Significance
Aqueous zinc-ion batteries (ZIBs) possess the advantages of low cost and high safety. However, their Zn anodes are usually confronting with the challenge of side reactions and Zn dendrite formation. Herein, we developed low ionic association electrolytes by introducing 2, 2, 2-trifluoroethanol (TFE) into 30 m ZnCl2. TFE will construct H-bonds with H2O in Zn2+ solvation structures and results in low ionic association due to the electron-withdrawing effect of -CF3 groups in TFE molecules. Therefore, Zn anodes in such electrolyte display a fast Zn2+ plating/stripping kinetics and high Coulombic efficiency. The resultant full batteries exhibit enhanced rate performance and stable cycling behavior.
Keywords: low ionic association, electron-withdrawing effect, aqueous zinc-ion batteries
Abstract
Aqueous zinc-ion batteries are emerging as one of the most promising large-scale energy storage systems due to their low cost and high safety. However, Zn anodes often encounter the problems of Zn dendrite growth, hydrogen evolution reaction, and formation of by-products. Herein, we developed the low ionic association electrolytes (LIAEs) by introducing 2, 2, 2-trifluoroethanol (TFE) into 30 m ZnCl2 electrolyte. Owing to the electron-withdrawing effect of -CF3 groups in TFE molecules, in LIAEs, the Zn2+ solvation structures convert from larger aggregate clusters into smaller parts and TFE will construct H-bonds with H2O in Zn2+ solvation structure simultaneously. Consequently, ionic migration kinetics are significantly enhanced and the ionization of solvated H2O is effectively suppressed in LIAEs. As a result, Zn anodes in LIAE display a fast plating/stripping kinetics and high Coulombic efficiency of 99.74%. The corresponding full batteries exhibit an improved comprehensive performance such as high-rate capability and long cycling life.
Aqueous zinc-ion batteries (ZIBs) hold great promise for large-scale energy storage because of their high safety, low cost, and environmental benignity (1–7). As a common anode material in ZIBs, Zn metal possesses the advantages of abundant resources, high theoretical specific capacity (820 mAh g−1 and 5,854 mAh cm−3), and low Zn/Zn2+ redox potential (−0.76 V vs. standard hydrogen electrode) (8–11). However, Zn metal also suffers from the issues of side reactions and Zn dendrite formation in traditional aqueous electrolytes. The side reactions, including Zn anode corrosion and cathodic hydrogen evolution reaction (HER), will consume electrolytes irreversibly and induce the formation of insoluble by-products on the surface of Zn anode (12–14). Zn dendrites that originated from inhomogeneous Zn deposition often pierce the separators, and this thus results in the short circuit of batteries. Simultaneously, some Zn dendrites are usually detached from the Zn anode easily and convert into form “dead Zn”. Therefore, side reactions and Zn dendrites will lead to low Coulombic efficiency and severe capacity degeneration of aqueous ZIBs.
Recently, “water-in-salt” electrolytes were developed to overcome the issues of Zn anodes (15–19). The H2O molecules in “water-in-salt” electrolytes are mainly restrained in Zn2+ ion-solvated shells. As a result, the redox potentials of these H2O molecules are largely extended and their activities are limited, achieving the suppression of HER. In addition, “water-in-salt” electrolytes can provide abundant Zn2+ ions at the surface of Zn anodes, which will homogenize the Zn2+ ion flux and thus inhibit the growth of Zn dendrites. Compared with other zinc salts that were used in “water-in-salt” electrolytes, ZnCl2 displays much lower cost, as a result, the high-concentration ZnCl2 “water-in-salt” electrolyte is attracting great attention (20–22). In 30 m ZnCl2 electrolyte, the activity of H2O molecules would be effectively suppressed in their Zn2+ solvation structures, however, the ionization of solvated H2O could be still induced by the interaction between Zn2+ ions and them. As a result, the HER often takes place at the surface of Zn anodes easily along with the formation of by-products (23–26). Furthermore, bivalent Zn2+ ions in the solvation structure will coordinate with the Cl− ions in neighboring solvation structures to form aggregates by the strong electrostatic interaction (27, 28). The existence of these aggregates results in that 30 m ZnCl2 electrolyte usually displays high viscosity with low ionic transport rate, limiting the power density of corresponding ZIBs. Thus, the solvation structures of 30 m ZnCl2 electrolyte have to be modified to inhibit the ionization of solvated H2O and weaken the interaction between Zn2+ ions and Cl− ions.
In this work, we designed the low ionic association electrolytes (LIAEs) by introducing 2, 2, 2-trifluoroethanol (TFE) into 30 m ZnCl2. In such LIAEs, owing to the electron-withdrawing effect of trifluoromethyl (-CF3) groups in TFE molecules, TFE will construct H-bonds with H2O in the Zn2+ solvation structure instead of coordinating with Zn2+ ions, which suppresses the ionization of H2O in Zn2+ solvation structure and reduces the electrostatic interaction between Zn2+ ions and Cl− ions (Fig. 1). Simultaneously, ionic association will be weakened and the aggregate clusters are divided into smaller solvation structures, resulting in enhanced ionic migration kinetics. Therefore, the Zn anodes in LIAE exhibit faster plating/stripping kinetics and higher Coulombic efficiency. The corresponding full batteries demonstrate superior rate performance and stable cycling behavior.
