Lithium-ion batteries (LIBs) have been extensively employed in portable electronics and electric vehicles because of their high energy/power density and long cycle life1-3. Nevertheless, they inevitably suffer from severe energy/power losses in cold environments, especially when the temperature drops below −20°C. For example, a commercial 18650 LIB only delivers ~5% of energy density and ~1% of power density at −40°C compared to those achieved at room temperature (+20°C)4,5. Such poor low-temperature performance limits their applications for aeronautics/space missions, polar expeditions and many military and civil facilities in cold regions, in which a battery operating temperature below −40°C is required4,6.
Searching for a system with high electrochemical reactivity beyond the Li technologies would be promising to address the challenges associated with low temperature operation. As an alkali metal, sodium (Na) stands out given that it shares many chemical and physical properties with Li, while being more reactive and substantially more naturally abundant7-9. With a lower first ionization energy than Li (495.8 vs. 520.2 kJ mol-1)10, Na chemistry can contribute to improved chemical/electrochemical reactivity and thus may facilitate electrochemical reactions in cold environments. Just like Li metal in Li-based batteries, Na metal plays a crucial role in Na batteries because of its low electrode potential (−2.714 V vs. standard hydrogen electrode) and high theoretical specific capacity (1166 mAh g-1)7-9. Nevertheless, investigation of Na batteries at low temperatures has been limited, and in particular understanding of the behavior of Na metal as an electrode is largely missing11-14.
Enabling low-temperature battery operation is furthermore highly dependent on the nature of the electrolyte17-19. The electrolyte resistance increases rapidly as temperature drops because of the relatively high freezing/melting points of traditional carbonate solvents and the reduced solubility of conducting salts5,17. Moreover, the structural and compositional evolution of the solid electrolyte interphase (SEI) formed on Na metal is unclear at low temperatures. Differences of the SEI are likely due to the change of the highest occupied molecular orbital (HOMO)-lowest unoccupied molecular orbital (LUMO) energies of electrolytes as a function of temperature6. Hence, the SEI-friendly electrolytes at ambient temperature might not be able to maintain the same protective capabilities to enable efficient cycling under cold conditions.
One viable solution to solve these problems is the formulation of electrolytes targeted for low-temperature operation using ether solvents with low melting points and SEI-benign salts. Using this strategy, we find here that a unary-solvent electrolyte using an acyclic ether is able to extend the Na metal operating temperature towards −40°C. Furthermore, adding a cyclic ether as a reconcilable anti-freeze agent in a binary-solvent electrolyte can expand the temperature threshold of thermostability down to −150°C. To achieve this superior performance, the trifluoromethanesulfonate (OTf) salt is found to play a critical role in stabilizing the SEI at low temperatures, as confirmed by X-ray photoelectron spectroscopy (XPS) and cryogenic transmission electron microscopy (cryo-TEM). We demonstrate stable Na plating/stripping in symmetric cells at an extreme temperature of −80°C, exhibiting low overpotentials of ~150 mV for over 750 hours. The achieved performance outperforms the state-of-the-art results of Na and Li metal anodes cycling at low temperatures (≤−40ºC) shown in Supplementary Fig. 1 and Supplementary Table 115-18. We further illustrate a full cell with a capacity decay of less than 0.089% per cycle and a high average Coulombic efficiency (CE) over 99.5% for over 100 cycles at temperatures as low as −60°C.
