Electrolyte design for high-rate graphite anodes. The energy storage rate relies on how fast the Li+ can migrate between the cathode and the anode. In principle, Li+ undergoes three steps during charging processes (Fig. 1a): (a) diffussion of solvated Li+ in the bulk electrolyte, (b) desolvation of the solvated Li+ before crossing the SEI and (c) naked Li+ crossing the SEI11. These three steps are highly intertwined. The physicochemical properties of the solvents and concentration determine the diffusivity of the Li+, while the components of the Li+ inner solvation shell affect the SEI compositions and thereby plays a decisive role in the Li+ crossing kinetics in SEI. Accelerating the Li+ transport in all three steps will dramatically enhance the electrochemical kinetics, which calls for the electrolyte with the high ionic conductivity (IC), low desolvation energy (ΔEdsv) and low Li+ crossing energy-barrier SEI (ΔESEI) with a low area specific resistance (ASR). To design an electrolyte with these properties, the IC (Fig. 1b), ΔEdsv (Fig. 1c), and reduction potential (Gred, Fig. 1d) for the commonly used salts and solvents were firstly measured via experiments and DFT calculations (Fig. 1c,d). According to the results, FSI− anion exhibits the lowest ΔEdsv (Fig. 1c) and highest reduction potential among the anions (Fig. 1d), which indicates that FSI− anion can endow low desolvation kinetics and inorganic-rich SEI. For the solvents, AN based electrolytes show superior ionic conductivity to enhance the Li+ migration and transport inside the electrolyte (Fig. 1b), while DOL solvent which demonstrates the weakest interaction with Li+ (Fig. 1c) should be a good choice to improve the desolvation kinetics. FEC solvent is favorable for the LiF-rich SEI as it can promote the formation of LiF at around 1.0 V. LiF-rich SEI is very thin and has a low ASR because LiF has a high ratio of ionic conductivity to electronic conductivity. The corresponding Quantum Chemistry calculations of reduction potentials for Li-solvent/anion with different dielectric constants are shown in Supplementary Fig. 1.
We selected LiFSI as the salt because it has the best dissociation and highest potential to form LiF (Fig. 1d) among the studied common salts, and DOL as the solvent for the lowest ΔEdsv of Li+ (Fig. 1c). To maximize the rate performance, 1.8 M LiFSI DOL was blended since the conductivity maxima locates at this concentration (Supplementary Fig. 2b). The correlating solvation structures were examined by Raman spectroscopy (Supplementary Fig. 2c). The C-O-C bending peak ascribed to DOL is slightly perturbed (721 to 730 cm− 1) by the dissolution of LiFSI, while the other three C-O stretching peaks of DOL (940, 958 and 1088 cm− 1 ) keep intact37, implying weak coordination of DOL molecules bound by Li+, id est, a low ΔEdsv of Li+ in 1.8 M LiFSI DOL electrolyte.
Molecular dynamics simulation was performed to analyze the detailed solvation structures of the electrolytes (Supplementary Fig. 3). Unlike the strong coordination of Li+ with solvent in 1.0 M LiPF6 EC/DMC electrolyte, Li+ prefers to bond with FSI− in 1.8 M LiFSI DOL electrolyte as evidenced by a much higher peak of Li-OFSI than that of Li-ODOL. The corresponding pie chart (Supplementary Fig. 3d) indicates a low solvent number of 1.26, which means an easier Li+ desolvation process. Besides the electrolytes, the properties of graphites (morphologies and particle sizes) will also affect the rate performance38. In current work, the intercalation/deintercalation behaviors of Li+ in different graphites were evaluated via Electrochemical Impedance Spectroscopy (EIS) tests under 100% state-of-charge (SOC) to screen a suitable graphite anode (Supplementary Fig. 4). The total resistances (Rcell) are in the descending order of mesocarbon microbeads (MCMB) > artificial graphite (AG) > natural graphite (NG) in both two electrolytes (Supplementary Fig. 4a,b). Three natural graphites with different average diameters of 2.5, 10 and 15 µm are named as 2.5NG, 10NG and 15NG. With the smallest average diameter, 2.5NG holds the lowest Rcell (15 Ω) in 1.8 M LiFSI DOL electrolyte (Supplementary Fig. 4c), which is benificial for a fast kinetics, and thus is served as the anode material. The morphologies of different graphites were characterized by scanning electron microscopy (SEM), and no defects or cracks that contribute to a capacitive process are obeserved (Supplementary Fig. 5).
