Geletu
Qing
a,
David
Thompson
a,
Mourad
Benamara
b,
Clemens
Heske
cd,
Lauren
Greenlee
*e and
Jingyi
Chen
*a
aDepartment of Chemistry and Biochemistry, University of Arkansas, Fayetteville, USA. E-mail: chenj@uark.edu
bInstitute of Nanoscience and Engineering, University of Arkansas, Fayetteville, USA
cDepartment of Chemistry and Biochemistry, University of Nevada, Las Vegas (UNLV), USA
dInstitute for Photon Science and Synchrotron Radiation (IPS) and Institute for Chemical Technology and Polymer Chemistry (ITCP), Karlsruhe Institute of Technology (KIT), Karlsruhe, Germany
eDepartment of Chemical Engineering, Pennsylvania State University, USA. E-mail: greenlee@psu.edu
First published on 8th September 2022
Oxygen vacancies in metal oxides can determine their properties. However, it is difficult to reduce the oxygen vacancy concentration in metal oxides without annealing them under high pressure. In this work, we develop a facile approach to control oxygen vacancy content via an ozone treatment under ambient pressure during cooling. This approach is demonstrated for the synthesis of La1−xSrxFeO3−δ (0 ≤ x ≤ 1, 0 ≤ δ ≤ 0.5x) perovskite oxides – an important class of energy-related materials due to their wide range of non-stoichiometry, mixed ionic and electronic conductivity, and the presence of a rare Fe(IV) oxidation state. A series of La1−xSrxFeO3−δ compounds was initially synthesized using a polymerized complex method. The concentration of oxygen vacancies and Fe(IV) were determined by redox titration, and the crystal structures were derived by analyzing X-ray diffraction patterns using Rietveld refinement. Significant amounts of oxygen vacancies were found in the as-synthesized compounds with x ≥ 0.8: La0.2Sr0.8FeO3−δ (δ = 0.066) and SrFeO3−δ (δ = 0.195). The ambient-pressure ozone treatment approach was able to substantially reduce the amount of oxygen vacancies in these compounds to achieve levels near the oxygen stoichiometry of 3 for La0.2Sr0.8FeO3−δ (δ = 0.006) and SrFeO3−δ (δ = 0.021). The oxygenation/deoxygenation kinetics can be tuned by the cooling rate after annealing. As the oxygen vacancy concentration decreases, the structure of SrFeO3−δ evolves from orthorhombic to cubic, demonstrating that the crystal structures in metal oxides can be highly sensitive to the number of oxygen vacancies. The ozone treatment approach developed in this study may thus offer a robust means to tune the properties of a wide variety of metal oxides.
The La1−xSrxFeO3−δ perovskites are also known to be one of the few compound suites that contain Fe(IV), where the substitution of La(III) with Sr(II) results in the oxidation of Fe(III) to Fe(IV); however, oxygen vacancies would also tend to compensate for the decrease in positive electric charge.19–21 The formation of oxygen vacancies is particularly common in compounds that theoretically contain a large amount of Fe(IV). For example, a full substitution of La(III) with Sr(II) to yield SrFeO3−δ is expected to contain only Fe(IV) if it is fully oxygenated (i.e., SrFeO3), but in practice, its oxygen deficiency can vary from δ = 0 to δ = 0.5, depending on the synthesis conditions.22–24 The oxygen stoichiometry of the La1−xSrxFeO3−δ perovskites strongly depends on the synthesis conditions, especially the final annealing temperature and oxygen partial pressure. With the substitution of La(III) with Sr(II), the formation energy of oxygen vacancies in La1−xSrxFeO3−δ decreases significantly and, as a result, the compound tends to be oxygen deficient.25 For example, the La1−xSrxFeO3−δ compounds with x ≤ 2/3 are essentially free of oxygen vacancies (i.e., δ = 0) when they are annealed in air and then cooled down slowly. However, for the compounds with x > 2/3, it is impossible to remove the oxygen vacancies by annealing in air, and the oxygen vacancy concentration increases with the Sr doping level.20,26