Screening of unary-solvent electrolytes at low temperatures
While 1M sodium hexafluorophosphate in diethylene glycol dimethyl ether (NaPF6-DEGDME) has been shown to be a suitable electrolyte for the Na metal anode at ambient conditions20, it is impractical for low-temperature operation, mainly due to the precipitation of salt as temperature drops (Supplementary Fig. 2). Therefore, we screened the electrolytes using a variety of salt-solvent combinations. Low melting point ethers, including DEGDME, 1,2-dimethoxyethane (DME) and 1,3-dioxolane (DOL) were chosen as candidate solvents (Supplementary Table 2). A variety of conducting salts, including NaPF6, sodium trifluoromethanesulfonate (NaOTf), sodium perchlorate (NaClO4), sodium bis(trifluoromethanesulfonyl)imide (NaTFSI) and sodium bis(fluorosulfonyl)imide (NaFSI) were investigated as the candidate salts. The concentration of salt in the solvent was kept as 1M for the homogeneity screening of these unary-solvent electrolytes at low temperatures. The results (Supplementary Table 3) suggest that the solubility of NaPF6 cannot reach 1M concentration in all three solvents, while the other four salts present good solubility in DEGDME. Moreover, DOL cannot easily solvate Na salts except for NaTFSI. In sum, eight out of the fifteen unary-solvent electrolytes are thermally stable at −35˚C at 1M concentration, and these were subjected to further electrochemical evaluation with metallic Na.
Electrochemical behavior of unary-solvent electrolytes
Galvanostatic cycling measurements were performed in symmetric Na/Na cells at a current density of 0.5 mA cm−2 with a cycling capacity of 0.5 mAh cm−2 at −20˚C (Fig. 1) and at +20˚C (Supplementary Fig. 3). At +20˚C, 1M NaOTf-DEGDME presents the most stable cycling with the smallest average overpotential (< 10 mV) for 600 hours (300 cycles). In contrast, 1M NaTFSI-DOL shows increasing overpotential which rapidly reaches 1 V, and the 1M NaTFSI-DEGDME and NaTFSI-DME exhibit early failure, reaching the protection voltage (5 V) at 32 hours (16 cycles) and at 44 hours (22 cycles), respectively. The rest of the electrolytes exhibit asymmetric voltage profiles over cycling (Supplementary Fig. 4).
In contrast to the behavior at +20˚C, no systems exhibit early failure at −20˚C (Fig. 1a), likely owing to the reduced reactivity of the electrolyte with Na. Among the eight electrolytes, the 1M NaOTf-DEGDME displays the highest stability with smooth and steady voltage plateaus as well as the lowest overpotential of ~16 mV (Fig. 1b). In comparison, unstable and fluctuating voltage spikes are observed for the rest of the electrolyte recipes (Fig. 1b and 1c). The overpotentials and voltage spike magnitudes of different electrolyte systems are compared in Supplementary Table 4. Because of the superior performance of the NaOTf-DEGDME at both +20˚C and −20˚C, it was selected for further testing in symmetric cells at −40˚C (Fig. 2a). Remarkably, the initial overpotential of 40 mV rises to only 50 mV even after 500 hours of cycling at 0.5 mA cm-2 (capacity: 1 mAh cm-2), corresponding to less than 0.2% increase per cycle. Even at a higher current density of 1 mA cm-2, an average overpotential of 100 mV can be maintained for over 300 hours. As for CE evaluation (Fig. 2b), the fluctuating profile at −40˚C compared to +20 ˚C suggests that the formation of stable SEI is difficult at lower temperatures.
Characterization of unary-solvent electrolytes
To further understand the effects of temperature on cycling, the morphology of the SEI was characterized using scanning electron microscopy (SEM) (Supplementary Figs. 5 to 8). It is noted that both NaTFSI and NaFSI salts result in damaged surfaces (pores and/or cracks) of the deposited Na21-25. In contrast, the NaOTf-DEGDME enables relatively smooth surfaces at both +20°C and −20°C, where leaf-vein-like texture could be observed on the surface and within the cross-section. Such microstructure becomes more distinct at −40°C (Fig. 2c).
