Next Article in Journal
Effect of an External Electric Field on the Kinetics of Dislocation-Free Growth of Tetragonal Hen Egg White Lysozyme Crystals
Next Article in Special Issue
Nb-Doped 0.8BaTiO3-0.2Bi(Mg0.5Ti0.5)O3 Ceramics with Stable Dielectric Properties at High Temperature
Previous Article in Journal
(Li1−xFex)OHFeSe Superconductors: Crystal Growth, Structure, and Electromagnetic Properties
Previous Article in Special Issue
Crystal Structures from Powder Diffraction: Principles, Difficulties and Progress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation into the Effect of Sulfate and Borate Incorporation on the Structure and Properties of SrFeO3-δ

School of Chemistry, University of Birmingham, Birmingham B15 2TT, UK
*
Author to whom correspondence should be addressed.
Crystals 2017, 7(6), 169; https://doi.org/10.3390/cryst7060169
Submission received: 28 April 2017 / Revised: 26 May 2017 / Accepted: 3 June 2017 / Published: 7 June 2017
(This article belongs to the Special Issue Crystal Structure of Electroceramics)

Abstract

:
In this paper, we demonstrate the successful incorporation of sulfate and borate into SrFeO3-δ, and characterise the effect on the structure and conductivity, with a view to possible utilisation as a cathode material in Solid Oxide Fuel Cells. The incorporation of low levels of sulfate/borate is sufficient to cause a change from a tetragonal to a cubic cell. Moreover, whereas heat treatment of undoped SrFeO3-δ under N2 leads to a transformation to brownmillerite Sr2Fe2O5 with oxygen vacancy ordering, the sulfate/borate-doped samples remain cubic under the same conditions. Thus, sulfate/borate doping appears to be successful in introducing oxide ion vacancy disorder in this system.

1. Introduction

Research into solid oxide fuel cells (SOFCs) as alternate energy materials has grown due to their high efficiency and consequent reduction in greenhouse gas emissions. Specifically, research into perovskite materials has attracted significant interest for potential SOFC materials including cathode, electrolyte and anode materials (see, for example, the review articles [1,2]). Traditionally, doping strategies for perovskite materials has involved the introduction of cations of a similar size e.g., Sr2+ for La3+, Mg2+ for Ga3+ to give the oxide ion conducting electrolyte La1−xSrxGa1−yMgyO3−x/2−y/2 [3]. More recently, we have applied oxyanion (silicate, phosphate, sulfate, borate) doping strategies to improve the properties of solid oxide fuel cell materials, initially demonstrating the successful incorporation into Ba2In2O5 [4]. Research on oxyanion incorporation into perovskites was initially reported for superconducting cuprate materials. This work showed that a wide range of oxyanions (carbonate, borate, nitrate, sulfate, and phosphate) could be incorporated into the perovskite structure [5,6,7,8,9,10,11,12]. In this doping strategy, the central ion of the oxyanion group is located on the perovskite B cation site, with the oxygens of this group occupying either three (borate, carbonate, nitrate) or four (sulfate, phosphate) of the available six anion positions around this site, albeit suitably displaced to achieve the correct coordination for the oxyanion group. As an extension to this work on cuprate superconductors, we have demonstrated successful oxyanion doping of perovskite materials for SOFC applications. This has included oxyanion (sulfate, silicate, phosphate) doping of Ba2(In/Sc)2O5 electrolyte materials [13,14,15,16]. Doping with these oxyanions introduces disorder on the oxygen sublattice, leading to improvements in the ionic conductivity. This work has also been extended to potential solid oxide fuel cell electrode materials, with successful oxyanion doping in the perovskite systems SrCoO3 [17,18,19,20], SrCo0.85Fe0.15O3 [21], SrMnO3 [18], CaMnO3 [22] and SrFeO3-δ [23], with the results suggesting improved performance/stability.
Following the prior work demonstrating the successful incorporation of silicate into SrFeO3-δ [23], we report in this paper an investigation into the effect of sulfate and borate doping. The undoped system SrFeO3-δ has attracted interest due to its high ionic and electronic conductivity, however under low p(O2) the oxide ion (and electronic conductivity) are significantly reduced. This is due to oxygen loss leading to a transformation to the oxygen vacancy ordered brownmillerite structure, Sr2Fe2O5 (Figure 1). We have therefore investigated the effect of borate/sulfate doping on this transformation.