Being able to tune or reduce the number of oxygen vacancies in the La1−xSrxFeO3−δ perovskites has drawn considerable attention, because they exhibit a mixed ionic and electronic conductivity,1,27,28 suited for electrocatalysis29,30 and oxygen storage,31 and Fe(IV) is particularly critical for understanding the oxygen evolution reaction (OER) mechanisms in Fe-doped NiOOH electrocatalysts.29 The oxygen deficiency of La1−xSrxFeO3−δ could be finely tuned in the range of δ = 0 to δ = 0.5 by annealing the compounds under a controlled environment. For example, either rapid quenching at high temperatures or annealing under a low oxygen partial pressure, an inert atmosphere (e.g., Ar), or a reducing atmosphere (e.g., 1% H2 in Ar), could introduce more oxygen vacancies into the crystalline structure.20,21,32 However, it is hard to eliminate oxygen vacancies at x ≥ 2/3 even when annealing under pure oxygen is performed.19–21,32 It is even more difficult to yield the fully oxygenated SrFeO3 without extreme conditions, that is, under pure oxygen at extremely high pressure (≥20 MPa).22,33,34
In this work, we have developed an ozone treatment method which can greatly suppress the formation of oxygen vacancies in La1−xSrxFeO3−δ with x ≥ 2/3, including SrFeO3−δ. Ozone is a strong oxidant that has been used to reduce oxygen vacancies; however, these ozone methods were limited to post-treatment, or in some cases coupled with ultraviolet (UV) radiation,35,36 for the removal of the surface oxygen vacancies of metal oxides.37 To our knowledge, ozone has not yet been demonstrated to reduce the oxygen vacancies of the bulk, especially for those cases in which full oxygenation of the bulk is difficult to achieve without a high-pressure oxygen treatment. This study demonstrates a new approach beyond the simple post-treatment in which ozone is applied in situ during the cooling process of the synthesis. A series of powder samples of La1−xSrxFeO3−δ (x = 0.0, 0.2, 0.4, 0.6, 0.8, and 1.0) were first synthesized using a polymerized complex method. The crystal structures of these samples were characterized using X-ray powder diffraction (XRD) and a Rietveld refinement. The content of Fe(IV) and oxygen vacancies in the samples were determined by redox titration. Ozone treatments were then applied to treat those samples with oxygen deficiency (i.e., La0.2Sr0.8FeO3−δ and SrFeO3−δ) while varying temperature, duration, and cooling rate. It was found that a slow cooling rate is the key to achieving high oxygen stoichiometry. As a result, highly oxygenated compounds, including La0.2Sr0.8FeO2.994 and SrFeO2.979, were successfully synthesized via ozone treatment at 225 °C for 6 h with subsequent slow cooling to 25 °C at 0.05 °C min−1. Our findings demonstrate that ozone treatment under atmospheric pressure can replace the high-pressure oxygen annealing in previous studies for the synthesis of highly oxygenated La1−xSrxFeO3−δ perovskites with high Sr/La ratios (x ≥ 2/3).
For the samples containing Sr, i.e., La1−xSrxFeO3−δ with x = 0.2, 0.4, 0.6, 0.8, and 1.0, the same method was used, but the La precursor was replaced by the Sr precursor in the synthesis, according to the corresponding molar ratio. For the synthesis of La0.2Sr0.8FeO3−δ and SrFeO3−δ, impurities were found in the final product when a large amount of the sample powder was calcined at 1300 °C using a single alumina crucible (cylindrical with covering lid, diameter: 40 mm, height: 40 mm, MTI corporation). To prevent this issue, high-purity samples were synthesized using the following protocols: for La0.2Sr0.8FeO3−δ, the sample powder was equally separated into 8 alumina crucibles, with 4 of them calcined at 1300 °C at a time. For SrFeO3−δ, the sample powder was equally separated into 16 alumina crucibles, with 2 of them calcined at 1300 °C at a time.
The Fe(IV) concentration (y) and the oxygen vacancy (δ) in La1−xSrxFe(III)1−yFe(IV)yO3−δ can be calculated using the eqn (1) and (2), based on the charge balance of the chemical formula and the number of moles of Fe(IV) (n, in mol) in a sample with measured mass (m, in g) determined by the titration.
