Open Access Article
K. Okada
ab,
I. Kimuraa and
K. Machida*a
aDivision of Applied Chemistry, Graduate School of Engineering, Osaka University, Suita, Osaka 565-0871, Japan. E-mail: machida@chem.eng.osaka-u.ac.jp
bDepartment of Materials Science, Graduate School of Engineering, Osaka Prefecture University, Sakai, Osaka 599-8531, Japan
First published on 6th February 2018
Enhanced electrochemical performance of LiFePO4 for Li-ion batteries has been anticipated by anion doping at the O-site rather than cation doping at the Fe-site. We report on the electrochemical performance of S-doped LiFePO4 nanoparticles synthesized by a solvothermal method using thioacetamide as a sulfur source. S-doping into the LiFePO4 matrix expands the lattice due to the larger ionic radius of S2− than that of O2−. The lattice parameters a and b increase by around 0.2% with sulfur content, while that of c remains almost unchanged with only 0.03% increase. The S-doping also contributes to the suppression of antisite defects (Fe occupying Li sites), which facilitates the easy migration of Li in the diffusion channels without blockage. Owing to these effects of S-doping, the S-doped LiFePO4 nanoparticles show enhanced electrochemical properties with a high discharge capacity of ∼113 mA h g−1 even at a high rate of 10C.
Olivine-structured LiFePO4 is considered as one of the most promising cathode materials for LIBs due to its high thermal stability, good cycling property, high energy density, theoretical capacity, and environmental friendliness as well as low cost.6,7 Although the theoretical capacity (170 mA h g−1) of LiFePO4 is a bit lower than that of conventional cathode materials (e.g., LiCoO2: 274 mA h g−1), its benefits such as low toxicity, high safety, low-cost and good cycling property will make LiFePO4 replaced to conventional cathode materials as now being produced commercially.8,9 However, its poor electronic conductivity (10−8 to 10−10 S cm−1) and low lithium ion diffusion coefficient (10−10 to 10−16 cm2 s−1) lead to a decreased capacity at a high charging/discharging rate, which has inhibited the use in high-rate applications.10–13 Many efforts have been made to solve the problems by mainly three strategies: (1) carbon coating,14,15 (2) reducing particle size,16,17 (3) doping with certain element.18–20 Excellent electrochemical properties have been reported by combining these three strategies. The carbon coating can increase the surface electronic conductivity, leading to an improvement of rate performance and cycling life. The method of carbon coating has been developed and established already as a thermal treatment of the mixture of LiFePO4 and carbon sources under inert gas. Reducing particle size can shorten the Li+ diffusion pathway length, that is an effective way to improve rate capability. Hydro/solvo-thermal synthesis approaches have been applied to obtain the LiFePO4 nanoparticles. As for the doping strategy, doping proper ions into the LiFePO4 can increase the electronic and ionic conductivity of the LiFePO4 and also expand lithium ions diffusion channels in the structure, leading to a significant improvement of the rate capability and cyclic performance. Various elements such as La, Cu, Y, V, Mn, Mg, Bi, Co, Pt, Pd, Ni, Zn, Mo and Cr have been previously used as dopant elements in cation-site (Fe-site) of LiFePO4. Certain cation doping elements like Ni and Zn can remarkably enhance the electrochemical performance of LiFePO4.21–24 There are much less reports on anion doping at O-site for LiFePO4. However, the O-site doping is expected to significantly enhance the conductivity of LiFePO4 rather than Fe-site doping as indicated by first-principle calculations.25,26 Indeed, significantly enhanced high rate capability and cycling performance have been reported for anion-doped (such as F and Cl) LiFePO4.27–29 Among considerable anion species, sulfur doping would be efficient to improve the electrochemical performance (rate capability and cyclic performance) because easy Li+ intercalation and extraction are expected by an expansion of lattice arising from larger anion doping and the lower bond dissociation energy of Li–S (312.5 kJ mol−1) than that of Li–O (340.5 kJ mol−1).30 Sulfur-doping has been only reported by Park et al. where S-doped LiFePO4 nanoparticles exhibit enhanced electrochemical properties compared to non-doped and N-doped LiFePO4.31 However, only surface substitution of LiFePO4 nanoparticles was achieved by exposing solvothermally-synthesized LiFePO4 nanoparticles to sulfur vapor at high temperature. The further enhancement of electrochemical properties can be expected for the sulfur doping into the LiFePO4 matrix.