2. Results and Discussion

2.1. SrFe1−xSxO3-δ

2.1.1. X-ray Diffraction Data

A range of samples of SrFe1−xSxO3-δ with increasing sulfate content (0 ≤ x ≤ 0.1) were prepared. X-ray diffraction analysis showed that without sulfate doping SrFeO3-δ forms a tetragonal perovskite in line with prior reports. This is illustrated in Figure 2 (expanded region 2 θ = 45° to 60°) where peak splitting can clearly be observed. Upon doping with sulfate, there is a transformation to a cubic cell (0.025 ≤ x ≤ 0.075), where no peak splitting is now observed (Figure 2). Above x = 0.075, small SrSO4 impurities appear, suggesting that the solubility limit of sulfate in SrFe1−xSxO3-δ is x ≈ 0.075.
In order to provide further support for the incorporation of sulfate, equivalent Fe-deficient samples were prepared without addition of sulfate. The X-ray diffraction data for SrFe0.95O3-δ (Fe-deficient, no sulfate) and SrFe0.95S0.05O3-δ are compared in Figure 3. These data show the presence of impurities for SrFe0.95O3-δ along with peak splitting, consistent with a tetragonal cell, as for the undoped SrFeO3-δ parent phase. In contrast, impurities are not observed for the sulfate-containing phase SrFe0.95S0.05O3-δ, and the cell is now cubic. Therefore, this comparison provides further evidence for the successful incorporation of sulfate.
Cell parameters for all these phases were determined using the Rietveld method (an example fit is shown in Figure 4). The variation of the cell parameters for SrFe1−xSxO3-δ with increasing sulfate content is given in Table 1. The data show a small general increase with increasing sulfate content, and this will be discussed in more detail in Section 2.1.4.
In addition to the determination of the unit cell parameters, site occupancies were refined for the Fe/S site. These occupancies are given in Table 1, and show that the refined values are in good agreement with the expected values.

2.1.2. Stability under N2

The effect of heating the SrFe1−xSxO3-δ samples under N2 was then examined. Upon heating under N2 to 950 °C, the X-ray diffraction data showed that SrFeO3-δ transforms to the oxygen vacancy ordered brownmillerite type Sr2Fe2O5 (Figure 1 and Figure 5). This would be expected to be unfavorable for fuel cell applications due to the ordering of oxygen vacancies, which is expected to lower the oxide ion conductivity. In contrast, for the sulfate-doped samples, the disordered cubic perovskite is retained under a similar heat treatment. For the x = 0.025 sample, there are some very weak peaks (see expanded XRD figure) associated with the brownmillerite structure, but for the higher sulfate contents (x ≥ 0.05), a single phase cubic cell is observed. In line with the reduction in the Fe oxidation state towards 3+, there is an increase in the cell parameters associated with the larger size of Fe3+ versus Fe4+ (Table 2).

2.1.3. Thermogravimetric Analysis

Thermogravimetric analysis (TGA) (heat treatment in N2 to reduce the Fe oxidation state to 3+) was then utilized to determine the oxygen contents of the samples as prepared in air. This analysis resulted in an interesting observation associated with Mass spectrometry analysis of the evolved gas. For the undoped sample SrFeO3-δ, the loss of mass is only associated with oxygen as indicated by the mass spectrometry data. However, for the sulfate-doped samples, a mass loss associated with CO2 was observed in addition to the expected mass loss due to O2.
Thus, the results indicated the presence of some carbonate in the sulfate-doped samples. In order to remove this carbonate, heat treatment in O2 (up to 900 °C) was carried out for these SrFe1−xSxO3-δ samples. After this oxygen treatment, TGA analysis indicated no presence of carbonate. This is illustrated in Figure 6, where a mass loss associated with CO2 is not observed after heat treatment in O2 for SrFe0.95S0.05O3-δ.
The TGA results indicating the presence of carbonate may at first glance suggest the existence of a small amount of SrCO3 impurity. However, the temperature at which the CO2 is lost is significantly lower than would be expected for SrCO3. In order to illustrate this, TGA data for SrCO3 were also collected and compared to the data for the SrFe0.95S0.05O3-δ. This experiment shows a significant difference in the temperature at which the loss of CO2 occurs for SrFe0.95S0.05O3-δ and SrCO3. In particular, the starting temperature for CO2 loss for SrFe0.95S0.05O3-δ occurs at a significantly lower temperature (≈490 °C vs. ≈760 °C for SrCO3), which may suggest that this carbonate is present in the perovskite structure, i.e., we have a mixed sulfate/carbonate-doped sample—SrFe1−x−ySxCyO3-δ. Further work is required to investigate this possibility, although as noted in the introduction, carbonate has been shown previously to be accommodated in perovskite materials.
Given the dual (O2 and CO2) mass loss for the air-synthesized SrFe1−xSxO3-δ samples, it is not possible to determine reliable oxygen contents for these samples. In addition, as detailed by Starkov et al., the determination of oxygen contents in partially substituted ferrites is non-trivial without a reliable fixed reference point [24]. Since it is possible that there may still be a small amount of Fe4+ in these samples after the N2 treatment, there is not a conclusive fixed reference oxygen point, and so any calculated oxygen contents would only be rough approximations.