The fast kinetics for graphite anodes. The rate performance of microsized natural graphite (2.5NG) was evaluated in different electrolytes between 1.0 V and 0.0 V using a NG||Li half-cell configuration. The referred rate of nC meant a full charge or discharge in 1/n h. Compared to the rapid capacity decline starting from 1C in the carbonate electrolyte, NG anode in 1.8 M LiFSI DOL electrolyte yield a revolutionary breakthrough by providing capacities of 315 and 180 mAh g− 1 even at 20C and 50C (Fig. 2a), respectively. The recoverable capacity of 370 mAh g− 1 is obtained when the rate returned to 0.2C, indicating no damage to the graphite anode. The corresponding charge/discharge curves from 1C to 50C reveal the intercalation process of Li+ into graphite (Fig. 2b) with three voltage plateaus emerging below ~ 0.2 V at 1C for 1.8 M LiFSI DOL due to the formation of multistage structures of LiCx25. However, no potential plateau can be observed for charging/discharging of graphite in commercial carbonate electrolytes (1.0 M LiPF6 EC/DMC) at a high rate (Supplementary Fig. 6a). At a low rate (< 1C), the Li insertion kinetics is quasi-independent of the electrolytes, showing similar voltage profiles in 1.8 M LiFSI DOL and commercial carbonate electrolytes. As the rate increased to 5C and above, significant differences appear between cells with two electrolytes, particularly in voltage profiles. At such high rates, the battery with 1.8 M LiFSI DOL electrolyte shows a more defined potential plateau and a much lower overpotential in comparison with the commercial carbonate electrolyte. Even when the rate increased up to 20C, the potential plateaus are still defined for the cell with 1.8 M LiFSI DOL electrolyte (Fig. 2b, magenta line). Since the rate performance was measured in a two-electrode cell, the total overpotentials are comprised of two parts: overpotentials for graphite anodes and Li metal electrodes. The real overpotentials for graphite anodes (Fig. 2c) were obtained by subtracting Li metal overpotentials in the galvanostatic Li plating/stripping tests from total overpotentials of the above charge/discharge curves (Supplementary Fig. 6c). When C-rate increases from 1C to 50C, the overpotential of graphite electrode in 1.8 M LiFSI DOL electrolyte is from 0.032 to 0.153 V, which is much lower than that in the carbonate electrolytes (from 0.037 to 0.335 V) (Fig. 2c). Considering the narrow voltage gap (~ 0.1 V) between the Li+ intercalation/de-intercalation potential of graphite with Li metal potential, this variation poses a vital impact on the rate performance of graphite anode.
Galvanostatic intermittent titration technique (GITT)39,40 using a high rate (10C) was performed to gain insights into the overpotentials in lithiation/delithiation processes of graphite electrodes/LiCx. During the lithiation process, a capacity of > 360 mAh g− 1 is achieved in the 1.8 M FSI DOL electrolyte in contrast to that of 80 mAh g− 1 in the commercial carbonate electrolyte (Fig. 2d). The representative potential changes in lithiation and relaxation processes at open-circuit in the middle profiles during GITT experiments are shown in the inset. The whole potential rise in the commercial carbonate electrolyte (142 mV) is 4.6 times higher than that in 1.8 M LiFSI DOL electrolyte (31 mV). Similar phenomena were also observed in the delithiation process with overpotentials of 152 mV and 31 mV in the commercial carbonate electrolyte and 1.8 M LiFSI DOL electrolyte (Supplementary Fig. 7), respectively. The extremely large overpotential of commercial carbonate electrolyte should be attributed to the high resistance of the interphases and Li+ desolvation41, which leads to sluggish kinetics. At the same lithiation/delithiation state of graphite, the similar values of equilibrated potentials (61 vs. 67 mV, 9.3 vs. 12 mV, 3.4 vs. 10 mV) in two electrolytes (Supplementary Fig. 7c,d) indicates that the thermodynamic equilibrium state is independent of the electrolyte. Based on the fitting of Nyquist plot, the cell resistance mainly comprises of bulk resistance (Rb), surface layer resistance (Rsei) and charge-transfer resistance (Rct) (Fig. 2e)6,42. The much lower values of Rsei and Rct (9.1 and 4.7 Ω) than those of the carbonate electrolyte (33.2 and 15.1 Ω), quantitatively verified a faster interfacial reaction kinetics in 1.8 M LiFSI DOL electrolyte. The rate performances of graphite electrode in both electrolytes were further investigated via cyclic voltammetry (CV) using NG||Li cells (Supplementary Fig. 8). Basically, the graphite electrode in 1.8 M LiFSI DOL electrolyte experiences significantly deintercalation peaks with lower voltages but intercalation peaks with higher voltages than those in carbonate electrolyte from 0.01 to 1 mVs− 1, due to faster Li-ion kinetics. The b-values of each peak, from which we can infer the nature of the redox reaction (limited by semi-infinite diffusion or a capacitive process), were obtained via mathematic fitting43–45, denoted as C1–C3 (charge) and D1–D3 (discharge), respectively (Supplementary Fig. 8b,d). Based on the fitting, the b-values obtained from cells in 1.8 M LiFSI DOL electrolyte are higher than those in carbonate electrolyte. Due to the fast kinetics in 1.8 M LiFSI DOL electrolyte, Li+ can easily reach a ready state to diffuse within the graphite, which endows a strong driving force for the Li+ diffusion in the graphite indicated by the large b-values of each redox peak.