Herein, we report on the synthesis of S-doped LiFePO4 nanoparticles by a single-step solution approach and its improved electrochemical performance compared to non-doped LiFePO4. Thioacetamide was used as a sulfur source, which was the conventional way for the synthesis of S-doped TiO2 and metal sulfide such as CdS, ZnS and Bi2S3 by the solution approach.32–35 The successful S-doping into LiFePO4 matrix was confirmed by X-ray diffraction (XRD), energy dispersive X-ray spectrometry (EDS) analysis and X-ray photoelectron spectrometry (XPS). The formation of LiFePO4 nanoparticles with ∼100 nm in diameter was observed by scanning electron microscopy (SEM). The S-doped LiFePO4 nanoparticles showed a high discharge capacity of ∼113 mA h g−1 even at a high rate of 10C.
:
Fe
:
P was 2.7
:
1
:
1. The reaction was carried out in an autoclave at 180 °C for 10 h. After the reaction, the resultant gray precipitate was washed with deionized water and ethanol for several times and then dried at 70 °C overnight under vacuum. S-doped LiFePO4 nanoparticles were prepared by adding thioacetamide with different amount (0.13, 0.40, 0.67, 1.33, 2.66 and 3.99 mmol) to the reaction solution containing Li, Fe and P sources before solvothermal treatment (the molar ratio; Li
:
Fe
:
P
:
S = 2.7
:
1
:
1
:
0.02, 0.07, 0.11, 0.22, 0.44 or 0.67). In the manuscript, the non-doped and S-doped LiFePO4 nanoparticles prepared with different amount of thioacetamide were denoted as LFP and LFP-S-x (where x = 0.02, 0.07, 0.11, 0.22, 0.44 or 0.67), respectively. For the electrochemical measurements, the dried samples with a certain amount of sucrose (10 wt%) were mixed in ethanol and calcined at 650 °C for 6 h under N2 atmosphere in order to obtain the carbon-coated samples.
:
2.5
:
1) in N-methyl-2-pyrrolidone (NMP) solution. Then, the slurry was coated on an Al foil by using a coater. After being dried under vacuum at 90 °C for 10 h, working electrodes were punched and weighed. Then, the cells containing 1.0 mol L−1 LiPF6 in ethylene carbonate (EC)-diethyl carbonate (DEC) (1
:
1, v/v) as electrolyte were assembled in a glove box under a dry and high purity argon atmosphere. Charge/discharge tests were performed between 2.0 and 4.3 V by using a battery tester (Kikusui, PFX2011, Japan) at room temperature at around 20 °C.![]() | ||
| Fig. 1 (a) XRD patterns of LFP and LFP-S-x (where x = 0.02, 0.07, 0.11, 0.22, 0.44 or 0.67) (silicon was used as a standard). (b) Changes in a and c lattice parameters of LFP-S-x (x: from 0 to 0.67). | ||
Fig. 2 shows the SEM images of LFP, LFP-S-0.22 and LFP-S-0.67. Uniform nanoparticles with ∼100 nm in diameter were observed for LFP (Fig. 2(a)). Similar nanoparticles formed in LFP-S-0.22 (Fig. 2(b)), indicating S-doping did not affect the morphology. In the LFP-S-0.67, larger and crystalline particles over 1 μm were observed with the presence of LiFePO4 nanoparticles (Fig. 2(c) and (d)). The large particles were concluded as an impurity FeS2 by the XRD result and EDS analysis (Fig. S2†). EDS mapping and analysis for S-doped LiFePO4 (LFP-S-0.22) confirmed that sulfur was uniformly dispersed as an atomic percent of ∼1.4% (Fig. 3 and S3†). Based on this result, the composition of LFP-S-0.22 was estimated as LiFePO3.9S0.1.