2.1.4. Heat Treatment under O2

The samples (described above) heated under O2 were also examined by X-ray diffraction (Figure 7). All the sulfate-doped SrFe1−xSxO3-δ samples were shown to retain their original cubic structure while undoped SrFeO3-δ remained tetragonal.
Cell parameters and site occupancies were determined by Rietveld refinement using the X-ray diffraction data. From these structure refinements (Table 3), there is a decrease in the unit cell parameters upon heating in O2. This can be more clearly seen in Figure 8 where the variation in cell volume versus sulfate content is compared for samples heated in O2 and those just heated in air. These data show a reduction in cell volume on heating in oxygen, which can be correlated with a greater concentration of the smaller Fe4+ as a result of an increase in oxygen content. The cell parameter data also show an interesting variation with sulfate content, with an approximately linear increase in cell volume up to x = 0.05. The fact that there is no further increase for x = 0.075 may therefore suggest that the sulfate solubility limit is closer to x = 0.05.
The increase in cell volume with increasing sulfate content is at first glance unexpected given that S6+ is significantly smaller than Fe3+/4+. One possible explanation could relate to changes in the Fe3+/Fe4+ ratio on sulfate incorporation. However, this cell volume increase is also seen for the N2 treated samples (Table 2), where we only have Fe3+, and thus another factor must be significant. In this respect, the cell volume increase may be associated with the extra oxygen associated with this dopant. Thus, if we take the case of the N2 treated samples, we are effectively replacing Fe(III)O1.5 with SO3, which means the introduction of an extra 1.5 oxide ions per sulfate, which might be expected to contribute to an expanded cell size.

2.1.5. Conductivity Data

Following the successful incorporation of sulfate, the conductivities of the SrFe1−xSxO3-δ samples were examined in air. In general, the SrFe1−xSxO3-δ samples were found to have similar conductivities, with the exception of a notable decrease at lower temperatures for the higher sulfate content (x = 0.075) sample (Figure 9). This observation of a decrease in conductivity for higher dopant levels is comparable to the silicon-doped SrFeO3-δ system where it was proposed that at higher doping levels the silicate disrupts the Fe-O network resulting in a decrease in conductivity. [23]. Another factor, however, could relate to low levels of insulating impurities in this high sulfate content sample, given that the cell parameter data suggest that the sulfate solubility limit may be closer to x = 0.05.

2.2. SrFe1−xBxO3-δ

2.2.1. X-ray Diffraction Data

The possible incorporation of borate into SrFeO3-δ was then examined, with samples of SrFe1−xBxO3-δ prepared for 0 ≤ x ≤ 0.15. The results showed that a higher borate content was achievable compared to the sulfate-doped samples, with single phase borate-doped samples for x ≤ 0.1 (Figure 10). In terms of the effect of borate incorporation on the structure, similar results were observed as for the SrFe1−xSxO3-δ samples. In particular, borate doping resulted in a similar cubic cell as for the sulfate-doped samples.
Cell parameters and site occupancies were determined using the Rietveld method (an example fit is shown in Figure 11).
The refinements gave Fe/B occupancies in good agreement with those expected (Table 4). Furthermore, in this case, a decrease in the unit cell was observed on borate doping in agreement with the smaller size of B3+ compared to Fe3+/4+, and the fact that unlike the situation for sulfate doping there is no additional oxygen associated with the dopant (i.e., we are effectively replacing Fe(III)O1.5 with BO1.5).

2.2.2. Stability under N2

Heat treatment of the SrFe1−xBxO3-δ samples at 950 °C under N2 showed similar results to those observed on sulfate doping, with the borate-containing samples remaining cubic after this heat treatment in nitrogen (Figure 12). In line with the reduction of the Fe oxidation state to Fe3+, there was a shift in the peak positions to lower angles, indicating a larger unit cell.

2.2.3. Thermogravimetric Analysis

These samples were then analysed by TGA (heat treatment under N2). In contrast to the air synthesised SrFe1−xSxO3-δ samples, there was no mass loss due to CO2 observed in these SrFe1−xBxO3-δ systems, thus indicating no carbonate present.