Charge balance:
3(1 − x) + 2x + 3(1 − y) + 4y = 2(3 − δ) | (1) |
Fe(IV) determined by titration:
ym/[MLa(1 − x) + MSrx + MFe + Mo(3 − δ) = n] | (2) |
δ = (242.747n − 51.285nx − mx)/(15.999n − 2m) | (3) |
With δ known, the Fe(IV) concentration can be calculated using eqn (4).
y = x − 2δ | (4) |
The stoichiometries of La (1 − x) and Sr (x, 0 ≤ x ≤ 1) are obtained from the molar ratio of the precursors and confirmed by inductively coupled plasma mass spectrometry (ICP-MS) measurement.
These La1−xSrxFeO3−δ compounds were characterized by XRD to determine their corresponding crystal structures. Fig. 2 displays the XRD patterns of the La1−xSrxFeO3−δ compounds with x = 0, 0.2, 0.4, 0.6, 0.8, and 1.0. The XRD peak positions and their relative intensities are, in general, consistent with previously reported data for these compounds, prepared using a solid-state reaction or the Pechini method.49–51 To further confirm the crystal structure and estimate structural parameters such as lattice constants and atomic distances, we performed a Rietveld refinement on each of these XRD patterns based on known structural information in the literature. For example, it is known that fully oxygenated La1−xSrxFeO3 (i.e., δ = 0) undergoes phase changes as the Sr composition increases, from an orthorhombic phase (Pbnm) for 0 ≤ x ≤ 0.2, to a rhombohedral phase (Rc) for 0.4 ≤ x ≤ 0.7, and a cubic phase (Pmm) for 0.8 ≤ x ≤ 1.0.19,32,34,50 We also took into consideration the oxygen vacancy effects on the crystal structure in choosing the initial model crystal structures for refinement. This is particularly important for the compounds with Sr composition x > 0.6, where a significant amount of oxygen vacancies were found,20 and where the ordering of oxygen vacancies had a profound effect on the crystal structure.22,32,50 For example, the fully substituted end-member, SrFeO3−δ, was found to have the following four different crystal structures, depending on the oxygen stoichiometry: SrFeO3 (cubic, space group Pmm), SrFeO2.875 (tetragonal, space group I4/mmm), SrFeO2.75 (orthorhombic, space group Cmmm), and SrFeO2.5 (orthorhombic, Ibm2).22 Based on this knowledge, the experimental XRD patterns of La1−xSrxFeO3−δ compounds with 0 ≤ x ≤ 0.6 fit well with the calculated patterns for the corresponding structures of the fully-oxygenated La1−xSrxFeO3 (δ = 0). For La0.2Sr0.8FeO3−δ, a reasonable agreement between the observed diffraction data and calculated profile was achieved using a Pmm cubic structural model with oxygen stoichiometry of 2.96 (i.e., La0.2Sr0.8FeO2.96). This model was chosen based on literature suggestions that La0.2Sr0.8FeO3−δ prepared using a solid-state reaction method with annealing at 1230 °C for 48 h and then furnace cooling to room temperature had an oxygen stoichiometry of 2.96.20 For SrFeO3−δ, the experimental XRD patterns were refined by the four above-mentioned crystal structures (Fig. S5 and S6, ESI†). Based on the results, there could be two phases co-existing in our SrFeO3−δ compound, with a major phase of SrFeO2.75 (orthorhombic, space group Cmmm) and a minor phase of SrFeO2.875 (tetragonal, space group I4/mmm).
Fig. 2 Powder XRD patterns of La1−xSrxFeO3−δ compounds, synthesized using a polymerized complex method and calcined in air, as described in Fig. 1. The patterns were indexed according to the calculated patterns generated using the VESTA software and the crystallographic information files obtained from the Rietveld refinement. |
The Rietveld refinement results are plotted in Fig. S7 (ESI†) by overlaying the experimental and calculated patterns for each compound. The structure of each compound refined by the Rietveld method is presented in Fig. 1. The data obtained from the Rietveld method are listed in Table 1. A gradual decrease in the Fe–O bond length with increasing Sr was found, from x = 0.0 to x = 0.8. The lengths of Fe–O bonds in iron oxides are typically 1.91–1.95 Å, 1.990–2.006 Å, and about 2.14 Å for Fe(IV), Fe(III), and Fe(II), respectively.49 These results indicate that an increased percentage of Fe in La1−xSrxFeO3−δ is oxidized from Fe(III) to Fe(IV) with increasing Sr content up to x = 0.8. For the fully-substituted end-member SrFeO3−δ, however, the presence of a large quantity of oxygen vacancies results in the reduction of Fe(IV) to Fe(III). The exact content of Fe(IV) in the La1−xSrxFeO3−δ compounds needs to be determined in order to verify the findings from the XRD analysis, as will be done in the following.