The cycling reversibility of 2.5NG in 1.8 M LiFSI DOL and commercial carbonate electrolytes was evaluated at 20C using NG||Li half cells (Fig. 2f). In stark contrast to only 40 mAh g− 1 for the cell with the carbonate electrolyte, the cell with 1.8 M LiFSI DOL electrolyte presents a highly reversible capacity of 315 mAh g− 1. At high current densities, the Li metal electrode have to be replaced every ~ 1000 cycles to eliminate the pernicious impact (high resistance, dendrites, and electrolyte consumption)46,47. Due to the surface corrosion of Li electrode, the cell capacity decreases to 180 mAh g− 1 after 4000 cycles without replacing the Li metal (Supplementary Fig. 9b,c), which can be recovered by replacing Li count electrodes. Different sized NGs in LiFSI-DOL at various concentrations were also evaluated (Supplementary Figs. 10 and 11). As can be seen, the cell comprised of 2.5NG and 1.8 M LiFSI DOL electrolyte displays the best rate performance. To reach a commercial level, NG electrodes with high loadings of 2 mAh cm− 2 and 3.5 mAh cm− 2 (Supplementary Fig. 12a) were tested with 1.8 M LiFSI DOL and showed a high reversible capacity of 340 mAh g− 1 and 325 mAh g− 1 at 4C respectively. Compared with the state-of-the-art value of 150 mAh g− 1 (Supplementary Fig. 12b), the high capacity of NG in our work brought a breakthrough for fast-charging battery chemistry.
SEI characterization. X-ray photoelectron spectroscopy (XPS) was performed to identify the chemical composition of the SEI on cycled graphite electrode with various time of Ar+ sputtering. For the SEI formed in carbonate electrolyte (Fig. 3a), both the organic (RCH2OLi, 290.7 eV) and inorganic (Li2CO3, 291.2 eV) species are detected in the C 1s spectra on the top surface. The content of carbon species is dominated (~ 80%) according to the XPS elemental analysis after 120 s sputtering with Ar+ (Fig. 3b), indicating most of the SEI components is derived from the EC decomposition. Meanwhile, the relatively weak signals for Li2CO3 and other inorganic species (O, C, P) (Supplementary Fig. 13a) during the sputtering process demonstrate the SEI formed in carbonate electrolyte is organic-rich.
Unlike SEI derived from carbonate electrolyte, the SEI of NG in 1.8 M LiFSI DOL electrolyte contains more inorganic (LiO, LiF) species (Fig. 3c,d and Supplementary Fig. 13b). LiF-rich feature can be confirmed by the intensive LiF signals in F 1s spectra (685.9 eV) and the high ratio of F content (~ 18% vs. 5% for SEI in carbonate electrolyte) in XPS elemental analysis. It is well known that LiF-rich SEI stabilizes the Li metal48 and alloy anodes49, but few has focused on LiF-rich SEI for graphite anode since the organic-rich SEI derived from the carbonate electrolyte sufficiently enables a decent cycling stability. Compared to the organic-rich SEI, the LiF-rich SEI is more promising for a fast kinetics since two or three atomic layers of LiF could block the side reactions thanks to the wide bond gap and high chemical/electrochemical stability50. Some other inorganic species contained N and S were also discovered in the SEI due to the decomposition of FSI− anions (Supplementary Fig. 14). The ab initio molecular dynamic (AIMD) simulated atomic SEI structures on the graphite was provided in Fig. 3e and Supplementary Fig. 15. For the carbonate electrolyte, the open-ringed EC, LixPF6, and Li2CO3 clusters can be found, validating that the SEI is mainly resulted from the decomposition of EC solvents. For 1.8 M LiFSI DOL electrolyte, LiF and LiNxSyOz clusters are formed at the interface without the capture of DOL decomposition, revealing that LiFSI is much easier to decompose at the interface, consistent with the XPS results. The formed LiF on graphite with low ionic electronic and high interface energy50 effectively blocks the continuous reaction of electrolytes, thus increasing the Coulombic efficiency.