The chemical composition of the SEI formed in NaOTf-DEGDME was revealed using X-ray photoelectron spectroscopy (XPS), as shown in Fig. 2c. The binding energies of all elements were calibrated with respect to C1s at 284.8 eV. In the C1s XPS profile, the peak with binding energy at 293.4 eV is assigned to -CF3, and the ones at 288.7 eV and 286.6 eV are ascribed to O-C=O and C-O-C, respectively26-31. In the O1s spectrum, the peaks at 533.3 eV and 531.2 eV correspond to polyether-like compounds and C-O-Na (e.g. RCH2ONa)32, respectively, and the one at 536.3 eV is assigned to Na KLL. For S2p, the doublets at 169.6 eV and 167.2 eV (based on the 2p3/2) are assigned to SO42- and SO32-, respectively. As to F1s, the peaks at 689.3 eV and 684.1 eV are assigned to C-F and NaF, respectively26-31. These XPS spectra at different temperatures reveal similar composition on the SEI surface (Fig. 2c), suggesting the existence of -CF3 containing compounds, Na2SO4, Na2SO3, Na2CO3, NaF, C-O-Na and organic debris. It is noteworthy that Na2SO3 can only be detected on the surface at +20°C. Besides, Na2S (S2-: 161.1 eV) cannot be detected in the inner SEI at −20°C and −40°C (Supplementary Fig. 9 to 11). The presence of the species with lower oxidation states at lower temperatures suggests that the reaction kinetics between the electrolyte and Na is suppressed. Moreover, the XPS peak intensities at −40°C are significantly lower than those at +20°C and −20°C, indicating more limited SEI formation with decreasing temperature, which is in accord with the CE measurements. Importantly, the high consistency in SEI composition formed in NaOTf-DEGDME across these temperatures cannot be achieved in the other electrolyte systems (Supplementary Figs. 12 to 23). Note that the discussion of morphology and chemical composition of the SEIs in the other systems is detailed in the Supplementary Information.
Design of binary-solvent electrolytes for extreme low temperatures
Motivated by the analysis of the nature of the SEI (Supplementary Tables 5 and 6)26-34, the NaOTf-DEGDME combination was selected as a launching point for further formulation of electrolytes that can operate below −40°C. A different solvent, DOL, which is a cyclic ether with a low melting point of −95°C, was introduced into DEGDME to produce binary-solvent electrolytes (Fig. 3a) to further improve the low-temperature properties of the electrolyte. However, the Na+ solvation ability of DOL is lower than that of DEGDME35. Therefore, the solvent volume ratios (DEGDME:DOL=8:2, 5:5 or 2:8 volume fraction) and salt concentration (0.5M or 1M) were evaluated in the screening of binary-solvent electrolytes (Supplementary Table 7).
It can be observed that the introduction of DOL drastically extends the low temperature operating range of the NaOTf-DEGDME system, where a higher DOL volume leads to better Na cycling stability in symmetric cells (Fig. 3b and Supplementary Fig. 24). Additionally, a lower salt concentration (0.5M) could accommodate an increased DOL proportion, further enhancing the performance (Fig. 3b). At −80°C, the symmetric cell overpotential (~35 mV) in 0.5M NaOTf-DEGDME/DOL (2:8) is less than half of that (~75 mV) in 0.5M NaOTf-DEGDME/DOL (5:5) and six times less than that (>200 mV) in 1M NaOTf-DEGDME/DOL (5:5). We found that both the NaOTf salt and the DEGDME solvent are critical for enabling this superior low-temperature performance. Replacing DEGDME with DME leads to increased overpotentials and/or an asymmetric voltage profile (Supplementary Fig. 25), which also occurs with the substitution of NaOTf by NaClO4 (Supplementary Fig. 26).
Evaluation of binary-solvent electrolytes
To understand the mechanisms for the enhanced performance at low temperatures, the electrolyte resistance was investigated at different temperatures (Fig. 3c). Remarkably, the resistance of 0.5M NaOTf-DEGDME/DOL (2:8) at −80°C is only 3.4 times higher than that at +20°C. This difference is less than half of that of 0.5M NaOTf-DEGDME/DOL (5:5) and ten times smaller than that of 1M NaOTf-DEGDME. The lower electrolyte resistance change is possibly attributed to the lower dynamic viscosity of DEGDME/DOL (2:8) and good homogeneity/uniformity of this electrolyte (Supplementary Figs. 27 and 28). Differential scanning calorimetry (DSC) further confirms the high thermal stability of 0.5M NaOTf-DEGDME/DOL (2:8), revealing no obvious phase transition even down to −150°C (Fig. 3d)36. Similarly, no phase transition is detected until −140°C for 0.5M NaOTf-DEGDME/DOL (5:5). In comparison, 1M NaOTf-DEGDME displays a second-order phase transition at −126°C during cooling and a first-order phase transition at −64°C (melting point of DEGDME) during the subsequent heating.