2.2.4. Conductivity Data

For the SrFe1−xBxO3-δ samples (x = 0.05, 0.1), the conductivity data showed significantly lower conductivities at lower temperatures compared with the undoped system (Figure 13). However, above 600 °C, the conductivities were comparable with the x = 0.1 sample, in particular showing a small improvement in the conductivity compared with undoped SrFeO3-δ. Notably, these temperatures are in the range where operation as a cathode in a solid oxide fuel cell would be.
Conductivity values (at 700 °C) for the borate- and sulfate-doped samples are shown in Table 5, and compared to the equivalent data for silicate-doped SrFeO3-δ (from reference [23]). The data show similar values for all samples at this typical solid oxide fuel cell operating temperature.

3. Experimental

High-purity SrCO3, Fe2O3, (NH4)2SO4, and H3BO3 were used to prepare SrFe1−xS/BxO3-δ samples. Stoichiometric mixtures of the powders were intimately ground and initially heated to 900 °C (4 °C/min) for 12 h. Samples were then ballmilled (350 rpm for 1 h, Fritsch Pulverisette 7 planetary Mill) and reheated to 1000, 1050 and 1100 °C for 12 h with ballmilling of samples between heat treatments. For the SrFe1−xBxO3-δ (x = 0.05, 0.1) samples, a higher temperature (1200 °C) heat treatment was required to achieve single phase samples. In order to ensure maximum oxygen content, all samples underwent a final heat treatment at 350 °C for 12 h in air.
Additionally, portions of the SrFe1−xSxO3-δ (x = 0, 0.025, 0.05, 0.075) samples were heated to 900 °C under oxygen for 12 h with slow cooling at 50 °C /h to 350 °C, with the samples then maintained at this temperature for 12 h followed by cooling at 50 °C/h to room temperature. To test the stability under low p(O2) conditions, both sulfate and borate-doped samples were heated under N2 to 950 °C for 12 h.
Powder X-ray diffraction data were used in order to determine lattice parameters and phase purity. For SrFe1−xSxO3-δ samples heated in air, X-ray diffraction data were collected on a Panalytical Empyrean diffractometer equipped with a Pixcel 2D detector (Cu K α radiation). For the remaining SrFe1−xSxO3-δ and SrFe1−xBxO3-δ samples, a Bruker D8 diffractometer with Cu K α 1 radiation was used.
Samples were also analysed using thermogravimetric analysis (Netzch STA 449 F1 Jupiter Thermal Analyser with mass spectrometry attachment). Samples were heated to 950 °C in N2 (10 °C/min) and held for 30 min to reduce the iron oxidation state to Fe3+.
Pellets for conductivity measurements were prepared as follows: powders of SrFe1−xSxO3-δ and SrFe1−xBxO3-δ heated in air were initially ball milled (350 rpm for 1 h), before pressing into compacts and sintering at 1100 °C (SrFe1−xSxO3-δ) and 1200 °C (SrFe1−xBxO3-δ) for 12 h in air. Four Pt electrodes were attached with Pt paste and the samples were heated at 900 °C for 1 h in air. Samples were then furnace cooled to 350 °C and held at this temperature for 12 h to ensure full oxygenation. Conductivities were measured using the four probe dc method.

4. Conclusions

The results presented in this paper demonstrate the first reports of successful incorporation of sulfate and borate into SrFeO3-δ. This doping strategy results in a change from a tetragonal to a cubic cell, which is maintained even after heating under N2, where undoped SrFeO3-δ transforms to the oxygen vacancy ordered brownmillerite structure. Conductivity data in air show that the borate/sulfate-doped samples have comparable conductivities (Table 5) to undoped SrFeO3-δ at solid oxide fuel cell operating temperatures. Given these initial promising results, further studies are warranted to investigate the performance of these doped systems as solid oxide fuel cell cathodes.

Acknowledgments

We would like to thank EPSRC for funding (studentship for Abbey Jarvis). The raw datasets associated with the results shown in this paper are available from the University of Birmingham archive: http://epapers.bham.ac.uk/3011/.