Cryogenic transmission electron microscopy (cryo-TEM) technique, which can retain the morphology of in-situ SEI51,52, was carried out to analyze the specific structures of the SEI films. The morphologies and elemental components of the SEI are explicated by high-resolution cryo-TEM with the corresponding fast Fourier transform (FFT) pattern (Fig. 3f,j). For the carbonate electrolyte, the generated SEI presents a classic mosaic-type structure with nano-scale Li2CO3 (orange and magenta circles) and Li2O particles (red circles) (Fig. 3g) dispersed into amorphous organic components53. However, the main inorganic components of the generated SEI in the 1.8 M LiFSI DOL electrolyte are LiF (yellow circle) and Li2O (red circles) with a size of 5–10 nm (Fig. 3j, k), which is also confirmed by the results of other SEI regions (Supplementary Fig. 16). Unlike the SEI with thickness of > 40 nm formed in the carbonate electrolyte (Fig. 3g), the SEI derived in 1.8 M LiFSI DOL electrolyte is only 15 nm (Fig. 3k), which is favorable for the fast transport of Li+ crossing the SEI. On the other hand, the interphase generated in 1.8 M LiFSI DOL electrolyte exhibits a homogeneous laminar structure (Supplementary Fig. 17b) compared to the irregular morphology formed in the carbonate electrolyte (Supplementary Fig. 17a). Since the detective depth of energy dispersive X-ray spectroscopy (EDX) is micrometer-scale, the detected signal reflects the overlap of both SEI and graphite, which leads to a less apparent content for all elements except C. However, we can still infer that the interphase formed in 1.8 M LiFSI DOL electrolyte is more than ten times richer in inorganic species (2.61 wt.% F and 4.08 wt.% S) (Fig. 3l) than that formed in the carbonate electrolyte (0.15 wt.% F and 0.01 wt.% P) (Fig. 3h) from the relative contents. The interphase formed in 1.8 M LiFSI DOL electrolyte was also identified by ADF STEM and EELS, showing edge enriched distribution of the detected elements (Fig. 3m), which differs from the broad distribution of elements in the carbonate electrolyte (Fig. 3i).
LiFePO 4 ||graphite full cells capable of fast charging and low-temperature operation. Ultrafast-charging capability of 1.8 M LiFSI DOL for Li-ion batteries was evaluated using microsized natural graphite anodes and LiFePO4 cathodes with a high loading, although the high loading would challenge the rate performance54–56. We first evaluated the cycle stability of 2 mAh cm− 2 graphite electrodes at different charging C-rate. Notably, the 1.8 M LiFSI DOL electrolyte enables the graphite electrodes to achieve a capacity of 320 mAh g− 1 at 4C and 150 mAh g− 1 at 10 C (Fig. 4a) under such high loading condition, in sharp contrast to a capacity of 20 mAh g− 1 for the carbonate electrolyte at 4C (Fig. 4a). Meanwhile, a significantly improved cycling performance with a high-capacity retention of 250 mAh g− 1 over 400 cycles is also achieved for 1.8 M LiFSI DOL electrolyte (Supplementary Fig. 18). In this harsh condition, the Li metal electrode has to be replaced by a fresh one at the 180th cycle to ensure that the degradation of the Li metal under a current density of 8 mA cm− 2 does not affect the cycling performance (Supplementary Fig. 18c). When the rate increased to 8C (16 mA cm− 2), the degradation of the Li metal would be more severe, leading to a more noticeable capacity decay (Supplementary Fig. 19). Another tough condition with the lithiation rate of 4C and delithiation rate of 0.3C (Supplementary Fig. 20) was also employed to verify the practical feasibility and a high capacity of 320 mAh g− 1 was retained over 200 cycles in 1.8 M LiFSI DOL electrolyte.
