: Ba3Sc2O5Cu2S2, Ba3InO5Cu2S2, Ba3In2O5Cu2Se2, and Ba3Y2O5Ag2Se2, although for the final example only the lattice parameters have been provided, with no further structural details.3,4,9 These stable layered oxychalcogenides provide a suite of homologous materials with which to compare our novel compounds, in the discussion below. We did attempt to repeat the synthesis of Ba3Y2O5Ag2Se2 for this work but found the material to be extremely air sensitive and were unable to collect diffraction data of sufficient quality to refine the atomic positions in order to provide a detailed structural analysis, however we are able to confirm the reported lattice parameters (ESI,† Fig. S1). We found that the related yttrium containing copper selenide, Ba3Y2O5Cu2Se2, is also air sensitive, but the decomposition is much slower, which allowed us to collect diffraction data of sufficient quality for Rietveld refinement, as described above.
![]() | ||
Fig. 3 Plot of the lattice parameters for the I4/mmm structured Ba3M2O5M′2Ch2 phase plotted against the ionic radius of the M3+ ion present in each sample. Guidelines indicate and connect the samples with the same heavy anion layer. The a parameter is shown in the lower graph with squares, and the c lattice parameter in the middle graph with triangles./the top graph depicts the cell volume versus M ionic radii. Open symbols indicate previously reported materials, from ref. 3, 4 and 9, closed symbols represent novel materials reported in this paper. |
![]() | ||
Fig. 4 Left schematic diagram of the unit cell of the Ba3M2O5M′2Ch2; [(M = Sc, In, Y), (M′ = Cu, Ag) and (Ch = S, Se)]. Centre and right plots of the key angles, bond lengths and distances within the structure for the oxide and chalcogenide blocks. Values plotted as a function of the M ion in the oxide block, with chalcogenide layer distinguished by black squares for [Cu2S2]2−, red circles for [Cu2Se2]2−, and blue triangles for [Ag2Se2]2−. Filled symbols are for new materials reported here, empty symbols are from prior literature, ref. 3, 4 and 9. The lines shown connect samples with the same chalcogenide layer as a guide to the eye. Where they are not shown, the error bars are smaller than the symbols used. |
For a given chalcogenide layer the expansion in a lattice parameter with the replacement of Sc by In is of course driven by the increased size of the M3+ ion, which can be seen in the observed increase in the M–O bond length and especially the axial bond lengths, with an increase of 0.08 Å. This leads to the expansion of the oxide block height seen in Fig. 4 in the increase in the Ba–Ba distance. The response of the chalcogenide layer is only a limited change in the chalcogenide bond length, with the flexibly of the layer leading to an increase of the cis-bond angle instead, and a correspondingly reduced chalcogenide layer block height. For example, comparing Ba3Sc2O5Ag2Se2 to Ba3In2O5Ag2Se2 the silver selenide bond angle increases by 1° and the heavy layer block height decreases by 0.015 Å. Despite this decrease in the height of the chalcogenide layer with exchange of scandium for indium in each of the chalcogenide series, the much larger increase in the height of the oxide layer predominates and accounts for the observed increase in c lattice parameter.
These trends in bond length and angle changes in the oxide and chalcogenide blocks extend to Ba3Y2O5Cu2Se2, yet despite this we observe the anomalous decrease in the c lattice parameter. This can be rationalised as the decrease in the cell height comes from a significant drop in the interlayer spacing controlled by the square cuboid environment of the interlayer barium. The significant expansion in the a lattice parameter, when comparing Ba3In2O5Cu2Se2 to Ba3Y2O5Cu2Se2, caused by the much larger yttrium ion results in a flattening of the cuboid barium environment, as seen in the sharp decrease in the O–Ba–Ch angle from 79.3° to 74.6°, and a significant decrease in the interlayer spacing, as measured by the M–Ch distance which spans this environment and accounts for a 0.49 Å decrease across the unit cell. This can be thought of as the chalcogenide and oxide layers expanding in the basal direction around the inter-layer barium and towards each other in the 001 direction, as the barium ion is no longer the ideal size for the site when compared to the pairings with smaller ions.
The calculations also allowed the structures to be predicted, including the material for which we could not refine a structure, Ba3Y2O5Ag2Se2. There was a good match between the structures derived from Rietveld modelling of experimental data and the structures derived from computational methods, with the lattice parameters of the models matching experimental results within 1% for all samples. Comparing the calculated bonds and angles, the observed differences between the two models were: barium–barium distance, less than 0.4%, barium to anion bonds less than 2%, M–O bonds <1.5% and O–M–O angle <2% (M = Sc, In, Y). Previous work has shown that the PBEsol functional slightly overestimates the chalcogenide bond angle compared to experiment, and this is reflected again here with both copper and silver chalcogenide angle. The computational model over predicts the Ch–M′–Ch angle by up to 3.5%, and underpredicts the bond length by 3%.