Author Contributions

P.R.S. conceived and designed the experiments; A.J. performed the experiments; A.J. and P.R.S. analysed the data; A.J. and P.R.S. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Orera, A.; Slater, P.R. New Chemical Systems for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 675–690. [Google Scholar] [CrossRef]
  2. Jacobson, A.J. Materials for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 660–674. [Google Scholar] [CrossRef]
  3. Ishihara, T.; Matsuda, H.; Takita, Y. Doped LaGaO3 Perovskite Type Oxide as a New Oxide Ionic Conductor. J. Am. Chem. Soc. 1994, 116, 3801–3803. [Google Scholar] [CrossRef]
  4. Hancock, C.A.; Porras-Vazquez, J.M.; Keenan, P.J.; Slater, P.R. Oxyanions in Perovskites: From Superconductors to Solid Oxide Fuel Cells. Dalt. Trans. 2015, 44, 10559–10569. [Google Scholar] [CrossRef] [PubMed]
  5. Francesconi, M.G.; Greaves, C. Anion Substitutions and Insertions in Copper Oxide Superconductors. Supercond. Sci. Technol. 1997, 10, A29–A37. [Google Scholar] [CrossRef]
  6. Slater, P.R.; Greaves, C.; Slaski, M.; Muirhead, C.M. Copper Oxide Superconductors Containing Sulphate and Phosphate Groups. Phys. C Supercond. Appl. 1993, 208, 193–196. [Google Scholar] [CrossRef]
  7. Letouzé, F.; Martin, C.; Maignan, A.; Michel, C.; Hervieu, M.; Raveau, B. Stabilization of New Superconducting Thallium Cuprates and Oxycarbonates by Molybdenum. Phys. C Supercond. 1995, 254, 33–43. [Google Scholar] [CrossRef]
  8. Kinoshita, K.; Yamada, T. A New Copper Oxide Superconductor Containing Carbon. Nature 1992, 357, 313–315. [Google Scholar] [CrossRef]
  9. Huvé, M.; Michel, C.; Maignan, A.; Hervieu, M.; Martin, C.; Raveau, B. A 70 K Superconductor. The Oxycarbonate Tl0.5Pb0.5Sr4Cu2(CO3)O7. Phys. C Supercond. 1993, 205, 219–224. [Google Scholar] [CrossRef]
  10. Goutenoire, F.; Hervieu, M.; Maignan, A.; Michel, C.; Martin, C.; Raveau, B. A 62 K Superconductor with an Original Structure: Sr4−xBaxTlCu2CO3O7. Phys. C Supercond. 1993, 210, 359–366. [Google Scholar] [CrossRef]
  11. Von Schnering, H.G.; Walz, L.; Schwarz, M.; Becker, W.; Hartweg, M.; Popp, T.; Hettich, B.; Müller, P.; Kämpf, G. The Crystal Structures of the Superconducting Oxides Bi2(Sr1−xCaX)2CuO8−δ and Bi2(Sr1−yCay)3Cu2O10−δ with 0 ≤ x ≤ 0.3 and 0.16 ≤ y ≤ 0.33. Angew. Chem. Int. Ed. Engl. 1988, 27, 574–576. [Google Scholar] [CrossRef]
  12. Maignan, A.; Pelloquin, D.; Malo, S.; Michel, C.; Hervieu, M.; Raveau, B. Stabilisation of Three New Oxycarbonates by V and Cr Substitutions The Superconductors (Tl,M) 1Sr4Cu2(CO3)O7 (M Cr, V) and (Hg,V) 1Sr4Cu2(CO3)O6+z. Phys. C Supercond. 1995, 249, 220–233. [Google Scholar] [CrossRef]
  13. Shin, J.F.; Orera, A.; Apperley, D.C.; Slater, P.R. Oxyanion Doping Strategies to Enhance the Ionic Conductivity in Ba2In2O5. J. Mater. Chem. 2011, 21, 874–879. [Google Scholar] [CrossRef]
  14. Pérez-Coll, D.; Pérez-Flores, J.C.; Nasani, N.; Slater, P.R.; Fagg, D.P. Exploring the Mixed Transport Properties of sulfur(VI)-Doped Ba2In2O5 for Intermediate-Temperature Electrochemical Applications. J. Mater. Chem. A 2016, 4, 11069–11076. [Google Scholar] [CrossRef]
  15. Shin, J.F.; Hussey, L.; Orera, A.; Slater, P.R. Enhancement of the Conductivity of Ba2In2O5 through Phosphate Doping. Chem. Commun. 2010, 46, 4613–4615. [Google Scholar] [CrossRef] [PubMed]
  16. Shin, J.F.; Apperley, D.C.; Slater, P.R. Silicon Doping in Ba2In2O5: Example of a Beneficial Effect of Silicon Incorporation on Oxide Ion/proton Conductivity. Chem. Mater. 2010, 22, 5945–5948. [Google Scholar] [CrossRef]