LIBs deployed at ultra-low temperatures (≤ -30°C) have limited success due to the dramatic capacity drop, which largely limits their practical application in the extreme environments57,58. Although some progress has been reported, the utilized cycling protocol was to charge at room temperature and discharge at low temperatures. Worse kinetics of graphite lithiation at low temperatures remains one of most challenges for the LIBs. Therefore, the low-temperature performances of the cells with the designed electrolytes were also evaluated. Compared to the rapid drop for the ICs of 1.0 M LiPF6 EC/DMC (11.59 to 1.11 mS cm− 1 from 25 to -30 ℃), the IC of 1.8 M LiFSI DOL strictly follows the Vogel-Tamman-Fulcher (VTF) empirical Eq. 57 and decreases slowly (15.14 to 7.49 mS cm− 1 from 25 to -30 ℃) (Supplementary Fig. 21), showing a great advantage for Li+ transport in bulk electrolyte at low temperatures. Further low-temperature test (Fig. 4b,c), id est, cycling at C/3 under gradually decreasing temperatures from 25 to -30 ℃, proves the superiority of the electrolyte by showing a capacity of 300 mAh g− 1 at -30 ℃ compared with the carbonate electrolyte (no capacity at -30 ℃).
LiFePO4 (LFP) cathode was selected to construct full cell due to its high safety, low cost, and stability59,60. Preliminary evaluation of the LFP||Li cells shows the advantages of 1.8 M LiFSI DOL electrolyte by presenting a capacity of 90 mAh g− 1 at 20C, four times higher than that in the carbonate electrolyte (20 mAh g− 1) (Supplementary Figs. 22 and 23). Coin and pouch full LFP||NG cells tests were also conducted to examine the long cycling stability at 20C (Supplementary Fig. 24 and Fig. 4d). As can been seen, the cell with 1.8 M LiFSI DOL electrolyte delivers an initial capacity of 90 mAh g− 1 and has no capacity decay after 1000 cycles (Supplementary Fig. 24c). The possible Li plating at high charging rate is the key issue restricting the LIB rate performance61,62. Both the digital and SEM images demonstrate that no Li plating occurring at any state of charge (SOC) during the fast-charging process (Supplementary Fig. 25), which guarantees the stable long cycling performance of full cells. At 60C (1 min to full charge and discharge respectively), the pouch cell with 1.8 M LiFSI DOL electrolyte exhibits a capacity of 60 mAh g− 1 without decay during long-time cycling along with a stable Coulombic efficiency of 99.99% (Fig. 4d and Supplementary Fig. 26), which far exceeds the cell performance with the carbonate electrolyte.
Electrolyte design for fast-charging graphite pairing with NCM811. High-voltage NCM811||graphite cells have received much attention because of their high energy density63. Following the criterion above, electrolyte with 1.0 M LiPF6 in FEC/AN (7:3 by vol.) was designed for NCM811||graphite cells. Compared to the carbonate electrolyte, 1.0 M LiPF6 FEC/AN exhibits a lower ΔEdsv (31.43 kJ mol− 1) and superior ionic conductivities at room temperature (13.22 mS cm− 1) (Supplementary Fig. 27a) and low temperature (2.50 mS cm− 1 at -30 ℃) (Supplementary Fig. 27b). The typical Li+ solvation structures with and without anions were presented in Fig. 5a,b. The RDF results demonstrate an average first shell solvation structure of Li(PF6)(FEC)3AN. The FEC solvent and PF6 anions in the solvation shell tend to be reduced to form LiF at the interface while the AN solvent accelerates the Li-ion transport within the electrolyte, thus meeting the requirements of our proposed mechanisms. Detailed solvation structures of 1.0 M LiPF6 FEC/AN are shown in Supplementary Fig. 28.
The 1.0 M LiPF6 FEC/AN electrolyte enables graphite electrode to achieve a higher capacity of 230 mAh g− 1 at 20C, which is superior to that of carbonate electrolyte (40 mAh g− 1) (Fig. 5c and Supplementary Fig. 29a,b). Moreover, the NCM811||NG cell in 1.0 M LiPF6/FEC:AN with the cathode loading of 2 mAh cm− 2 maintains a highly reversible capacity of 170 mAh g− 1 under a harsh cycling condition (4C charge 0.3C discharge).