Jacinthe
Gamon‡
a,
Arnaud J.
Perez‡
a,
Leanne A. H.
Jones
bc,
Marco
Zanella
a,
Luke M.
Daniels
a,
Rhun E.
Morris
a,
Chiu C.
Tang
d,
Tim D.
Veal
bc,
Laurence J.
Hardwick
ab,
Matthew S.
Dyer
a,
John B.
Claridge
a and
Matthew J.
Rosseinsky
*a
aDepartment of Chemistry, University of Liverpool, Crown Street, Liverpool, L69 7ZD, UK. E-mail: rossein@liverpool.ac.uk
bStephenson Institute for Renewable Energy, University of Liverpool, Chadwick Building, Peach Street, Liverpool, L69 7ZF, UK
cDepartment of Physics, University of Liverpool, Oliver Lodge Laboratory, Oxford Street, Liverpool, L69 7ZE, UK
dDiamond Light Source Ltd, Harwell Science and Innovation Campus, Didcot, OX11 0DE, UK
First published on 21st September 2020
Multiple anion materials are of particular interest for the discovery of new crystal structures and offer an original way to modulate physical properties, including energy storage materials with enhanced performances. Through careful synthesis optimization, a new Na2Fe2OS2 phase was prepared by two different routes: high temperature solid-state synthesis and simple mechanochemical synthesis. The long-range and local structure of Na2Fe2OS2 was studied by Rietveld refinement of neutron and X-ray diffraction data combined with EXAFS data refinement. The phase comprises an amorphous and a crystalline part which has an anti-K2NiF4 structure, corresponding to the n = 1 member of the homologous anti-Ruddlesden–Popper [AX][ABX3]n series. Its electrochemical properties as a cathode material were studied in Na half cells and Na-ion full cells, revealing that the material becomes fully amorphous upon initial desodiation to Na0.5Fe2OS2, but maintains a reversible capacity of 135 mA h g−1 in full cells where up to 1.2 Na+ can be reversibly extracted and reinserted when compensating for the Na lost in SEI formation. The stability of the pristine material and its structural evolution upon charging are discussed, paving the way for further optimization of this material. Being composed exclusively of earth-abundant elements and stable under dry air, Na2Fe2OS2 perfectly illustrates the great opportunity of multiple anion chemistry to explore new structure types and develop better energy storage systems.
While many studies focus on the substitution of cations within an oxide lattice, the combination of two anions in one structure type offers alternative ways to modulate structure and properties.3 Hetero-anionic materials containing anions of different sizes and electronegativity tend to form layered structures, favorable for chemical intercalation, in which each cation bonds to one of the anions according to their preferred chemical interactions.4–8 The use of oxyfluorides as cathode materials is an active field of research,9,10 as the higher electronegativity of fluoride ions leads to high voltage electrode materials. Less electronegative anions, such as chalcogenides (Ch = S2−, Se2−), are more easily oxidized than oxides, favoring Ch2− anionic redox activity11 and higher capacities.12 The presence of highly polarizable chalcogenide anions is also a strong advantage to increase electronic conductivity, which often limits the electrochemical properties in oxide systems.
In that matter, oxysulphides represent a promising family of materials, which is relatively unexplored, mostly due to their lower voltage compared to oxides. However, recent results on a new anti-perovskite Li2FeOS have shown that the low voltage can be compensated by large capacities alongside the advantage of using abundant elements.13
Reported for its interesting electronic and magnetic properties,14 Na2Fe2OSe2 is a good candidate for Na-ion batteries with a reasonable amount of Na that gives a theoretical capacity of 162 mA h g−1. Its layered structure should lead to good Na-ion conductivity and good electronic conductivity in the Fe2O layer. The sulphide equivalent Na2Fe2OS2, which has not been reported, has an even higher theoretical capacity of 225 mA h g−1, while being composed of elements that are among the most abundant on our planet.1
Due to its reactivity against oxidation with oxygen and hydrolysis with water, synthesis and processing of sulphide materials is expensive and often requires the use of toxic gases and/or solvents which generate safety and environmental issues at the industrial level. Developing low temperature and solvent-free synthesis routes of sulphide materials has therefore become an important research topic.15–17 We report here on the synthesis of the promising oxysulphide Na2Fe2OS2 through mechanochemistry, a scalable, room temperature route, with thorough characterization of its structural features, and unravel its electrochemical properties as a cathode material for Na-ion batteries through a combination of advanced characterization techniques.
Mechanochemical synthesis has been shown to be a powerful tool for the synthesis of metastable and soft materials such as chalcogenides,19–21 while avoiding severe limitations on the use of toxic sulphurizing agents and therefore gaining substantial interest from industry.22,23 It additionally has the benefit of producing sub-micron particles with appreciable amount of structural disorder and improved ionic conductivity,24–26 leading scientists to adopt this method for battery material processing.27–30 Mechanochemical synthesis was thus used for the preparation of Na2Fe2OS2. Details on the optimization procedure for the mechanochemical synthesis can be found in the Methods section and in part 2.1 of the ESI.† By adding an excess of Na2S to the reaction media (1Na2O + 2FeS + 0.75Na2S), we were able to isolate the target Na2Fe2OS2 phase. During high energy ball milling, high temperatures in excess of 700 °C can be reached through friction at the contact point of colliding balls and over a very short period of time (10−4 to 10−3 s). These high temperatures have been described by the plasma-magma and impulse models,31 and are due to the large amount of energy transferred to the reagents, estimated as the kinetic energy of the ball before collision32 (of the order of ∼100 J).33 The Na2Sx polysulphides,34 with melting points ranging from 490 °C for Na2S2 to 252 °C for Na2S6, are therefore likely to act as a flux, known to favor the crystallization of sulphide materials.35 The excess, unreacted Na2S was removed through washing the powder in anhydrous methanol, resulting in a powder free from any Na3Fe2S4 and FeO crystalline impurities (cf. ESI, part 2.1†). In-depth structural and electrochemical properties characterization were then performed on this mechanosynthesized and methanol-washed sample, denoted Na2Fe2OS2-MW.
The micro-structural analysis (implemented in Fullprof, cf. Methods)36 revealed an average coherent scattering domain size of 22(2) nm (based on the Debye–Scherrer model) with no particular size anisotropy, and an average maximum strain of 31.98(5)‰ (upper limit of the apparent strain defined by Stokes and Wilson)37 compared to 136.8(2) nm and 5.51(1)‰, respectively for Na2Fe2OS2-SS. The SXRD pattern shows a broad contribution between 1–2 Å−1, which can be partially attributed to the capillary used for the measurement (Fig. S10†). However, the significant contribution of the background to the SXRD pattern indicates the potential presence of an additional amorphous phase which does not contribute to the Bragg scattering, and is not unheard of through mechanosynthesis methods.38
The relative proportions of crystalline and amorphous components in the material were assessed using the Rietveld quantitative amorphous content analysis method by adding crystalline Na3Fe2S4 as a suitable internal standard (cf. ESI part 2.2.2†). This analysis revealed that 39(5) wt% of the phase is crystalline, i.e. 61(5) wt% of the material is not observed using diffraction experiments. This result highlights the importance of using the quantitative phase analysis method with an internal standard to study mechanosynthesized materials. Attempts to increase the extent of crystallization in the compound by annealing at 400 °C or 600 °C, or by extended ball milling resulted in the decomposition of the phase, further confirming the metastability of Na2Fe2OS2.
The elemental analysis performed by Inductively Coupled Plasma Optical Emission Spectrometry (ICP/OES), CHNS, together with Scanning Electron Microscopy coupled with Wavelength-Dispersive X-ray spectroscopy (SEM-WDX) revealed an overall composition: Na2.2(1)Fe2.00(1)OS2.29(1) (Table S4 and Fig. S11†) close to the expected composition Na2Fe2OS2. The excess Na and S content can be attributed to some remaining sodium sulphide species (2.5(5)% according to the mass balance from the washing procedure).
Transmission Electron Microscopy (TEM) imaging and Energy Dispersive X-ray spectroscopy (EDX) mapping were performed to gain more information about the microstructure of the sample. The imaging revealed the presence of large micrometric particles composed of agglomerated nanograins (Fig. 1a). The elementary ∼20 nm coherent scattering domains revealed by XRD could not be isolated. EDX mapping revealed a homogeneous distribution of the four elements over the whole particle, aside from the 70 nm wide Fe metal grain impurity (Fig. 1b). These results, added to the estimated lateral resolution from the EDX probe of ∼2–5 nm,39 rule out the presence of any other impurity with grains larger than 7 nm and show that the Na2Fe2OS2 composition is homogeneous on the nanometer scale.
![]() | ||
Fig. 1 Characterisation of Na2Fe2OS2-MW. (a) TEM imaging of a particle of Na2Fe2OS2-MW and (b) the corresponding EDX mapping images of the elemental distribution. (c) Infrared spectra of Na2Fe2OS2-SS, Na2Fe2OS2-MW and of a Na3Fe2S4 reference prepared by solid-state synthesis. Rietveld refinement of Na2Fe2OS2-MW against (d) synchrotron X-ray diffraction and neutron powder diffraction data from (e) bank 2 (2θ = 168.330°) and (f) bank 5 (2θ = 30.000°) of Polaris instrument, with Iobs (red dots), Icalc (black line), Iobs–Icalc (blue line), and Bragg reflections (black tick marks for Na2Fe2OS2, blue tick marks for Fe, and pink tick marks for ZrO2). Contribution of the capillary to the background of the pattern is shown in Fig. S10.† |
Fourier Transform Infrared spectroscopy (FTIR) was used to further probe the compositional homogeneity on the local scale and discriminate the presence of amorphous impurities. Fig. 1c shows the spectra of the crystalline and mechanosynthesized samples along with a Na3Fe2S4 reference material. Both Na2Fe2OS2-MW and Na2Fe2OS2-SS samples show similar features, with one group of broad and low intensity peaks between 700 and 550 cm−1, a broad and intense signal between 550 and 400 cm−1, and another group of different intensity peaks between 400 and 280 cm−1. Additionally, Na2Fe2OS2-SS shows a signal with increasing intensity below 280 cm−1, which is not observed in Na2Fe2OS2-MW. This signal can be attributed to the Na3Fe2S4 impurity (Fig. 1c and Table S7†).
Peaks were fitted using the Gaussian function in order to compare positions and full width at half maximum (FWHM). Fig. S15† shows the result of the fit and fitting parameters are reported in Table S8.† Vibration modes with wavenumbers between 400 and 280 cm−1, which are also observed for Na3Fe2S4 (Fig. 1c), can be attributed to Fe–S stretching modes.40–42 For Na2Fe2OS2-SS this region consists of 3 distinct peaks of FWHM ranging from 10 to 24 cm−1 whereas for Na2Fe2OS2-MW it consists of a much broader signal. This could be fitted by three overlapping contributions and the two main peaks at 292(2) and 310(2) cm−1 show large FWHM of 109(3) and 50(2) cm−1, respectively. The comparison between FWHM of Na2Fe2OS2-SS and Na2Fe2OS2-MW reflects the high amount of structural disorder as well as the large strain present in the mechanosynthesized sample, therefore susceptible to show a broader range of Fe–S distances. Vibrations at higher wavenumbers, between 550 and 400 cm−1, can be attributed to metal–oxygen stretching bonds.43 Interestingly, these peaks have comparable FWHM for Na2Fe2OS2-MW (52(3) and 40(2) cm−1) and Na2Fe2OS2-SS (66(7) and 64(3) cm−1), which is consistent with the rigid nature of the short Fe–O bonds compared to Fe–S bonds.
Overall, no additional peaks could be identified for Na2Fe2OS2-MW. This observation suggests that the intimately combined coherent and incoherent scattering domains have similar local structure with a broader range of interatomic distances compared to the purely crystalline phase (Na2Fe2OS2-SS).
a (Å) | 4.0325(1) (Na2Fe2OS2-MW); 4.04222(1) (Na2Fe2OS2-SS) | |||
c (Å) | 14.077(1) (Na2Fe2OS2-MW); 14.07319(9) (Na2Fe2OS2-SS) | |||
V (Å3) | 228.5(2) (Na2Fe2OS2-MW); 229.950(2) (Na2Fe2OS2-SS) | |||
Site | Na | Fe | O | S |
Wyckoff position | 4e | 4c | 2a | 4e |
Site symmetry | 4mm | mmm | 4/mmm | 4mm |
(x, y, z) | (0, 0, 0.176(2)) | (0, 0.5, 0) | (0, 0, 0) | (0, 0, 0.376(3)) |
sof | 1 | 0.99(2) | 1 | 0.96(4) |
B iso/B11, B22, B33, Beq (Å2) | 1.6(3) | 2.0(4), 1.0(4), 1.8(4), 2.1(5) | 0.1(2), 0.1(2), 1.6(9), 1.2(5) | 2.3(5) |
Coordination | 6 | 6 | 6 | 9 |
BVS | 1.06(2) | 1.80(1) | 2.14(1) | 1.84(1) |
Distance (Å) | d Na–S: 2.808(3) | d Fe–S: 2.670(2) × 4 | — | — |
d Na–S: 2.9410(8) × 4 | d Fe–O: 2.0152(1) × 2 | — | — | |
d Na–O: 2.478(2) | d Fe–Fe: 2.8500(2) | — | — | |
d Na–Fe: 3.1936(17) | — | — | ||
Angles (°) | S–Na–S: 104.30(10) × 4 | S–Fe–S: 98.02(4) × 2 | — | — |
S–Na–S: 86.50(2) × 4 | S–Fe–S: 81.98(9) × 2 | — | — | |
S–Na–S: 151.41(2) × 2 | S–Fe–S: 180.00(9) × 2 | — | — | |
S–Na–O: 75.70(6) × 4 | S–Fe–O: 90.00(5) × 8 | — | — | |
S–Na–O: 179.97(12) | O–Fe–O: 180 | — | — |
There was no indication of anisotropic displacements of S and Na from the refinement, but anisotropic displacement parameters were found for Fe and O (Table 1), suggesting possible displacement of these atoms. However, a model with Fe atoms displaced from the 4c (0, 0.5, 0) onto a 16n (x, 0.5, z) site did not improve the fit (Fig. S13†) and was therefore discarded. During independent refinement, sof of Na refined to a value higher than 1, and was therefore fixed to 1, while the sof of Fe, S and O refined to 0.99(2), 0.96(4) and 0.98(1) respectively. In contrast to Na2Fe2OS2-SS for which the refinement highlighted an Fe deficiency (Table S2†), the refined composition of Na2Fe2OS2-MW, Na2Fe1.98(2)S1.96(4)O0.98(1), is closer to the ideal stoichiometry and could explain why this sample has a smaller a lattice parameter.
Na2Fe2OS2 was found to be isotypic with its selenium analogues.14 The phase crystallizes in an anti-K2NiF4 type structure (Fig. 2a). K2NiF4 corresponds to the n = 1 member of the Ruddlesden–Popper (RP) homologous series of general formula [AX][ABX3]n, where A is a large cation, B a smaller cation and X an anion (Fig. 2b). In the anti-structure of Na2Fe2OS2, the [ABX3] layer is an anti-perovskite slab which can be written as [SO(Fe2Na)]+ and the rocksalt [AX] layer as [SNa]−, with large sulphur anions on the A site, small oxygen anions on the B site, and iron and sodium cations ordered on the two inequivalent X sites. As expected from the size and electronegativity difference of O2− and S2−, anion ordering is clearly evidenced in the structure. Interestingly, the layers of the anti-RP phases are charged, which is in contrast with the regular n = 1 RP phase.