  17. Hancock, C.A.; Slade, R.C. T.; Varcoe, J.R.; Slater, P.R. Synthesis, Structure and Conductivity of Sulfate and Phosphate Doped SrCoO3. J. Solid State Chem. 2011, 184, 2972–2977. [Google Scholar] [CrossRef]
  18. Hancock, C.A.; Slater, P.R. Synthesis of Silicon Doped SrMO3 (M = Mn, Co): Stabilization of the Cubic Perovskite and Enhancement in Conductivity. Dalton Trans. 2011, 40, 5599–5603. [Google Scholar] [CrossRef] [PubMed]
  19. Liu, Y.; Zhu, X.; Yang, W. Stability of Sulfate Doped SrCoO3-δ MIEC Membrane. J. Memb. Sci. 2016, 501, 53–59. [Google Scholar] [CrossRef]
  20. Zhu, Y.; Zhou, W.; Sunarso, J.; Zhong, Y.; Shao, Z. Phosphorus-Doped Perovskite Oxide as Highly Efficient Water Oxidation Electrocatalyst in Alkaline Solution. Adv. Funct. Mater. 2016, 26, 5862–5872. [Google Scholar] [CrossRef]
  21. Li, M.; Zhou, W.; Xu, X.; Zhu, Z. SrCo0.85Fe0.1P0.05O3-δ Perovskite as a Cathode for Intermediate-Temperature Solid Oxide Fuel Cells. J. Mater. Chem. A 2013, 1, 13632–13639. [Google Scholar] [CrossRef]
  22. Porras-Vazquez, J.M.; Kemp, T.F.; Hanna, J.V.; Slater, P.R. Synthesis and Characterisation of Oxyanion-Doped Manganites for Potential Application as SOFC Cathodes. J. Mater. Chem. 2012, 22, 8287–8293. [Google Scholar] [CrossRef]
  23. Porras-Vazquez, J.M.; Pike, T.; Hancock, C.A.; Marco, J.F.; Berry, F.J.; Slater, P.R. Investigation into the Effect of Si Doping on the Performance of SrFeO3-δ SOFC Electrode Materials. J. Mater. Chem. A 2013, 1, 11834–11841. [Google Scholar] [CrossRef]
  24. Starkov, I.; Bychkov, S.; Matvienko, A.; Nemudry, A. Oxygen Release Technique as a Method for the Determination Of “δ-pO2-T” diagrams for MIEC Oxides. Phys. Chem. Chem. Phys. 2014, 16, 5527–5535. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of Sr2Fe2O5 (showing oxygen vacancy ordering) (left) and SrFeO3 (right).
Figure 1. Structure of Sr2Fe2O5 (showing oxygen vacancy ordering) (left) and SrFeO3 (right).
Crystals 07 00169 g001
Figure 2. X-ray diffraction patterns for (a) SrFeO3-δ, (b) SrFe0.975S0.025O3-δ, (c) SrFe0.95S0.05O3-δ, (d) SrFe0.925S0.075O3-δ, (e) SrFe0.9S0.1O3-δ, and (f) SrFe0.85S0.15O3-δ. The data show a tetragonal cell for SrFeO3-δ, (as highlighted by the peak splitting: see expanded region Figure 2 θ = 45° to 60°), while the sulfate-doped samples are cubic. SrSO4 impurities (highlighted by an asterisk) are observed for samples with higher sulfate contents (x = 0.1 and 0.15).
Figure 2. X-ray diffraction patterns for (a) SrFeO3-δ, (b) SrFe0.975S0.025O3-δ, (c) SrFe0.95S0.05O3-δ, (d) SrFe0.925S0.075O3-δ, (e) SrFe0.9S0.1O3-δ, and (f) SrFe0.85S0.15O3-δ. The data show a tetragonal cell for SrFeO3-δ, (as highlighted by the peak splitting: see expanded region Figure 2 θ = 45° to 60°), while the sulfate-doped samples are cubic. SrSO4 impurities (highlighted by an asterisk) are observed for samples with higher sulfate contents (x = 0.1 and 0.15).
Crystals 07 00169 g002aCrystals 07 00169 g002b
Figure 3. X-ray diffraction patterns for (a) SrFe0.95O3-δ (Fe-deficient, no sulfate) and (b) SrFe0.95S0.05O3-δ. The data show impurities (highlighted by an asterisk) and a tetragonal cell for SrFe0.95O3-δ, while the sulfate-containing sample SrFe0.95S0.05O3-δ is phase pure and cubic.
Figure 3. X-ray diffraction patterns for (a) SrFe0.95O3-δ (Fe-deficient, no sulfate) and (b) SrFe0.95S0.05O3-δ. The data show impurities (highlighted by an asterisk) and a tetragonal cell for SrFe0.95O3-δ, while the sulfate-containing sample SrFe0.95S0.05O3-δ is phase pure and cubic.