Sample (x) | 0.0 | 0.2 | 0.4 | 0.6 | 0.8 | 1.0 |
---|---|---|---|---|---|---|
Note: a, b, and c are the lattice constants, V is the cell volume, and χ2 is the goodness-of-fit. | ||||||
Model | LaFeO3 | La0.8Sr0.2FeO3 | La0.6Sr0.4FeO3 | La0.4Sr0.6FeO3 | La0.2Sr0.8FeO2.96 | SrFeO2.75 |
Space group | Pbnm | Pbnm | Rc | Rc | Pmm | Cmmm |
a (Å) | 5.5560 | 5.5521 | 5.5301 | 5.4989 | 3.8728 | 10.9522 |
b (Å) | 5.5666 | 5.5249 | 7.7098 | |||
c (Å) | 7.8558 | 7.8196 | 13.4294 | 13.4221 | 5.4707 | |
V (Å3) | 242.96 | 239.87 | 355.67 | 351.48 | 58.09 | 461.94 |
χ 2 | 1.41 | 1.29 | 1.56 | 1.64 | 1.77 | 1.52 |
<Fe–O> (Å) | Fe–O1: 2.0235 | Fe–O1: 1.9986 | Fe–O: 1.9660 | Fe–O: 1.9484 | Fe–O: 1.9364 | Fe1–O1: 1.9244 |
Fe–O1: 1.9849 | Fe–O1: 1.9641 | Fe1–O3: 1.8806 | ||||
Fe–O2: 2.0174 | Fe–O2: 1.9915 | Fe2–O2: 1.9356 | ||||
Fe2–O3: 2.0028 |
The content of Fe(IV) and the associated oxygen stoichiometry in the La1−xSrxFeO3−δ compounds were determined using redox titration. As illustrated in Fig. 3a, the titration involves a two-step process in which the Fe(IV) in the La1−xSrxFeO3−δ compound is first reacted with an excess amount of Fe(II), and then the remaining Fe(II) is titrated with a standard Ce(IV) solution. As a result, the number of moles of Fe(IV) in the La1−xSrxFeO3−δ compound equals the difference between the numbers of moles of the added Fe(II) and Ce(IV) consumed during the titration. The concentrations of Fe(IV) (y) and oxygen vacancies (δ) in each La1−xSrxFe(III)1−yFe(IV)yO3−δ compound were calculated based on the amounts of the added Mohr's salt or Fe(II), the sample used, and the 0.015 M Ce(SO4)2 solution consumed during titration. The calculation is detailed in the Experimental Method section. The titration results are summarized in Table S2 (ESI†) and are highly reproducible, as indicated by the standard errors of less than 0.004.
Fig. 3 (a) Schematic illustration of the working principle of the redox titration of Fe(IV). (b) Plot of the Fe(IV) (y, left ordinate) and oxygen vacancy (δ, right ordinate) content against Sr content (x) in the as-synthesized La1−xSrxFe(III)1−yFe(IV)yO3−δ compounds calcined in air. The y and δ values were determined from the titration results detailed in the ESI.† The standard error is too small to be seen on the graph, but a full list of the data (including the standard errors) is listed in Table S2 (ESI†). |
The stoichiometry of Fe(IV) (y) and oxygen vacancy (δ) are plotted against the Sr content (x) in the La1−xSrxFe(III)1−yFe(IV)yO3−δ compounds in Fig. 3b. A linear increase in y, the stoichiometry of Fe(IV), with increasing Sr composition (x) was observed until x = 0.6 (y = 0.985x with R2 = 0.9989). The slope of the linear fit is 0.985, suggesting that the substitution of La(III) by Sr(II) is nearly completely compensated by the conversion of Fe(III) to Fe(IV) in the La1−xSrxFeO3−δ perovskites (for 0 ≤ x ≤ 0.6). In other words, the number of oxygen vacancies (δ) in these compounds is very small. However, when the Sr composition (x) was further increased from 0.6 to 1.0, the curve deviated from the low-x linear relationship, only reaching y = 0.667 at x = 0.8. At x = 1.0, we even find a slight decrease to y = 0.611. The results suggest that significant amounts of oxygen vacancies are introduced to the La1−xSrxFeO3−δ perovskites with x > 0.6. These observations are consistent with previous studies that show how the content of the oxygen vacancies increases with increased Sr composition and that it is impossible to achieve fully oxygenated compounds with Sr composition x > 0.6 in air.20,26 The results also corroborate the XRD analysis that the SrFeO3−δ compound with oxygen stoichiometry of 2.805 would likely be a mixture of two different phases. This, in turn, agrees well with the previous finding that SrFeO2.75 (orthorhombic, space group Cmmm) and SrFeO2.875 (tetragonal, space group I4/mmm) can coexist in SrFeO2.80.23