Given the low resistance and good thermal behavior, long-term Na/Na symmetric cells cycling at −80°C were compared in Supplementary Figs. 29 and Fig. 3e. The cells using 0.5M NaOTf-DEGDME/DOL (2:8) display small overpotentials of ~50 mV without increasing for over 2,000 hours at 0.2 mA cm-2 with 0.1 mAh cm-2 cycled (Supplementary Figs. 29). Even at a higher current of 0.5 mA cm-2 with a higher capacity of 0.25 mAh cm-2, a stable operation over 750 hours can still be achieved (Fig. 3e). Enlarged voltage profile (Fig. 3f) further suggests that the overpotential at 0.5 mA cm-2 starts at 147 mV (after stabilization process) and slightly increases to 150 mV after 500 hours, revealing a smooth voltage profile without spikes. In comparison, the 0.5M NaOTf-DEGDME/DOL (5:5) electrolyte contributes to large spikes (> 4,000 mV) and asymmetric voltage profile over cycling.
Characterization of SEI in binary-solvent electrolyte
On top of the improvements in electrolyte properties, the binary-solvent system can alter the features of the SEI in comparison to the unary-solvent one (Fig. 4a). Specifically, the gradual increase of the volume fraction of DOL results in the vanishing of leaf-vein-like texture that was observed in NaOTf-DEGDME at low temperatures (Supplementary Fig. 30 and Fig. 4b). As to the composition, XPS profiles of the NaOTf-DEGDME/DOL electrolyte at −80°C (Fig. 4c and Supplementary Figs. 31 to 34) suggest similar SEI components detected in its unary-solvent counterpart. Cryo-TEM (Fig. 4d and Supplementary Figs. 35 and 36) further indicates the crystalline nature of the SEI (likely containing Na2SO4 and Na2CO3) via identification of lattice fringes and analysis of electron diffraction patterns.
The uniformity/homogeneity of SEI achieved at −80°C using the binary-solvent electrolyte supports the long-term stability observed in Na/Na symmetric cells (Fig. 3e). The high integrity of the SEI also contributes to a significantly improved CE behavior at −40°C compared to that at +20°C, showing an average value of ~97.6% (Supplementary Fig. 37).
Demonstration of full cells working at low temperatures
Full cells comprising Na3V2(PO4)3 as the cathode8 and Na metal as the anode in the 0.5M NaOTf-DEGDME/DOL (2:8) electrolyte were demonstrated at 0.2C (1C=110 mA g-1) at low temperatures down to −80°C (Fig. 5). Galvanostatic cycling (Fig. 5a) reveals that the discharge capacity decreases with the cell being sequentially cooled. At −60°C, ~42% of the specific capacity obtained at −20°C can be retained. The capacity loss from the temperature stepping is reversible, which is fully recovered as the temperature rises. The charge-discharge voltage profiles from 0°C to −80°C (Fig. 5b) suggest that the hysteresis is also temperature sensitive. At −60°C, the hysteresis is about four times than that of at −40°C and five times than that at −20°C. For long-term cycling (Fig. 5c), a full cell at −20°C exhibits an initial discharge capacity of 91.8 mAh g-1 (~83.5% of the theoretical value) with a low decay rate (< 0.055% per cycle) and a high average CE value (> 98.1%) for over 100 cycles. At −40°C, the initial capacity decreases to 67.5 mAh g-1 while a higher CE value (> 99.6%) can be achieved. Even at −60°C, a similarly high CE (> 99.5%) can still be maintained with a high capacity retention of 91.1% (a decay rate < 0.089% per cycle) after 100 cycles. At −40˚C and −60˚C, C-rate performance up to 1C (Fig. 5d) further confirms the superiority of Na metal batteries using the designed electrolyte at extremely cold conditions.