Crystals 07 00169 g003
Figure 4. Observed, calculated and difference X-ray diffraction profile for SrFe0.975S0.025O3-δ.
Figure 4. Observed, calculated and difference X-ray diffraction profile for SrFe0.975S0.025O3-δ.
Crystals 07 00169 g004
Figure 5. X-ray diffraction patterns of (a) SrFeO3-δ, (b) SrFe0.975S0.025O3-δ, and (c) SrFe0.95S0.05O3-δ after heating under N2 to 950 °C.
Figure 5. X-ray diffraction patterns of (a) SrFeO3-δ, (b) SrFe0.975S0.025O3-δ, and (c) SrFe0.95S0.05O3-δ after heating under N2 to 950 °C.
Crystals 07 00169 g005
Figure 6. Plot of mass vs. temperature and ion current (for m/z = 44; CO2) vs. temperature (under N2) for SrFe0.95S0.05O3-δ prepared in air (black) and O2 (red), showing a mass loss associated with CO2 in the air synthesised sample, which is eliminated after heat treatment in O2. Solid lines indicate %mass and dashed lines indicate ion current.
Figure 6. Plot of mass vs. temperature and ion current (for m/z = 44; CO2) vs. temperature (under N2) for SrFe0.95S0.05O3-δ prepared in air (black) and O2 (red), showing a mass loss associated with CO2 in the air synthesised sample, which is eliminated after heat treatment in O2. Solid lines indicate %mass and dashed lines indicate ion current.
Crystals 07 00169 g006
Figure 7. X-ray diffraction patterns of (a) SrFeO3-δ, (b) SrFe0.975S0.025O3-δ, (c) SrFe0.95S0.05O3-δ, and (d) SrFe0.925S0.075O3-δ after heating under O2.
Figure 7. X-ray diffraction patterns of (a) SrFeO3-δ, (b) SrFe0.975S0.025O3-δ, (c) SrFe0.95S0.05O3-δ, and (d) SrFe0.925S0.075O3-δ after heating under O2.
Crystals 07 00169 g007
Figure 8. Plot of unit cell volume vs. x for SrFe1−xSxO3-δ heated in air (●) and heated in O2 (■).
Figure 8. Plot of unit cell volume vs. x for SrFe1−xSxO3-δ heated in air (●) and heated in O2 (■).
Crystals 07 00169 g008
Figure 9. Plot of log σ vs. 1000/T for SrFeO3-δ (●), SrFe0.975S0.025O3-δ (○), SrFe0.95S0.05O3-δ (■) and SrFe0.925S0.075O3-δ (□) in air.
Figure 9. Plot of log σ vs. 1000/T for SrFeO3-δ (●), SrFe0.975S0.025O3-δ (○), SrFe0.95S0.05O3-δ (■) and SrFe0.925S0.075O3-δ (□) in air.
Crystals 07 00169 g009
Figure 10. X-ray diffraction patterns for (a) SrFeO3-δ, (b) SrFe0.95B0.05O3-δ, (c) SrFe0.9B0.1O3-δ.
Figure 10. X-ray diffraction patterns for (a) SrFeO3-δ, (b) SrFe0.95B0.05O3-δ, (c) SrFe0.9B0.1O3-δ.
Crystals 07 00169 g010
Figure 11. Observed, calculated and difference X-ray diffraction profile for SrFe0.95B0.05O3-δ.
Figure 11. Observed, calculated and difference X-ray diffraction profile for SrFe0.95B0.05O3-δ.
Crystals 07 00169 g011
Figure 12. X-ray diffraction patterns for (a) SrFe0.95B0.05O3-δ, (b) SrFe0.95B0.05O3-δ after heating under N2, (c) SrFe0.9B0.1O3-δ, (d) SrFe0.9B0.1O3-δ after heating under N2.
Figure 12. X-ray diffraction patterns for (a) SrFe0.95B0.05O3-δ, (b) SrFe0.95B0.05O3-δ after heating under N2, (c) SrFe0.9B0.1O3-δ, (d) SrFe0.9B0.1O3-δ after heating under N2.
Crystals 07 00169 g012
Figure 13. Plot of log σ vs. 1000/T for SrFeO3-δ (●), SrFe0.95B0.05O3-δ (○), SrFe0.9B0.1O3-δ (■) in air.
Figure 13. Plot of log σ vs. 1000/T for SrFeO3-δ (●), SrFe0.95B0.05O3-δ (○), SrFe0.9B0.1O3-δ (■) in air.
Crystals 07 00169 g013
Table 1. Lattice parameters and Fe/S site occupancies obtained from Rietveld refinement using XRD data for SrFe1−xSxO3-δ. The structure of SrFeO3-δ was refined using a tetragonal space group (P4/mmm). The structures of the sulfate-doped samples were refined using a cubic space group (Pm 3 ¯ m).