Overall, the accurate match between the DFT calculations and the empirically derived structural models provides confidence that we can reliably interpret the model derived for Ba3Y2O5Ag2Se2 for which it was not possible to collect Rietveld quality data. For Ba3Y2O5Ag2Se2, the DFT structural model indicates a continuation of the structural trends observed in the rest of the series. The reduction in the length of c parameter compared to Ba3In2O5Ag2Se2, which we can observe experimentally, is due to a significant decrease in the inter-layer spacing, as the expansion in the a lattice parameter driven by the larger yttrium ion forces a flattening of the square cuboid environment of the inter layer barium. There is almost no expansion in the oxide block in the 001 direction despite the larger yttrium cation, and only the expected slight decrease in the chalcogenide block height.
![]() | ||
Fig. 5 Electronic band structures of the eight thermodynamically stable Ba3M2O5M′2Ch2 compounds calculated using the HSE06 DFT functional. Red and green dots mark the conduction band minimum (CBM) and valence band maximum (VBM), respectively. Compositions marked with an asterisk have been published previously.3,4 |
Band gap/eV | HSE06 optical band gap/eV | Light hole mass/me | VBM composition | CBM composition | |
---|---|---|---|---|---|
a Previously reported. b Uncertainty due to sample air sensitivity. Light hole effective masses for the newly reported compounds are broken down further by reciprocal space direction in Table S5 (ESI). | |||||
Ba3Sc2O5Cu2S2a | 3.27a | 3.24 | 0.6–0.64 | S 2p | Ba 5d, Cu 4s |
Ba3Y2O5Cu2Se2 | 3.04 | 0.46–0.5 | Se 3p | Ba 5d, Cu 4s | |
Ba3Sc2O5Cu2Se2 | 3.09 | 2.95 | 0.45 | Se 3p | Ba 5d, Cu 4s |
Ba3Y2O5Ag2Se2a | 2.63 | 0.46–0.48 | Se 3p | Ag 5s | |
Ba3Sc2O5Ag2Se2 | 2.58 | 2.60 | 0.44–0.56 | Se 3p | Ag 5s |
Ba3In2O5Cu2S2a | 1.77a | 1.84 | 0.54–0.56 | S 2p | In 5s |
Ba3In2O5Ag2Se2 | 1.44 | 1.58 | 0.37–0.56 | Se 3p | In 5s |
Ba3In2O5Cu2Se2a | 1.32a | 1.48 | 0.40–0.49 | Se 3p | In 5s |
The band gaps of the novel materials were experimentally determined from analysis of the diffuse reflection spectra, using the method recently outlined by Poeppelmeier, as suitable for crystalline, non-degenerate semiconductors.16 In brief, the Kubelka–Munk function was used to estimate the sample absorption from the reflectivity,37 and then a plot of the square of the absorption against the photon energy was used to extrapolate a tangent from the point of inflection to determine the band gap of the sample (Fig. 6). This was carried out for three of the novel materials, Ba3Sc2O5Cu2Se2 (band gap of 3.1 eV), Ba3Sc2O5Ag2Se2 (2.6 eV), Ba3In2O5Ag2Se2 (1.4 eV). These band gaps values are given in Table 2, alongside those previously reported for Ba3Sc2O5Cu2S2, Ba3In2O5Cu2S2 and Ba3In2O5Cu2Se2 and the values determined computationally from calculated optical absorption plots (Fig. S2, ESI†). For the yttrium containing samples experimental measurements could not be made due to their air sensitivity, so only the modelled band gaps are reported. Overall, there is good agreement between the experimental and calculated band gaps, as would be expected based on what we have shown previously for this family of materials.3
The trends in the band gaps observed for this series of compounds can be explained by considering the composition of the band edges derived from the electronic structure modelling. For all of the compounds, the valence band maximum (VBM) is comprised predominantly of the filled chalcogenide p orbitals, while the conduction band minimum (CBM) is predominantly composed of the empty coinage metal s states, except for the indium containing samples with the [Ba3In2O5]2+ block, where the In 5s states dominate then CBM. Ignoring the for the moment indium containing samples, this means that in the majority of the samples the size of the band gap is dependant only on the composition of the chalcogenide layer. We would expect sulfides to have larger band gaps than selenides, as the more electronegative sulfur lowers the VBM, while compounds with copper in the heavy layer would have a larger band gap than silver containing samples, as the more electronegative silver lowers the position of the CBM.