![]() | ||
| Fig. 3 Secondary electron image (SEI), EDS mapping images and line scan profile of LFP-S-0.22 for P, O, Fe and S. The line scan profile was collected in the red dot line of the SEI image. | ||
The chemical states of sulfur, phosphorous and oxygen ions were investigated by XPS (Fig. 4). Fig. 4(a) shows the XPS spectra for the S 2p electrons of LFP, LFP-S-0.22 and LFP-S-0.67. A broad peak at around 160–163 eV was confirmed in the LFP-S-0.22 although no peaks was detected in LFP. The peak was fitted as the core levels of S 2p1/2 (at 162.4 eV) and S 2p3/2 (at 161.4 eV) separated by a spin–orbit splitting of 1.0 eV, demonstrating that the sulfur atoms are doped in the state of S2−.48,49 On the other hand, a broad peak at ∼1.0 eV higher than that of LFP-S-0.22 was detected in LFP-S-0.67. The peak at higher energy could be assigned to that of disulfide derived from FeS2.50,51 The presence of FeS2 in LFP-S-0.67 was consisted with the XRD results. The XPS spectra for the P 2p electrons of LFP, LFP-S-0.22 and LFP-S-0.67 are shown in Fig. 4(b). As shown in Fig. 4(b), the sulfur doping shifted the P 2p XPS peak to lower binding energy. Similar chemical shift of binding energy has been reported for the nitrogen doping at Li3PO4 and Na3PO4 where a decrease in the P 2p binding energy was explained by simple charged-shell models52,53 as reducing the average ionic charge on the phosphorus ions due to the replacement of P–O bonds with less ionic P–N bonds.54,55 The binding energy chemical shift by sulfur doping to LiFePO4 can be explained by the similar effect; reducing the average ionic charge on the phosphorus ions due to the replacement of P–O bonds with less ionic P–S bonds results in the lower P 2p binding energy. O 1s XPS peaks in LFP-S-0.22 also showed lower binding energy compared to LFP, which is similar to N-doped Li3PO4 and Na3PO4 (ref. 54) (Fig. S4†). These XPS results for the S 2p, P 2p and O 1s electrons clearly indicate the successful S2− doping at O-site of LiFePO4 matrix.
Electrochemical measurements were carried out to test the electrical properties of LFP and LFP-S-0.22 after carbon coating by following a reported method.56 Fig. 5(a) shows the charge/discharge curves of carbon-coated LFP and LFP-S-0.22 measured at 5C (850 mA g−1) between 2.0 and 4.3 V vs. Li+/Li. Both samples showed a charging and discharging plateau region at ∼3.4 V which is correspond to the redox reaction between FePO4 and LiFePO4.57 A smaller voltage difference between the charge and discharging plateaus was confirmed in LFP-S-0.22 compared to LFP, implying that LFP-S-0.22 possess the better electronic conductivity and higher reaction reversibility.58 The non-doped LFP presented a discharge capacity of 99.0 mA h g−1 at 5C. In contrast, S-doped LFP-S-0.22 led to a higher discharge capacity of 121.6 mA h g−1 at 5C. The discharge capacities of LFP and LFP-S-0.22 were investigated at different rates from 0.5C to 10C in order to examine the rate capabilities (Fig. 5(b)). The LFP-S-0.22 showed much higher discharge capacities (131.7, 128.5, 121.6 and 112.7 mA h g−1 at 0.5, 1, 5 and 10C, respectively) at each rates compared to LFP (120.6, 116.4, 99.0 and 81.1 mA h g−1 at 0.5, 1, 5 and 10C, respectively). In addition, the discharge capacity of LFP-S-0.22 varied much smaller from 128.5 to 112.7 mA h g−1 at rates of 1C to 10C, retaining 87.7% of discharge capacity. There was no difference between LFP and LFP-S-0.22 for coulombic efficiency at each C rate (Fig. S5†). These electrochemical investigations exhibited that sulfur doping into LiFePO4 effectively improved the rate capability. It has been considered that high rate capability can be achieved by an increase in electric conductivity of LiFePO4. The significantly enhanced high rate capability achieved in LFP-S-0.22 was considered due to an increase in electric conductivity by sulfur doping as expected by first-principle calculations.31 The differential capacity (dQ/dV) studies showed smaller peak voltage separation between the anodic and cathodic peaks in LFP-S-0.22 compared to LFP, indicating the smaller electrochemical polarization and better electrochemical kinetics of the S-doped LiFePO4 (Fig. S6†).59,60 Also, the LFP-S-0.22 exhibited higher internal electronic conductivity and lower resistance than LFP (Fig. S7†). In addition, the sulfur doping results in the suppression of
antisite defects (Fe occupying Li sites) in LiFePO4 which is important as it facilitates the easy migration of Li in the diffusion channels without blockage.61 The antisite defect is considered to be present in the one-dimensional channel of LiFePO4 and effectively block the Li+ pathways,62–64 which reduces the capacity, rate capability and cycle life of a battery. The suppression of
antisite defects can be investigated by Fourier transform infrared (FTIR) spectra.39,65 Fig. 6 shows the FTIR spectra of LFP and LFP-S-0.22. The Infrared absorption spectra in the range 800–1200 cm−1 and 400–700 cm−1 were assigned to the symmetric (ν1) and asymmetric (ν3) P–O stretching, symmetric (ν2) and asymmetric (ν4) O–P–O bending in (PO4)3− tetrahedral, respectively.66 As shown in Fig. 6, the sulfur doping shifted a peak at ∼975 cm−1 to lower wavenumber, while there were no significant changes for the other peaks. This result is clearly consistent with the
antisite defects-suppressed LiFePO4, suggesting sulfur doping into the LiFePO4 matrix can also suppress the formation of
antisite defects. In the LiFePO4 structure, lithium and iron ions are surrounded by six oxygen atoms and occupy the different octahedral interstitial sites; Li ions occupy edge-sharing sites (M1) and Fe ions occupy corner-sharing sites (M2). The
antisite defects reportedly form by cation intermixing (cation-exchanging) between the M1 and M2 sites.62 Our results indicate that the sulfur doping at O-site makes the Fe ions preferentially located in the M2 sites. In the reaction media, the sulfur prefers to react with iron ions as our XRD investigation revealed the formation of FeS2 with excess of sulfur sources and also the preferential formation of Fe–S bonds was reported in the sulfur-modified LiFePO4 on the surface by theoretical and experimental investigations.31 Then, the formed Fe–S species react with the H3PO4 by replacing O to S in the reaction media, which results in the preferential location of Fe in M2 sites by corner-sharing oxygen or sulfur atoms with PO4 tetrahedral. Consequently, the occupation of Fe in the M1 sites (
antisite defects) would be suppressed by sulfur doping at O-site. The suppression of
antisite defects by S-doping would also contribute to improvement in the capacity and the rate capability compared to non-doped LiFePO4. It should be also mentioned that the discharge capacity of 112.7 mA h g−1 at 10C achieved in LFP-S-0.22 is much higher than that of sulfur-doped LiFePO4 only on the surface (86.4 mA h g−1 at 10C) reported in the literature.31 Thus, it can be concluded that anion doping into the LiFePO4 matrix is more important than anion doping only on the surface of LiFePO4 particle in order to enhance the electrochemical performance. The synthesis of cation-doped LiFePO4 has been reported by similar solvothermal approaches.67,68 Thus, our system would be simply applied to the cation-doping approach. The further enhancement of electrochemical performance is expected by multi-elements co-doped LiFePO4 with certain elements.
![]() | ||
| Fig. 5 (a) Charge–discharge voltage curves of LFP and LFP-S-0.22 at 5C. (b) Rate performance of LFP and LFP-S-0.22. | ||
antisite defects in lithium ions diffusion channels. Both the lattice expansion and the suppression of
antisite defects by S-doping contribute to the enhancement of electrochemical properties due to the easy migration of Li in the diffusion channels without blockage.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c7ra12740e |
| This journal is © The Royal Society of Chemistry 2018 |