Table 1. Lattice parameters and Fe/S site occupancies obtained from Rietveld refinement using XRD data for SrFe1−xSxO3-δ. The structure of SrFeO3-δ was refined using a tetragonal space group (P4/mmm). The structures of the sulfate-doped samples were refined using a cubic space group (Pm 3 ¯ m).
SrFe1−xSxO3-δ
S (x)00.0250.050.075
a (Å)3.8648(1)3.8723(1)3.8776(1)3.8766(1)
c (Å)3.8487(1)---
V (Å3)57.486(4)58.066(2)58.303(4)58.260(4)
Rwp (%)1.841.672.011.97
Rexp (%)0.920.920.900.90
Fe occupancy10.98(1)0.96(1)0.94(1)
S occupancy-0.02(1)0.04(1)0.06(1)
Fe/S Uiso0.003(1)0.009(1)0.011(1)0.008(1)
Table 2. Lattice parameters for SrFe1−xSxO3-δ after heating in air and N2.
Table 2. Lattice parameters for SrFe1−xSxO3-δ after heating in air and N2.
SrFe1−xSxO3-δ
S (x)0.0250.050.075
AirDry N2AirDry N2AirDry N2
a (Å)3.8723(1)3.9231(1)3.8776(1)3.9256(1)3.8766(1)3.9280(1)
c (Å)------
V (Å3)58.066(2)60.379(1)58.303(4)60.496(1)58.260(4)60.606(1)
Rwp (%)1.673.102.013.091.973.20
Rexp (%)0.922.590.902.510.902.50
Table 3. Lattice parameters and Fe/S site occupancies for SrFe1−xSxO3-δ, obtained from Rietveld refinement using XRD data for SrFe1−xSxO3-δ after heating in O2.
Table 3. Lattice parameters and Fe/S site occupancies for SrFe1−xSxO3-δ, obtained from Rietveld refinement using XRD data for SrFe1−xSxO3-δ after heating in O2.
SrFe1−xSxO3-δ Heated in O2
S (x)00.0250.050.075
a (Å)3.8651(1)3.8641(1)3.8692(1)3.8691(1)
c (Å)3.8477(1)---
V (Å3)57.349(3)57.694(2)57.924(2)57.922(3)
Rwp (%)4.163.293.793.93
Rexp (%)3.712.812.732.79
Fe occ1(-)0.97(2)0.94(2)0.93(2)
S occ-0.03(2)0.06(2)0.07(2)
Table 4. Lattice parameters and Fe/B site occupancies obtained from Rietveld analysis using X-ray diffraction data for SrFe1−xBxO3-δ. The structure of SrFeO3-δ was refined using a tetragonal space group (P4/mmm). The structures of the doped samples were refined using a cubic space group (Pm 3 ¯ m).
Table 4. Lattice parameters and Fe/B site occupancies obtained from Rietveld analysis using X-ray diffraction data for SrFe1−xBxO3-δ. The structure of SrFeO3-δ was refined using a tetragonal space group (P4/mmm). The structures of the doped samples were refined using a cubic space group (Pm 3 ¯ m).
SrFe1−xBxO3-δ
B (x)00.050.1
a (Å)3.8648(1)3.8593(1)3.8561(1)
c (Å)3.8487(1)--
V (Å3)57.486(4)57.483(2)57.336(4)
Rwp (%)1.843.233.67
Rexp (%)0.922.302.52
Fe occupancy10.92(1)0.89(1)
B occupancy-0.08(1)0.11(1)
Fe/B Uiso0.003(1)0.015(1)0.022(1)
Table 5. Conductivity data in air at 700 °C for SrFe1−xMxO3-δ where M = Si [23], S and B.
Table 5. Conductivity data in air at 700 °C for SrFe1−xMxO3-δ where M = Si [23], S and B.
Si (x)S (x)B (x)
00.050.10.1500.0250.050.0750.050.1
Conductivity 700 °C (S cm−1)26213518262530253032

Share and Cite

MDPI and ACS Style

Jarvis, A.; Slater, P.R. Investigation into the Effect of Sulfate and Borate Incorporation on the Structure and Properties of SrFeO3-δ. Crystals 2017, 7, 169. https://doi.org/10.3390/cryst7060169

AMA Style

Jarvis A, Slater PR. Investigation into the Effect of Sulfate and Borate Incorporation on the Structure and Properties of SrFeO3-δ. Crystals. 2017; 7(6):169. https://doi.org/10.3390/cryst7060169

Chicago/Turabian Style

Jarvis, Abbey, and Peter Raymond Slater. 2017. "Investigation into the Effect of Sulfate and Borate Incorporation on the Structure and Properties of SrFeO3-δ" Crystals 7, no. 6: 169. https://doi.org/10.3390/cryst7060169

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop