2022 Volume 90 Issue 3 Pages 037011
All-solid-state sodium secondary batteries are expected to be low-cost, next-generation batteries. NiS2 and FeS2 are potential candidates as positive electrode materials owing to their high theoretical capacities. However, it is difficult to achieve sufficient capacity with bulk FeS2. In this study, pyrite Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) electrodes are prepared by a mechanochemical process. The all-solid-state sodium cells with Ni1−xFexS2 show higher discharge–charge potentials than those with NiS2, and higher capacities than those with FeS2. In addition, Ni1−xFexS2 exhibits a higher rate performance than those of NiS2 and FeS2. The all-solid-state cells using Ni1−xFexS2 (x = 0.3, 0.5, and 0.7) are discharged and charged with a high capacity of approximately 390 mAh g−1, without significant capacity fading for at least 30 cycles. The solid-solution formation of NiS2 and FeS2 results in lower material cost, higher rate performance, higher discharge–charge potential than those of NiS2, and higher capacity than that of FeS2. Pyrite Ni1−xFexS2 is a promising positive electrode material for all-solid-state sodium secondary batteries.
Conventional lithium-ion batteries (LIBs) are widely used in smartphones and other devices. From a safety standpoint, all-solid-state batteries are considered as next-generation batteries owing to the use of non-flammable inorganic solid electrolytes. Furthermore, all-solid-state sodium secondary batteries are expected to be low-cost batteries due to the abundance of sodium sources.1,2
It is necessary to develop a positive electrode active material with a high capacity and potential to obtain an all-solid-state sodium secondary battery with a high energy density. Sulfur is a promising candidate as a high-capacity material due to its high theoretical capacity (1672 mAh g−1). However, it is challenging to use sulfur directly as an active material, because it has a low electrical conductivity of 5 × 10−30 S cm−1 at 25 °C.3 The preparation of a composite that includes conductive carbon is an effective strategy for increasing electron conduction. In fact, our research group reported that a sulfur-carbon composite showed a high reversible capacity in all-solid-state sodium secondary batteries.4
In addition to sulfur-carbon composites, transition metal sulfides are promising electrode materials for high-energy batteries because of their high theoretical capacities and electrical conductivities. Among them, transition-metal polysulfide shows a high capacity derived from the anionic redox reaction involving dissociation and formation of disulfide bonds.5 We also reported that amorphous MoS3 showed a high capacity in all-solid-state sodium secondary batteries.6 In addition, metal sulfides with pyrite-type structures that have disulfide bonds (e.g., NiS2, FeS2, and CoS2) are expected to be high-capacity materials. In particular, NiS2 is expected to be a promising electrode material for all-solid-state sodium secondary batteries because it has been reported that NiS2 shows a high capacity in sodium-ion batteries using organic electrolytes.7,8 FeS2 is also a promising candidate because of its abundance in the Earth’s crust, low material cost, and high theoretical capacity.9–11 Additionally, FeS2 showed a higher discharge–charge potential than that of NiS2, as shown in this study. However, it is difficult to achieve sufficient capacity using bulk FeS2 due to the slow diffusion of Na and Fe in FeS2, which decreases the utilization of FeS2 during the discharge–charge reaction.12
In this study, we focused on the solid solutions Ni1−xFexS2 prepared from NiS2 and FeS2. We aimed to lower the cost and improve the discharge–charge potential of NiS2, and improve the capacity of FeS2, by combining NiS2 and FeS2. Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) were prepared by a mechanochemical (MC) process. MC processes are often used for the synthesis of metal sulfides.6,13–16 The structures and electrochemical properties of Ni1−xFexS2 electrodes in all-solid-state sodium secondary batteries were evaluated.
Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) samples were prepared using an MC process. The starting materials, Ni (99.99 %, Aldrich), Fe (99.9 %, Kojundo Chemical), and S (99.9 %, Kojundo Chemical) were mixed in a molar ratio of 1 − x : x : 2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) to obtain a total mass of 0.5 g. The mixtures were placed in a 45 mL zirconia pot with zirconia balls (90 g, diameter 4 mm) in a dry Ar glove box and mechanically milled using a planetary ball mill apparatus (Pulverisette 7, Fritsch). The rotational speed and milling time were set to 510 rpm and 10 h, respectively.
2.2 Material characterizationX-ray diffraction (XRD) measurements were performed using an X-ray diffractometer (SmartLab, Rigaku) equipped with a one-dimensional X-ray detector employing CuKα radiation (λ = 1.54 × 10−10 m). CHNS elemental analysis was performed using an elemental analyzer (Vario EL cube, Elementar). The surface of the Ni1−xFexS2 particles was analyzed using a field-emission scanning electron microscope (FE-SEM, SU8200; Hitachi High-Technologies Corp.). The electronic structures of sulfur in Ni1−xFexS2 were analyzed using X-ray photoelectron spectroscopy (XPS) (K-Alpha, Thermo Fisher Scientific) with a monochromatic AlKα source (1486.6 eV). The observed binding energies were calibrated against the adventitious C1s peak at 284.7 eV.
2.3 Electrochemical measurementsThe all-solid-state cells for discharge–charge measurements were fabricated as follows: composite positive electrodes were prepared by mixing Ni1−xFexS2, Na3PS4 glass-ceramic solid electrolyte (SE),17 and acetylene black (AB) in a mass ratio of 40 : 60 : 6 in an agate mortar. SE and AB provided sodium-ion and electron conduction paths to the active material in the composite positive electrodes. In this study, a sufficient amount of SE (60 wt%) was also added to the composite positive electrodes to evaluate the potential performance of Ni1−xFexS2 as an active material. Tri-layer pellets (10 mm in diameter) consisting of a Ni1−xFexS2-Na3PS4-AB positive electrode layer (6 mg), a Na3PS4 glass-ceramic separator layer (80 mg), and a Na15Sn4-Ketjen black (KB) negative electrode layer (40 mg)18 were prepared by uniaxial pressing under 360 MPa for 5 min in a polycarbonate cylinder with stainless steel rods acting as current collectors. All preparation steps were conducted in a dry-Ar-filled glovebox. The discharge–charge tests were conducted in a glovebox at 24–27 °C with a cut-off voltage range of 0.92 to 2.72 V vs. Na15Sn4 using an electrochemical workstation (VMP3, Bio-Logic Co.).
Figure 1a shows the XRD patterns of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1). Figure 1b and Table 1 show the lattice constants determined from the XRD patterns of Ni1−xFexS2. All compositions showed similar XRD patterns with broad peaks, and the diffraction patterns of Ni1−xFexS2 could be indexed by the pyrite structure of the space group Pa-3 (No. 205). On the XRD patterns of Ni0.3Fe0.7S2, Ni0.1Fe0.9S2, and FeS2, the small peaks derived from the impurity phase Fe7S8 were observed. Figure 1c shows the structural model of the pyrite. All sulfur atoms in the pyrite-type structure formed disulfide bonds. With an increase in the Fe content of Ni1−xFexS2, the diffraction pattern of Ni1−xFexS2 shifted to higher angles and the lattice constant of Ni1−xFexS2 decreased, suggesting that solid solutions of Ni1−xFexS2 were obtained.
(a) XRD patterns of Ni1−xFexS2. (b) The lattice constants for x in Ni1−xFexS2. (c) The structural model of the pyrite.
Lattice constants (Å) | |
---|---|
NiS2 | 5.681 (7) |
Ni0.7Fe0.3S2 | 5.602 (7) |
Ni0.5Fe0.5S2 | 5.552 (3) |
Ni0.3Fe0.7S2 | 5.516 (17) |
Ni0.1Fe0.9S2 | 5.456 (3) |
FeS2 | 5.426 (3) |
Figures 2a–2f show SEM images of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1). The particle sizes of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) samples ranged from the submicron level to a few microns, and there were no significant differences in their sizes. From the EDS mapping results of Ni0.3Fe0.7S2 (Fig. S1), it was found that the constituent elements were basically uniformly distributed.
(a–f) The SEM images of Ni1−xFexS2.