The structure can also be described as alternating layers of [Fe2OS2]2− and sodium double layers [Na2]2+ (Fig. 2c). Other phases containing this isostructural [M2OQ2]2− motif (M = Fe, Ti; Q = S, Se, As, Sb) have been reported in the literature. Among them, Na2Ti2Pn2O (Pn = As, Sb)44 is isostructural to Na2Fe2OS2 and Na2Fe2OSe2 (with Ti in the Fe site and Pn in the S or Se site); BaTi2Pn2O shows a barium single layer (Fig. S14b†), and La2O2Fe2OSe2 a fluorite type [La2O2]2+ layer (Fig. S14c†)45,46 instead of a sodium double layer [Na2]2+. Several phases containing the anti-[CuO2] square planar configuration within the [M2OQ2]2− layer have been studied as potential superconductors,47,48 as the square planar lattice is an important feature of the well-known high temperature cuprate superconductors.49,50 They revealed interesting physical properties such as anti-ferromagnetic ordering in the 50–100 K range14,51 or spin-density-wave/charge-density-wave instabilities.48,52,53
Looking more closely at individual atomic environments in Na2Fe2OS2, Fe is coordinated by two oxygen atoms in the anti-[CuO2] type square planar layer, and four S atoms above and below the plane (Fig. 2d). Fe therefore sits in an octahedral site compressed along the axial direction (Fe–O), with the four S atoms in the equatorial plane forming a rectangle with two obtuse (99.02(6)°) and two acute (80.98(6)°) S–Fe–S angles (Fig. 2d). The Fe–O distance (dFe–O = 2.0152(1) Å, interatomic distances are reported in Table 1) is shorter than that observed in FeO (dFe–O = 2.16885 Å) where Fe2+ is also 6-fold coordinated. On the contrary, the Fe–S distance (dFe–S = 2.670(2) Å) is longer than in most iron sulphide compounds, yielding a bond valence sum (BVS) of 1.80(1) on the Fe atom, close to the expected +2 theoretical value. This geometry around the Fe site is reflected in the values of the adp: the displacement ellipsoid of Fe shows a marked elongation along the a and c axis, revealing a disk-shaped displacement in the equatorial plane of the FeO2S4 octahedra (Fig. 2d). This is also consistent with the higher polarizability of the S2− anions in comparison to O2− which yields more rigid bonds. The rigidity of the Fe–O bond is further featured in the comparison with the structure of other compounds presenting the Fe2OCh2 (Ch = S, Se) layer. For instance, the a lattice parameter (twice the Fe–O distance) in Na2Fe2OSe2 (4.107(8) Å)45 and La2O2Fe2O2Se2 (4.084466(9) Å)14 is very close to that of Na2Fe2OS2 (1.8 and 1.2% difference, respectively). On the contrary, the longer c lattice parameter (14.641(8) and 18.59798(7) Å) accommodates the larger Se2− ions and [La2O2]2+ slabs (3.9 and 24% difference, respectively). Oxygen anions are coordinated by four iron atoms in the ab plane and two sodium atoms along c (Fig. 2g). The adp of oxygen highlight a large elongation of the displacement ellipsoid of oxygen along the c axis. This has been observed previously in compounds with similar structures and was attributed to the preferred five-fold coordination of O over the six-fold coordination.54 Sulphur atoms are slightly off plane with the sodium atoms of the same layer (z(Na)–z(S) = 0.052(5)), such that S is displaced towards the closest Fe2O layer (Fig. 2f). The [SNa]− layer therefore differs from a pure [AX] rocksalt layer in which Na and S atoms occupy the same plane. Finally, sodium atoms lie in a distorted NaOS5 octahedral site (Fig. 2e), with one short, four long Na–S distances, and one short Na–O distance.
The stability of the phase under different atmospheres was further assessed by first synthesizing a sample under Ar and then storing it under dry air or under ambient atmosphere (relative humidity ∼40%) for 20 hours. Within seconds, the powder left under ambient atmosphere changed texture (formation of agglomerates) and became amorphous according to X-ray diffraction (red line, Fig. 3a). In contrast, no loss in crystallinity was observed for the sample left under dry air as attested by the similarity of the XRD pattern with that of the sample stored under Ar (green line, Fig. 3a). This suggests that Na2Fe2OS2 decomposes in the presence of humidity.
The infrared spectra of the three samples are shown on Fig. 3b, with assignment of the peaks in Table S7.† The spectra of the sample synthesized or stored under dry air are similar to that of the pristine material, whereas that of the sample left under humid atmosphere shows features from carbonate and hydroxide decomposition products, further demonstrating the stability of both the crystalline and amorphous content of Na2Fe2OS2 under dry air.