The ozone treatment was carried out under min. 5% O3/O2 at ambient pressure. We investigated the effects of temperature, duration time, and cooling rate during ozone treatment on the Fe(IV) content and oxygen stoichiometry of the highest Sr containing compound, SrFeO3−δ, which is the most difficult one to be fully oxygenated. Fig. 4a–c, displays the correlation of the ozone treatment conditions and the changes of the Fe(IV) (y) and oxygen vacancy (δ) content of SrFe(III)1−yFe(IV)yO3−δ, monitored by the redox titration. The detailed parameters of each experiment are listed in Table S3 (ESI†). For a fully oxygenated compound, it is expected to have a stoichiometry of SrFe(IV)O3 with y = 1 and δ = 0. We first studied the effect of temperature in the region of 200 to 400 °C (Fig. 4a; E2 to E7 in Table S3, ESI†). It was hypothesized that an increased temperature could lower the activation barrier of the reaction based on the transition state theory (Arrhenius equation). However, the result indicated that with the increase of the ozone treatment temperature, the value of y first increased to reach the highest value of Fe(IV) at 0.887 obtained at 225 °C and then decreased (Fig. 4a). This implies that the temperature effect is not only to increase the rate of the forward reaction (kf) – oxygenation reaction of the ozone treatment, but also to increase the rate of the reverse reaction (kb) – deoxygenation reaction that causes the formation of oxygen vacancies.23 The difference of the temperature effects on kf and kb appear to suggest an optimal temperature for the oxygenation ozone treatment at 225 °C. However, these experiments were done at a similar cooling rate by allowing the furnace to naturally cool to 100 °C, followed by opening the furnace cover for fast cooling to room temperature. In fact, we found that the content of Fe(IV) significantly increased when the cover of the furnace was opened at a lower temperature (Fig. 4b). Varying the temperature at which the furnace cover was opened, the value of y increased from 0.777 to 0.891 for opening at 150 and 80 °C, respectively. This result suggests that the cooling rate plays a significant role in controlling the kf and kb and shifts the oxygenation/deoxygenation reaction equilibrium during ozone treatment.
To further examine the effect of the cooling rate, the ozone treatment of SrFeO3−δ was performed under a series of controlled cooling rates (E10 to E12 in Table S3, ESI†). The plot of y and δ against cooling rate in Fig. 4c indicates a significant increase in the content of Fe(IV) with decreased cooling rate. The values of y reached 0.950, 0.957, and 0.958 at cooling rates of 0.5, 0.05, and 0.01 °C min−1, respectively. The value of y increased only slightly when the cooling rate decreased from 0.05 to 0.01 °C min−1, indicating that the maximum value of y that could be obtained using this method is about 0.958. The corresponding highest oxygen stoichiometry can reach 2.979, which is within the range of those obtained via annealing under pure oxygen at high pressure (Table 2). The ozone treatment under controlled cooling rates was also carried out for La0.2Sr0.8FeO3−δ (E13 to E15 in Table S3, ESI†). Similar to the trend of SrFeO3−δ, a significant increase in y was observed with decreasing cooling rate (Fig. 4d). The highest y value, reached at a cooling rate of 0.05 °C min−1, was 0.789, which is only 0.13% different from the theoretical maximum value of 0.8, resulting in a stoichiometry of La0.2Sr0.8FeO2.994 with nearly full oxygen occupation. The results demonstrate that the ozone treatment indeed offers a robust alternative to reverse the oxygen loss and fully oxygenate perovskites at atmospheric pressure.