Figure 3a shows the S2p XPS spectra of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1), which were composed of two sets of doublet peaks. Each doublet comprises the S2p3/2 and S2p1/2 components. The peak energy separation (1.2 eV) and the area ratio of S2p3/2/S2p1/2 = 2/1 for each doublet were set during peak fitting. The S2p spectra of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) showed one S2p3/2 peak at 162.7–163.0 eV. This is because there is only one type of sulfur in the pyrite-type structure, which forms disulfide bonds. Figures 3b and 3c show the Ni2p and Fe2p XPS spectra of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1), where a Ni2p3/2 peak was observed at 853.2–853.4 eV and an Fe2p3/2 peak at 707.0–707.3 eV. The broad peak near 711 eV in Fig. 3c is attributed to the Auger line of Ni KL1L1. These results indicate that the electronic states of the Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) samples were similar. In all compositions, the electronic states of Ni and Fe are similar because Ni and Fe exist in the cationic sites in the pyrite structure.
(a) S2p (b) Ni2p and (c) Fe2p XPS spectra of Ni1−xFexS2.
Figures 4a–4f show the discharge–charge curves of the all-solid-state cells fabricated using Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) as the positive electrode active material, respectively. The discharge–charge measurements were performed at a current density of 0.013 mA cm−2. The discharge–charge potentials of Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) were higher than that of NiS2 and lower than that of FeS2. Thus, the discharge–charge profiles of Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) were intermediate between those of NiS2 and FeS2. The plateaus observed during the first discharge of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) were attributed to the structural changes associated with the initial Na ion insertion. With the insertion of Na during the initial discharge, the crystal structure changes and becomes amorphous. The crystal structure does not return to its original state even after the Na is removed during charging. The charging and discharging continue in an amorphous state.19 The reversible capacities of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) were 492, 485, 411, 390, 418, and 227 mAh g−1, respectively. NiS2 showed irreversible capacity in the first cycle, and then the capacity decreased slightly over five cycles. In contrast, Ni1−xFexS2 (x = 0.3, 0.5, 0.7, 0.9, and 1) showed reversible capacities in the first cycle, and no decrease in capacity with each cycle. Furthermore, the potentials of Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) increased compared to those of NiS2. This was attributed to the combination of the characteristics of NiS2 and FeS2. Here, the reaction equation between Na and MS2 (M: Ni, Fe) is as follows:19 MS2 + 2Na+ + 2e− $ \rightleftarrows $ Na2MS2. This reaction is an anion redox reaction in which the disulfide bond ((S-S)2−) dissociates and changes to S2−, unlike the intercalation reaction of the layered compound. The anion redox reaction of Ni1−xFexS2 is a two-electron reaction with a theoretical capacity of approximately 440 mAh g−1. The capacity of 227 mAh g−1 for the all-solid-state cell using FeS2 was clearly lower than the theoretical capacity. The electronic conductivity of Ni1−xFexS2 too high to obtain an accurate value. Therefore, the influence of electronic conductivity on the difference in capacity and overvoltage is almost negligible, and factors such as sodium diffusion during the discharge-charge are considered to be more influential. Additionally, it has been reported that bulk FeS2 does not exhibit sufficient capacity.12,20 This is considered to be due to the slow diffusion of Na and Fe in FeS2; controlling the particle size at the nanoscale is one method for solving this problem.12 In fact, FeS2 (99.9 %, Aldrich) with a controlled particle size at the nanoscale showed a higher capacity at 298 mAh g−1 (Fig. S2) than FeS2 prepared by the MC process. Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9), in which a portion of Fe in FeS2 was replaced with Ni, showed capacities close to the theoretical capacity.
(a–f) Discharge-charge curves of Na15Sn4/Na3PS4/Ni1−xFexS2 in all-solid-state cell.
Figure 5a shows the rate performance of the all-solid-state cells using Ni1−xFexS2 (x = 0, 0.5, and 1) as the positive electrode active materials. The discharge–charge measurements were performed with current densities of 0.013, 0.038, 0.064, 0.13 and 0.19 mA cm−2. Both NiS2 and FeS2 showed low capacities of approximately 80 mAh g−1 and 25 mAh g−1 at 0.13 mA cm−2 and 0.19 mA cm−2, respectively. Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) showed a high rate performance compared to those of NiS2 and FeS2. In particular, Ni0.5Fe0.5S2 showed high capacities of 256 and 162 mAh g−1 at 0.13 and 0.19 mA cm−2, respectively. It was found that the rate performance of Ni1−xFexS2 was improved by substituting a portion of Ni in NiS2 with Fe (or Fe in FeS2 with Ni) compared to those of NiS2 and FeS2. Recently, it was reported that Co substitution into FeS2 enhances Na diffusion,21 and it was considered that a similar phenomenon occurs in Ni1−xFexS2 in this study. However, the detailed mechanism of Na diffusion enhancement has not been revealed.
(a) Rate performance of Ni1−xFexS2 in all-solid-state cell at 0.013–0.19 mA cm−2 for 0.92–2.72 V vs. Na15Sn4. (b) Cyclic performance of Ni1−xFexS2 in all-solid-state cell at 0.038 mA cm−2 for 0.92–2.72 V vs. Na15Sn4.
Figure 5b shows the cycle performance of the all-solid-state cells using Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) as the positive electrode active materials. The discharge–charge cycle measurements were performed at a current density of 0.038 mA cm−2. The capacities at the 30th cycle of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) were 370, 381, 385, 395, 313, and 97 mAh g−1, respectively, while the capacity retention rates of Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1) were 81.9 %, 86.5 %, 89.8 %, 93.6 %, 85.0 %, and 78.1 % between the 2nd and 30th cycles, respectively. NiS2 and FeS2 showed a decrease in capacity in the first few cycles, but then showed a stable capacity up to 30 cycles. Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) had improved capacity degradation results in the first few cycles compared to those observed in NiS2 and FeS2; Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9), and did not show significant capacity fading for at least 30 cycles. This indicated that Ni1−xFexS2 exhibited a high cycle performance for 30 cycles in all-solid-state sodium cells. The Ni1−xFexS2 (x = 0.3, 0.5, and 0.7) samples, which have higher discharge–charge potentials than that of NiS2, showed almost the same capacities as that of NiS2 after 30 cycles. These results show that Fe substitution in NiS2 can reduce the material cost and improve the energy density compared to those of NiS2.
Pyrite Ni1−xFexS2 (x = 0, 0.3, 0.5, 0.7, 0.9, and 1), which is a solid solution of NiS2 and FeS2, was prepared by the MC process. The XRD results suggested that a solid solution of Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) was obtained. The all-solid-state sodium cells with Ni1−xFexS2 (x = 0.3, 0.5, 0.7, and 0.9) positive electrodes showed high reversible capacities comparable to those of the cell with NiS2. Ni1−xFexS2 showed a higher discharge–charge potential than that of NiS2. Ni1−xFexS2 also showed a higher rate performance than those of NiS2 and FeS2. In particular, Ni0.5Fe0.5S2 showed high capacities of 485, 256 and 162 mAh g−1 at 0.013, 0.13, and 0.19 mA cm−2, respectively. In addition, Ni1−xFexS2 exhibited an excellent cycle performance for 30 cycles in all-solid-state sodium cells. The solid-solution formation of Ni1−xFexS2 resulted in lower material cost, higher rate performance, and higher discharge–charge potential than those of NiS2, and higher capacity than that of FeS2. Furthermore, it could be determined that the formation of Ni1−xFexS2 solid solutions is advantageous for Na ion diffusion due to the combination of different discharge–charge potentials. This study suggests that Ni1−xFexS2 is a promising positive electrode material for all-solid-state sodium secondary batteries.
This study was supported by JSPS KAKENHI Grant Numbers JP20K05688 and JP21H04701.
The data that support the findings of this study are openly available under the terms of the designated Creative Commons License in J-STAGE Data at https://doi.org/10.50892/data.electrochemistry.19106876.
Gaku Shirota: Data curation (Lead), Formal analysis (Lead), Writing – original draft (Lead)
Akira Nasu: Data curation (Supporting), Investigation (Supporting), Writing – review & editing (Supporting)
Atsushi Sakuda: Conceptualization (Lead), Project administration (Lead), Supervision (Equal), Writing – review & editing (Lead)
Minako Deguchi: Data curation (Supporting), Formal analysis (Equal)
Kota Motohashi: Supervision (Supporting), Writing – review & editing (Equal)
Masahiro Tatsumisago: Supervision (Equal), Writing – review & editing (Supporting)
Akitoshi Hayashi: Funding acquisition (Equal), Supervision (Equal), Writing – review & editing (Equal)
The authors declare no conflict of interest in the manuscript.
Japan Society for the Promotion of Science: JP20K05688
Japan Society for the Promotion of Science: JP21H04701
A. Nasu, A. Sakuda, K. Motohashi, and A. Hayashi: ECSJ Active Members
M. Tatsumisago: ECSJ Fellow