The final structural model defined for the intermediate disordered phase is presented in Fig. 5f. The refined composition, Na1.6(1)Fe2.0(1)OS1.98(10), is close to the expected Na1.7Fe2OS2 composition, suggesting that sodium is homogeneously removed from the amorphous and crystalline parts of the pristine sample. The 16n site is occupied by 62(12)% Fe and 38(12)% Na, while the 4e site contains 41(3)% Na and 38(4)% Fe. This partial occupation represents a high degree of cation disorder, with almost random occupation of Fe and Na on both sites. The BVS value of the 16n site (2.28(7)) is slightly higher than expected for Fe2+ cations, suggesting that potentially oxidized Fe3+ species are most probably located in these positions rather than the 4e site. The larger size of the 4e site and smaller BVS value (1.08(8)), should be more favourable to host Fe2+, which is also more mobile than Fe3+ cations. Assuming that Fe3+ remains in the 16n site, the composition of the initial [Fe2OS2]2− layer becomes [(Fe1.2(2)Na0.8(2))OS2]2.4−, whereas that of the former sodium double layer becomes [Na0.76(8)Fe2+0.82(6)]2.4+.
The decrease of the a lattice parameter reflects the presence of Fe3+, which affects the average Fe–O distance in the ab plane. By considering the rms displacements of both Fe, S and O atoms from the value of their adp, the Fe–O distances range between 1.84(13) and 2.18(13) Å and the Fe–S distances range from 2.080(5) and 3.416(6) Å. It is probable that, on a local scale, S and Fe or Na will arrange within their potential sites so as to prevent the smallest distance which is not favourable for Fe or Na and S bonding. In turn, the increase in the c lattice parameter counterbalances this effect and acts to maintain satisfactory Na–Na and Na–S distance. Overall, a unit cell volume shrinkage is observed (214.8(1) Å3 compared to 228.91(3) Å3 for the pristine sample).
Rietveld analysis was used to calculate the phase fraction of the crystalline material compared to the Fe impurity, expected to be insensitive to cycling, and thus check whether part of the phase becomes amorphous during the phase transition from Na2Fe2OS2 to Na1.7Fe2OS2. The amount of crystalline phase decreases slightly from 98.2(1) wt% for Na2Fe2OS2 to 96.1(3) wt% for Na1.7Fe2OS2, suggesting that only a small part of the crystalline Na2Fe2OS2 phase (2.1(3) wt%) loses crystallinity during the charge to 2.05 V. This shows that the amorphous and crystalline parts of the starting material behave similarly during the electrochemical process and is reinforced by the comparison of the apparent size of the crystallites. Microstructural analysis of Na1.7Fe2OS2, performed in the same way as for Na2Fe2OS2-MW, did not reveal any particular size anisotropy, with a volume-averaged apparent size of the crystallites of 24(5) nm and an average maximum strain of 106.4(6)‰. The apparent particle size did not decrease during the charge to 2.05 V, which indicates that little side amorphization occurs during this first process. However, the anomalously large strain could find its origin through increased structural disorder. This seems consistent with the wide range of bond lengths which indicates significant local relaxation from the average refined cell metrics.
The successive electrochemical reactions happening in Na2Fe2OS2 can be summarized as follow:
Na2Fe2OS2 (crystalline) → Na1.6Fe2OS2 (crystalline) + 0.4Na+ + 0.4e− | (1) |
Na1.6Fe2OS2 (crystalline) → Na0.5Fe2OS2 (amorphous) + 1.1Na+ + 1.1e− | (2) |
Na0.5Fe2OS2 (amorphous) ↔ Na1.95Fe2OS2 (amorphous) + 1.45Na+ + 1.45e− | (3) |
Limiting the cycling between 1.5 and 2.05 V to retain the crystallinity of the material only gives a capacity of 34 mA h g−1 (0.3 Na+/e− exchanged), which is not large enough to represent a good compromise between capacity and structural stability. Despite the material's loss of crystallinity, an initial charge capacity of 170 mA h g−1 (1.5 Na+/e−) and a discharge capacity of 162 mA h g−1 are obtained when cycling between 1.5 and 3 V. This is a significant capacity and it is therefore important to obtain some insight on the local structure evolution of the amorphous compounds upon cycling of the material.
The EXAFS data were fitted to obtain more precise information on the bond distances in the different samples. In the pristine material, structural information for the crystalline part of the sample is available up to R-values of 5–6 Å, and bond distances can be constrained based on the crystallographic parameters obtained from diffraction data (a, c, zNa, zS, see Fig. S22 and Table S11 in ESI† for more detailed information). This model gives a good fit of the data beyond the first shell, but a poor result in the 1–2.4 Å R-range. This suggests that additional contributions in the first shell of atoms around the Fe* absorbing atom, i.e., Fe*–O and/or Fe*–S distances, have to be included to model the data. A good fit could be obtained by adding a short Fe*–S distance (2.290(6) Å) in the model (Fig. 6b), whereas the addition of a Fe*–O distance gave large bond distance values (>2.3 Å), highly correlated to the Fe*–S scattering path, and did not improve the fit significantly. This contribution was therefore not included in the final model. The final fit of the EXAFS data is shown in Fig. S23a and b, and fitting parameters in Table S12.† Interestingly, the short Fe–S distance (2.290(6) Å) is much shorter than the average Fe–S distance of the crystallographic model (2.615(10) Å). This distance could reflect local distortions in the crystalline part of the sample, which can be inferred from the large atomic displacement parameters on Fe and S atoms. The root mean square (rms) displacements of both atoms along the Fe–S vector were calculated from the outcome of the refinement of the diffraction data (detail in part 2.4 of ESI and Table S8†). It shows that the Fe–S distance in the pristine material can be as short as 2.34(15) Å if the rms displacement is considered (green line in Fig. S24†), therefore matching, within error, the distance obtained with EXAFS. Alternatively, the short Fe–S distance could reflect the environment of Fe in the amorphous part of the sample. Future work will consist of an in-depth structural and compositional characterization of this amorphous content produced by ball milling.
The phase charged to 2.05 V, which retains crystallinity according to X-ray diffraction, shows a similar EXAFS signal compared to the fully amorphous samples, whereas one would expect some signal to persist beyond the first shell of atoms. Simulating EXAFS data for different models considering the multiple Fe–Fe distances in Na1.7Fe2OS2, due to antisite defect (38% of Fe on the 4e site) and/or displacement of Fe from the 4c to 16n site, results in a significant reduction of the intensity beyond the first shell (Fig. S25 in ESI†). This explains why the EXAFS data of the Na1.7Fe2OS2 material resembles that of the amorphous samples (3 V and 1.5 V), despite being crystalline. Almost no information beyond the first shell (Fe*–O and Fe*–S) can be obtained because of the increased disorder leading to multiple Fe–Fe distances in Na1.7Fe2OS2.
For the ex situ cycled samples charged at 3 V (Na0.5Fe2OS2) and discharged at 1.5 V after a complete cycle (Na1.95Fe2OS2), the EXAFS fit was restrained to the first two shells, Fe*–O, Fe*–S and Fe*–Fe. A comparison of different fitting models gives good results for mixed oxide/sulphide coordination environments (Rf < 0.1%, see Fig. S27†), whereas pure oxide and sulphide models result in poor fits (Rf > 2.5%). The result of the final fits is shown in Fig. S23† with refined parameters in Table S12.† This analysis suggests that iron in the amorphous material formed upon cycling is still coordinated by oxygen and sulphur atoms, although it is difficult to discriminate between heteroleptic Fe environments (FeOxSy), as in the crystalline Na2Fe2OS2 compound, and/or an oxysulphide with a distribution of homoleptic polyhedra (FeOx and FeSy). The Fe–S distances obtained for the cycled samples (between 2.256(3) and 2.296(5) Å) are close to the short Fe–S distance found for the pristine sample (2.290(6) Å). These distances are typical of Fe–S bonds in tetrahedral FeS4 environments, and the Fe–O distances in the amorphous samples (between 1.878(6) and 1.927(19) Å) are typical of tetrahedral FeO4 environments, which suggests that the amorphisation of the material is associated with a large reorganization of cations/anions in the material. Both Fe*–O and Fe*–S distances decrease upon oxidation of the material to 2.05 V (1.917(8) and 2.268(2) Å, respectively) and further to 3 V (1.878(6) and 2.256(3) Å), and increase again on reduction to 1.5 V (1.927(19) and 2.296(5) Å), suggesting some degree of Fe oxidation/reduction upon cycling.
To remove the effect of SEI formation on the hard carbon and access more reversible capacity from Na2Fe2OS2, we precycled a hard carbon electrode in a Na half cell to form the SEI, recovered it after 5 cycles, washed it in DMC and assembled a full cell with that electrode which was again cycled between 0 and 3 V. This time, the CE on the first cycle reaches 88.4%, suggesting that Na2Fe2OS2 also contributes to the irreversible capacity during the first cycle, and the specific energy of the cell reaches 215 W h kg−1. The CE then settles around 99% for the next 20 cycles, and the capacity further decreases with cycling (77% capacity retention after 20 cycles), although more slowly than with the pristine hard carbon anode (65% capacity retention). Cells opened after 10 and 50 cycles were first compared visually, showing very different results. After 10 cycles, no obvious degradation was observed, but after 50 cycles the glass fiber separator clearly appeared dark where the cathode was placed (Fig. S27a†). XPS measurements on the hard carbon anode after 50 cycles (Fig. S27c and d†) show signals corresponding to Fe 2p and S 2p peaks, thus proving that long term cycling results in iron and sulphur dissolution, migration and deposition on the anode side. This explains, at least in part, the low CE efficiency and the capacity loss with cycling.
To further explore Fe dissolution in this material, a 3-electrode cell (hard carbon counter electrode, Na reference electrode) was used to charge the material above 3 V vs. Na+/Na and extract the remaining 0.5 Na+ from Na0.5Fe2OS2 (Fig. S27b†). The voltage stopped increasing at 3.2 V vs. Na+/Na, dropped to 3.15 V, then remained at a flat potential of 3.05 V. This gives an apparently unlimited capacity, well above the theoretical capacity of Na2Fe2OS2 or the hard carbon electrode, and leads to a complete loss of cycling performances afterwards. The color of the separator recovered from the cell appeared dark (Fig. S27a†), and XPS analysis of the hard carbon anode (Fig. S27c and d†) recovered from a coin cell after cycling in equivalent conditions, shows that the Fe/S dissolution can also be triggered by overcharging the cathode above 3.2 V vs. Na+/Na. Soluble species in the electrolyte can potentially act as redox shuttles, as can be the case in Na–S systems,57 thus providing such large capacity.
Overall, the material itself shows an intrinsic instability when charged above 3.2 V vs. Na+/Na, which cannot be prevented. However, cation dissolution over long term cycling has been mitigated by using surface coatings or adding electrolyte additives for other cathode materials in the past, such as LiFePO4,58,59 and we expect that further studies will be successful in preventing this issue and improving the long term cycling properties of Na2Fe2OS2.
The Na2Fe2OS2 material gives a theoretical capacity of 225 mA h g−1, 75% of which is obtained on the first charge to 3 V. The reversible capacity is 162 mA h g−1, which corresponds to an approximated volumetric capacity of 552 mA h cm−3 (considering the density of the crystallized Na2Fe2OS2 phase). For comparison, materials considered as potential cathode material for Na-ion batteries such as Na3V2(PO4)2F3 and Na2/3(Fe1/2Mn1/2)O2 have volumetric capacities of 349 and 570 mA h cm−3, respectively.2 In that regard, the performances of Na2Fe2OS2 fall within the right range for commercial application (cf. Table S15†).
In contrast to the parent material Na2Fe2OSe2, which can easily be prepared by a solid-state route between 600 and 700 °C,14 Na2Fe2OS2 is metastable and decomposes at high temperature. An interesting point of comparison is the metastable anti-perovskite phase Li2FeOS, whose synthesis requires quenching from 750 °C to room temperature.13 A lower temperature synthesis route, via mechanochemistry, was developed to obtain the kinetically-stabilized Na2Fe2OS2 phase. Mechanosynthesis has been shown to be a powerful tool for the synthesis of sulphide and oxysulphide materials in the past.70–72 This method has the advantage of operating at room temperature and ambient pressure, which leads to easy processing/up-scaling and enables the formation of metastable phases with high doping concentration for instance.25,45 The stability and conceivable synthesis of Na2Fe2OS2 under dry air will also facilitate its handling for any future application. The study of the microstructure of the material showed that it is composed of an amorphous part and a crystalline part, having coherently scattering domains of 20 nm, both of composition Na2Fe2OS2.
Using X-ray and neutron diffraction data, the crystalline component was fully characterized. Na2Fe2OS2 is the n = 1 member of the anti-Ruddlesden–Popper [SNa][SO(Fe2Na)]n homologous series, with divalent iron and two sodium ions. The structural evolution of the “anti” structure Na2Fe2OS2 upon desodiation can be compared to the topochemical oxygen deintercalation from the “normal” Ruddlesden–Popper phase Ca2MnO4 to form the oxygen deficient Ca2MnO3.5 phase under reducing conditions (Fig. 8).73,74 In the oxygen deficient RP phase, oxygen vacancies are preferentially found in the [MnO2] square plane of the RP phase, with an ordering that depends on the nature of the A cation. In the sodium deficient anti-RP structure, with refined composition Na1.6(1)Fe2.0(1)OS1.98(10), the sites left vacant by the deintercalated Na atoms are filled by the migration of Fe atoms, so that the vacancies preferentially sit in the [Fe2O] plane. Interestingly, the number of vacancies in Na1.6(1)Fe2.0(1)OS1.98(10) is close to that found in the Ca2MnO3.5 phase, but they are disordered over all the cation sites, a difference that could simply arise from the presence of two different cations Fe2+ and Na+ in the anti-RP phase, as opposed to only one anion O2− in the RP phase. Furthermore, the displacement of Fe atoms from the 4c to 16n position takes after the displacement of some O atoms into tetrahedral OCa2Mn2 sites for the oxygen deficient RP phase, in order to compensate for the loss of oxygen anions and reduce the repulsion between cations. Further reduction of the RP phase leads to a loss of cationic order to form the disordered rocksalt Ca2MnO3, whereas it is possible to oxidize the anti-RP phase NaxFe2OS2 to x = 0.5, which leads to the complete amorphisation of the material. Fe atoms in the amorphous phase formed are coordinated by O and/or S atoms, with Fe–O/Fe–S distances consistent with a tetrahedral environment, e.g. FeO2S2. As the loss of crystallinity of the material is triggered by the favorable vacancy formation on the Fe site of the [Fe2O] square plane, the question arises how to stabilize this layer and avoid or delay the amorphisation of the material. A possible answer could come from the study of the n = 2 members of the “normal” Ruddlesden–Popper series. By moving from Y3+ cations in YSr2Mn2O7 to a larger Nd3+ cation in NdSr2Mn2O7, the oxygen vacancies in the corresponding (Y/Nd)Sr2Mn2O5.5 phases are displaced from the [MnO2] planes to the apical oxygen sites corresponding to the Na sites in the anti-structure of Na2Fe2OS2. Following this reasoning, using larger Se2−/Te2− anions instead of S2− could stabilize the vacant sodium sites, thus preventing the migration of Fe atoms and the loss of crystalline order, or at least delay it to higher desodiation levels.
This new structure type among cathode materials paves the way for the exploration of new phases with other transition metal cations and/or higher order members of the series. A few n = 1 compounds have been reported in the anti-RP series,14,44 but the only higher n member reported is the n = 3 ((Ba,Sr)3n+1ONn−1)Bin+1 anti-RP compound.54 The theoretical capacity is expected to decrease with increasing values of n (191 mA h g−1 for n = 2, 178 mA h g−1 for n = 3 and converging to 150 mA h g−1 for n = ∞, corresponding to the anti-perovskite NaFe2OS) but these values remain attractive and higher order structures could present improved structural stability with cycling or other more valuable features compared to Na2Fe2OS2. It is interesting to note that the previously reported anti-perovskite Li2FeOS, which has a much higher theoretical capacity (455 mA h g−1), is the end member of another anti-RP series [FeS][Li2FeOS]n, and shows more promise to increase the capacity. We were not able to stabilize the Na equivalent Na2FeOS phase, or alternatively the anti-RP Li2Fe2OS2 phase. The relative size of cations and anions seems to play a role in the stabilization and local structural distortions of anti-perovskite and anti-Ruddlesden–Popper phases, in the same way it does for perovskite and Ruddlesden–Popper phases, which are prone to distortions and octahedral tilting that have been extensively studied.75 The larger difference in ionic radii between sodium (r(Na+, [VI]) = 1.16 Å) and iron (r(Fe2+, [VI]) = 0.78 Å) cations is well suited to the anti-RP phases with two distinct cations sites corresponding to the axial and equatorial anion sites in the RP structure itself, whereas iron and lithium (r(Li+, [VI]) = 0.76 Å) have similar radii that favor mixed occupancy on the unique cation site of the anti-perovskite structure. Substituting iron with earlier transition metals (Ti, V, Cr, Mn), which have larger ionic radii, could help in stabilizing both Na anti-perovskite and Li anti-RP structures by decreasing/increasing the difference in ionic radii, respectively.
Despite the full amorphization of the phase, a good capacity is obtained and maintained throughout cycling. The charge compensation mechanism is expected to involve both Fe and S species and remains to be unraveled. In that regard, the disentanglement of redox contributions will represent a fascinating challenge. Analysis of EXAFS data at the Fe K-edge enabled us to understand the local structure surrounding Fe. It showed that the oxysulphide phase is maintained both in the amorphous part of the pristine sample and throughout the charge–discharge process during cycling, with Fe being coordinated to both S2− and O2− anions. This feature could help maintain the electrochemical properties of the material upon long-term cycling and shows that the crystallinity is not a prerequisite for good cyclability in this material. Amorphization has been shown to play a crucial role in promoting the electrochemical performance of Na–Fe–P–O based cathodes,76,77 and, in this regard, it would also be interesting to further study sodium oxysulphide glass materials for Na-ion batteries. Oxysulphides represent a very promising opportunity as cathode materials for sodium-ion batteries, with the possibility for both transition metal cations and sulphur anions to participate in the redox process to enhance overall capacity.
Transmission Electron Microscopy (TEM) and energy dispersive X-ray spectroscopy (EDX) mapping was performed with a JEOL 2100+ equipped with an EDX detector from Oxford instruments. Powder sample was ball milled in methanol under inert atmosphere to reduce particles size. Few drops of the methanol suspension were deposited on a holey carbon film Ni TEM grid. The grid was loaded into a beryllium TEM holder and inserted in the JEOL 2100+. In order to avoid exposing the material to air the sample preparation and loading was perform under inert atmosphere.
The program FullProf was used to obtain information about the microstructure, following the method described by Rodriguez-Carjaval et al. This model uses the Scherrer formula, which considers that the size broadening can be written as a linear combination of spherical harmonics. Peak shapes were modelled using the spherical harmonics expansions36 in a tetragonal material with Laue class 4/m. The microstructural information given in the output files were then treated using GFOURIER program81 to visualize the particle shapes.
The amount of amorphous content was determined by using the quantitative phase analysis method described in the literature.82–84 This method consists of mixing a known amount of the Na2Fe2OS2 material with an internal standard of similar absorption coefficient, Na3Fe2S4, and to compare the relative weight percent calculated by the program with the relative weight percent determined experimentally, which includes the contribution of amorphous phases. This method is described in the ESI, part 2.3.4.†
Footnotes |
† Electronic supplementary information (ESI) available: Synthesis details for the solid state synthesis and mecanosynthesis of Na2Fe2OS2, details of the different synthesis attempts and of the optimisation and washing procedure, morphology of the sample prepared by mechanosynthesis, result of the elemental analysis, details of the Rietveld refinement and structure outcomes of the solid state and mechanosynthesis samples before and after washing, bond distance and angles in Na2Fe2OS2-MW, XRD results of the stability test, in situ XRD of Na2Fe2OS2-MW, EXAFS fits of the ex situ samples and outcome of the refinement, details of the structure determination and outcome of the Rietveld refinement of Na1.7Fe2OS2. CIF file for Na2Fe2OS2. Associated structural information of Na2Fe2OS2. CCDC 1990654. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d0ta07966a |
‡ These authors contributed equally to the work. |
This journal is © The Royal Society of Chemistry 2020 |