Compound | Temperature, time | Oxygen pressure | Cooling condition | Method for δ determination | Ref. |
---|---|---|---|---|---|
SrFeO2.979 | 225 °C, 6 h | 101.325 kPa | 0.01 °C min−1 to RT | Redox titration | This work |
SrFeO2.998 | 400 °C, 24 h | 30 MPa | 1 °C h−1 to 200 °C | Thermogravimetric | 22 |
SrFeO2.961 | 450 °C, 24 h | 20 MPa | N/A | Redox titration | 19 |
SrFeO2.98 | 900 °C, 12 h | 20 MPa | 5 °C min−1 to RT | Thermogravimetric | 33 |
SrFeO2.94 | 900 °C, — | 35.5 MPa | Slowly cooled | Thermogravimetric | 34 |
The corresponding structures of the SrFeO3−δ and La0.2Sr0.8FeO3−δ compounds with different degrees of oxygenation were then characterized by XRD. The XRD patterns of the SrFeO3−δ and La0.2Sr0.8FeO3−δ compounds are shown at the top of Fig. S8 and S9 (ESI†), respectively. The oxygen vacancy concentration has a significant impact on the crystal structures of the SrFeO3−δ and La0.2Sr0.8FeO3−δ compounds, as seen by comparing the peak positions and shapes of the XRD patterns. Detail plots of the major XRD peaks are also shown in Fig. S8 and S9 (ESI†). The peak(s) of each XRD pattern at 2θ = 47° are displayed in Fig. 5 to further elaborate the changes obtained in the XRD patterns for these two sets of compounds, along with the crystal structure indices generated by the Rietveld refinement. This particular 2θ angle is chosen because one peak would arise for the cubic crystal structure, corresponding to the reflection plane (200) that is perpendicular to the c-axis of the cubic structure; however, two peaks would appear for the orthorhombic structure, with one indexed as (402) and the other indexed as (040).
Fig. 5 XRD region centered at 2θ = 47° for the ozone treated compounds: (a) SrFeO3−δ, and (b) La0.2Sr0.8FeO3−δ. The full XRD patterns are shown in Fig. S8 and S9 (ESI†). |
For SrFeO3−δ, both the peak positions and shapes of the diffraction pattern at 2θ ≈ 47° change significantly with increasing oxygen vacancy content (Fig. 5a). For the two near fully-oxygenated SrFeO2.979 samples, a peak at 47.22° with a shoulder (of about half its intensity) at 47.34° can be assigned to the (200) reflection of the cubic structure. The peak splitting arises from the Cu Kα doublet of the X-ray source. The simulated pattern from Rietveld refinement of the experimental data yields a doublet at 47.18° and 47.30°, in agreement with our experimental observation. For the highly oxygen-deficient SrFeO2.805, three peaks were identified (at 46.97°, 47.13°, and 47.29°), which can be indexed to the two doublet peaks for the (402) and (040) reflections of the orthorhombic structure. The Rietveld refinement of the orthorhombic structure, based on the experimental XRD data, yields simulated (402) reflections at 46.91° and 47.03°, and (040) reflections at 47.11° and 47.23°. The shoulder of the (402) peak at 47.03° overlaps with the major (040) peak at 47.11°, resulting in one broad peak observed in the experimental pattern. A structural transition from a cubic to an orthorhombic structure is observed in SrFeO2.944, with the peak at 47.22° being broadened towards lower angles. The transition is more obvious in SrFeO2.910, with the presence of a shoulder at 47.07°. As a result, the experimental diffraction patterns of SrFeO2.910 and SrFeO2.944 could not be fit using a single structural model in the Rietveld refinement.
For La0.2Sr0.8FeO3−δ, the peak shape remains very similar for all three compounds (La0.2Sr0.8FeO2.994, La0.2Sr0.8FeO2.988, and La0.2Sr0.8FeO2.934), as shown in Fig. 5b. The observed doublet can be assigned to the (200) reflection of the cubic structure based on the Rietveld refinement. As the number of oxygen vacancies increases, the peak shifts to lower angles, suggesting an increase of the lattice constant based on Bragg's Law. This finding is consistent with the Rietveld refinement results.
Table 3 summarizes the results of the Rietveld refinement of the experimental XRD patterns (cubic structures) for the near fully-oxygenated SrFeO3−δ and La0.2Sr0.8FeO3−δ compounds. A comparison of the simulated and experimental XRD patterns is shown in Fig. S10 (ESI†). A decrease in the Fe–O bond distance was observed when decreasing the number of oxygen vacancies in the SrFeO3−δ and La0.2Sr0.8FeO3−δ compounds, which indicates an increase in the oxidation state of Fe. These results are thus consistent with those obtained from the redox titration. For SrFeO3−δ, the Fe–O bond distances estimated from the refinement are 1.9262 Å for SrFeO2.979 (0.05 °C min−1) and 1.9248 Å for SrFeO2.979 (0.01 °C min−1). These values are consistent with the literature reported data of 1.91–1.95 Å for Fe(IV) in iron oxides,49 indicating an oxidation state close to Fe(IV).
Samples | SrFeO2.979 (0.05 °C min−1) | SrFeO2.979 (0.01 °C min−1) | La0.2Sr0.8FeO2.988 | La0.2Sr0.8FeO2.994 |
---|---|---|---|---|
Note: a, b, and c are the lattice constant. V is cell volume. χ2 is goodness-of-fit. | ||||
Model | SrFeO2.96 | SrFeO2.96 | La0.2Sr0.8FeO2.96 | La0.2Sr0.8FeO2.96 |
Space group | Pmm | Pmm | Pmm | Pmm |
a (Å) | 3.8525 | 3.8496 | 3.8673 | 3.8665 |
b (Å) | ||||
c (Å) | ||||
V (Å3) | 57.18 | 57.05 | 57.84 | 57.80 |
χ 2 | 1.99 | 2.17 | 1.92 | 1.71 |
〈Fe–O〉 (Å) | Fe–O: 1.9262 | Fe–O: 1.9248 | Fe–O: 1.9337 | Fe–O: 1.9333 |
The ozone treatment at a slow cooling rate thus provides a robust approach to reverse the loss of oxygen in the SrFeO3−δ and La0.2Sr0.8FeO3−δ perovskite oxides under ambient pressure, as schematically shown in Fig. 6. The ozone generates radicals that can effectively react with the perovskite lattice and fill the oxygen vacancies in highly oxygen-deficient structures; however, simple ozone applied in post-treatment is not sufficient to fully oxygenate the bulk of these materials. The ozone treatment is required to be applied in situ during the cooling process, starting from an optimized temperature of 225 °C, at which the oxygenation is favorable. More importantly, a slower cooling rate facilitates the complete filling of the oxygen vacancies of the bulk. In this study, the cooling rate of 0.05 °C min−1, starting from 225 °C, resulted in near-zero oxygen vacancies for the perovskites with high Fe(IV) content, quantified by the redox titration as SrFeO2.979 and La0.2Sr0.8FeO2.994. The crystal structures and stoichiometry of SrFeO2.979 and La0.2Sr0.8FeO2.994 compounds are stable in air at room temperature after storage for months, as confirmed by XRD and titration.
After ozone treatment, little changes were found in the morphology of the samples for both SrFeO3−δ and La0.2Sr0.8FeO3−δ compounds, as shown in Fig. S11 and S12 (ESI†). They consist of particulates with different sizes ranging from tens of nanometers to tens of micrometers. Based on the EDS analysis, the ratio of Sr to Fe in SrFeO3−δ is approximately 1:1, and the ratio of La:Sr:Fe is close to the stoichiometry in formula of La0.2Sr0.8FeO3−δ. After storage under ambient conditions for 9 months, the crystal structures of the ozone-treated compounds (i.e., SrFeO2.979 and La0.2Sr0.8FeO2.994) remain the same as the freshly-prepared samples (Fig. S13, ESI†). The titration results indicate that the Fe(IV) stoichiometry of ozone-treated SrFeO3−δ at 0.01 °C min−1 only shows a slight decrease by 0.8%, and the Fe(IV) stoichiometry of ozone-treated SrFeO3−δ and La0.2Sr0.8FeO3−δ at 0.05 °C min−1 decreases by 1.8% and 1.5%, respectively (Table S3, ESI†). This method can couple to the commonly-used calcination method to offer a facile means to reduce oxygen vacancies without the need of high-pressure environments.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ma00604a |
This journal is © The Royal Society of Chemistry 2022 |