This explains why Ba3Sc2O5Cu2S2 has the largest band gap (3.24 eV, modelled value), due to the electronegative sulphur lowering the VBM, and the relatively electropositive copper raising the CBM. As both Ba3Y2O5Cu2Se2 and Ba3S2O5Cu2Se2 share the copper selenide heavy layer they have similar modelled band gaps (3.04 eV and 2.95 eV), due the VBM and CBM being composed of the same orbitals, but smaller than Ba3Sc2O5Cu2S2 due to the contribution of the more electropositive selenium to the VBM. The silver selenides; Ba3Y2O5Ag2Se2 and Ba3S2O5Ag2Se2 both have smaller band gaps of 2.63 eV and 2.60 eV, due to presence of the lower lying silver 5s states. The indium containing compounds have much smaller band gaps, as indium 5s states are much lower in energy than either copper or silver ns states giving band gap energies of 1.84 eV, 1.58 eV and 1.48 eV for the copper sulfide, silver selenide and copper selenide, respectively. The small differences observed in the modelled band gaps for the pairs Ba3Sc2O5Cu2Se2 and Ba3Y2O5Cu2Se2 (2.95 eV to 3.04 eV); Ba3Sc2O5Cu2Se2 and Ba3Y2O5Cu2Se2 (2.60 eV to 2.63 eV); and Ba3In2O5Cu2Se2 and Ba3In2O5Ag2Se2 (1.48 eV to 1.58 eV), despite each pair having identical band edge compositions, is due to a secondary effect where an increase in the basal lattice parameter reduces overlap and decreases the dispersion in the bands leading to the slight increase in band gap.8
From the novel compounds reported in this work we can conclude from the electronic structure data that Ba3Sc2O5Cu2Se2 is of most interest as a possible p-type transparent conductor. It has a calculated hole mobility equivalent to Sr3Sc2O5Cu2S2, and a sufficient large band gap to prevent visible light absorption and is air stable. We can also conclude that Ba3Sc2O5Ag2Se2 might be a functional visible light photocatalyst, based on its reasonable transport properties and 2.6 eV band gap.
The visible light photocatalytic potential of Ba3Sc2O5Ag2Se2 was assessed using a dye degradation test, where a solution of dichloroindophenol (DCIP) dye undergoes reductive decolouration in the presence of glycerol as a sacrificial oxidant, enhanced by the photocatalyst. This is an effective and simple test to assess visible light photocatalysts.17 The sample of Ba3Sc2O5Ag2Se2 was loaded with a cobalt oxide co-catalyst to enhance its activity,18 and allow for measurements to be conducted in a 3 hour time window. A 5 sun solar simulator with a UV light filter was used as the light source. This found that 10 mg of CoOx@Ba3Sc2O5Ag2Se2 in 5 mL of 6.44 × 10−5 M dye solution was able to degrade 15.1% of the dye during the three hour experiment. No degradation of the dye was observed without the catalyst present, or in the presence of the catalyst without the light source. This confirms that Ba3Sc2O5Ag2Se2 is a visible light photocatalyst, comparable to other layered oxychalcogenides tested using the same procedure, with a similar rate to Ba3In2O5Cu2S2 (14.5% dye degradation in three hours), but lower than Ba3Sc2O5Cu2S2 (27.0% degradation).4
Additionally, room temperature conductivity measurements were conducted on a pellet of Ba3Sc2O5Cu2Se2, which was found to have a density of 73% relative to the theoretical density derived from diffraction data. Four point probe measurements found that the pellet had a conductivity of 5.0(1) × 10−5 S cm−1, which is comparable to previous reports on other promising for pristine undoped layered oxychalcogenides such as LaOCuS (1 × 10−4 S cm−1),10 and Sr2GaO3CuS (2.2 × 10−4 S cm−1),38 although not as high as values reported for LaOCuSe (24 S cm−1) or Sr3Sc2O5Cu2S2 (2.8 S cm−1).2,39 However, it should also be considered that the value reported here for Ba3Sc2O5Cu2Se2 are for an annealed pellet, rather than a plasma sintered one.
We have not been able to form any compound containing lanthanum, i.e. those with a [Ba3La2O5]2+ layer. All Bragg peaks observed in the diffraction patterns of our attempts at these samples could be indexed to elemental or binary by-products, as summarised in Table S4 (ESI†), with no indication of the expected peaks of the target phases. These syntheses were speculative, as there are no prior examples of lanthanum in the 5 coordinate M3+ ion environment in the 32522-structure type. The lanthanum ion, with an ionic radius of 1.03 Å, is considerably larger than the yttrium ion (0.90 Å) which is the largest ion found on the M site in the known examples of the structure type. We hypothesise that the limiting size for ions in the oxide block can be predicted based on the Goldschmidt tolerance factor, t, originally derived for predicting the formability of the perovskite structure, using the relationship of the relative ion sizes rA, rB and rO for the 12 coordinate cation site, the 6 coordinate cation site and the anion site, as shown in eqn (3). Although not always a guarantee of stability, it is successful as a filter for unstable compositions in the perovskites.40 A survey of all of the known 32522 structure types (and related 42622 structure) finds that these all have oxide layers with tolerance factors in the range of 0.9 to 1.0, which matches that observed for undistorted, cubic perovskites, implying that these oxychalcogenide structure types have a relative tight tolerance for the size of the ions. Calculating the tolerance factor for [Ba3La2O5]2+ give a value of 0.88, which falls outside of this stable range, and indicates that the barium ion is too small for the lanthanum ion. The tolerance factors for the oxide layers can be found in Table 3. Perovskite type compounds can be made with tolerance factors less than 0.9, but these undergo distortion through polyhedral tilting, to reduce the size of the 12-coordinate A site. It is likely that such distortions are not possible in the 32522 structure due to the need for commensurate bonding to the chalcogenide layer.
![]() | (3) |
Oxide layer | Tolerance factor, t |
---|---|
[Ba3Sc2O5]2+ | 0.99 |
[Ba3In2O5]2+ | 0.97 |
[Ba3Y2O5]2+ | 0.93 |
[Ba3La2O5]2+ | 0.88 |
We were also unable to experimentally synthesise any compound containing a silver sulfide layer, nor were any predicted to be stable from computational methods. At first this is surprising, given the ability to synthesis examples of all the other copper and silver chalcogenide layers. However, in each of the attempts at forming a silver selenide, the same by-products were observed, elemental silver and barium sulphate. Our hypothesis for this instability is that silver is too readily reduced by sulfide, with BaSO4 acting as a thermodynamic sink. In contrast the silver selenides are stable because of the lower reduction potential of Se2− compared to S2−, and the consequent lower enthalpy of formation of BaSeO4. This also helps to explain the lack of prior examples of layered silver sulfides in the literature, where to our knowledge there are only two examples, these are LaOAgS,41 and Sr3Fe2.5O5Ag1.5S2; in the former there is no alkaline earth to form a sulphate and in the latter the composition is stabilised by partial substitution of silver for iron in the sulfide layer.42
The final compound that could not be synthesized was Ba3Y2O5Cu2S2. In our attempt at the synthesis of this we could not identify any of the expected peaks of the target phase in the diffraction pattern collected on the sample, which instead was found to contain Bragg peaks which could indexed to a mixture of Y2O3, BaS, Cu, BaSO4 and BaYO4. This inability to combine the [Ba3Y2O5]2+ layer with the [Cu2S2]2− layer can be rationalised based on the expected size. Yttrium is the largest of the three ions considered in the oxide block, while the [Cu2S2]2− layer contains the smallest paring of ions for the chalcogenide layer, so it is likely that the [Ba3Y2O5]2+ is simply too large to be accommodated in the structure with the copper sulfide layers. Although predicted to be unstable, it was still possible to computationally model and produce a relaxed structure for Ba3Y2O5Cu2S2, which had predicted lattice parameters of 4.34 Å and 26.48 Å. The largest previously reported a lattice parameter for a layered oxychalcogenide containing copper sulfide layers is the value of 4.18 Å in Ba3In2O5Cu2S2.4 In order to achieve this theoretical basal expansion to 4.34 Å the model requires an extreme stretching of the copper sulfide geometry, such that the S–Cu–S angle becomes 125°, even accounting for the tendency to over-predict the angle value, this is an extreme distortion from the ideal tetrahedral geometry. This would support the argument of size incompatibility as the reason for the inability to synthesis Ba3Y2O5Cu2S2, and also suggest the possible upper size limit for inclusion of the copper sulfide layer is approximately 4.20 Å and certainly no larger than 4.35 Å.
Footnote |
† Electronic supplementary information (ESI) available: Fig. S1: Rietveld refinement of attempt at forming Ba3Y2O5Ag2Se2. Fig. S2: Calculated optical absorption data. Table S1: List of competing phases used in calculating energy above hull. Table S2: Selected band distances and angles from refined structural models. Table S3: Structural details from computational models. Table S4: Details of refinement of compounds which could not be synthesized. Table S5: Light hole masses for newly reported compounds. All data supporting this study are openly available from the University of Southampton repository at DOI: http://10.5258/SOTON/D2122. See DOI: 10.1039/d1tc05051f |
This journal is © The Royal Society of Chemistry 2022 |