Fig. 1.
Schematic diagrams of the Zn2+ solvation structures and interfacial side reactions in 30 m ZnCl2 and LIAE-20%.
Results
The LIAEs were prepared by adding TFE into 30 m ZnCl2 electrolyte under stirring until the solution became transparent and stable. The molar ratios of TFE and H2O are adjusted in a range of 10 to 50% (SI Appendix, Fig. S1). In the LIAEs, there are some aggregate clusters with polymer-like characteristic in electrolytes, which can be reflected by their glass transition temperature (Tg) in differential scanning calorimetry (DSC) (SI Appendix, Fig. S2). The DSC curves of LIAEs display an endothermic peak that represents the Tg of electrolytes. When more TFE was added into the LIAEs, the Tg values of LIAEs decreased obviously, indicating that aggregate clusters are divided into smaller parts. Dynamic light scattering (DLS) further displays the size distribution of aggregate clusters accurately, as shown in Fig. 2A and SI Appendix, Fig. S3. The average size of aggregates in LIAEs varies from 2.94 to 1.31 nm with the addition of TFE molecules from 10 to 50%. Simultaneously, the amount of aggregates with smaller sizes will be increased, as suggested by the Tyndall effect (Fig. 2B). When the amount of TFE is increased, the intensity of scattered light in LIAEs is gradually enhanced, indicating the existence of more aggregates with smaller sizes (29). Combining with low viscosity and shear stress (Fig. 2C), LIAE-20% displays an improved ionic conductivity of 17.7 mS cm−1 (Fig. 2D), which is higher than the case of 30 m ZnCl2 electrolyte. However, when more TFE is added in the range of 30 to 50%, the ionic conductivity of LIAEs will be gradually reduced due to the drop of ZnCl2 concentration.
Fig. 2.
The characteristics of LIAEs with different molar ratios. (A) Distribution of particle size. (B) Tyndall effect. (C) Viscosity and shear stress. (D) Ionic conductivity. (E) Raman spectra. (F) FTIR spectra. (G) LSV curves.
To further understand the solvated structures of aggregate clusters in LIAEs, the LIAEs with different TFE contents were characterized by Raman spectra. As shown in Fig. 2E, the peak at around 240 cm−1 is attributed to the ionic aggregates of [ZnxCl2x+2]2− (30), which are induced by electrostatic interaction between Zn2+ ions and Cl− ions in neighboring solvation structures. The peaks at around 293 and 337 cm−1 are assigned to the solvation structures of [Zn(H2O)2Cl4]2− and [Zn(H2O)Cl3]−, respectively. After introducing TFE into LIAEs, the peak intensity of [ZnxCl2x+2]2− is weakened, whereas [Zn(H2O)2Cl4]2− and [Zn(H2O)Cl3]− peaks are significantly enhanced. These suggest the solvation structure conversion from [ZnxCl2x+2]2− to [Zn(H2O)2Cl4]2− and [Zn(H2O)Cl3]− in LIAEs due to the addition of TFE, which is different from the case of dilute electrolytes (SI Appendix, Fig. S4). However, it is noted that the peaks of [Zn(H2O)2Cl4]2− and [Zn(H2O)Cl3]− shift to high wavenumber with the increase of TFE contents. It is ascribed to the H-bonds interaction between TFE and the H2O molecules in solvation structures, which were further identified by the FTIR spectra, as shown in Fig. 2F and SI Appendix, Fig. S5. In FTIR spectra, the characteristic peak at 1,277 cm−1 corresponds to the -CF3 group of TFE molecules (31). It is gradually enhanced and shifts to low wavenumber with the addition of TFE molecules. Besides, the peaks of O–H bending vibration (1,615 cm−1), O–H symmetric (3,200 cm−1), and asymmetric (3,380 cm−1) stretches in H2O are weakened and exhibit a blue shift. These indicate the formation of H-bonds between TFE and H2O molecules. Furthermore, with the increase of TFE contents in LIAEs, more H-bonds between TFE and H2O molecules will be achieved.
The electron-withdrawing effect of -CF3 group in TFE determines the Zn2+ solvation structures and the H-bonds between TFE and H2O molecules in LIAEs. In order to demonstrate the distribution of charges on the surfaces of solvent molecules, the molecular electrostatic potential (MESP) mappings of solvents were calculated, as shown in Fig. 3A. The O atom in H2O molecules exhibits a negative MESP (blue region) of −0.0672 Ha, which tends to coordinate with positive Zn2+ ions to induce Zn2+-H2O coordination structures. In contrast, the electron density of O atom in TFE molecule is reduced and corresponding MESP value is more positive since -CF3 group possesses a strong electron-withdrawing effect. As a result, TFE molecules are hard to coordinate with Zn2+ ions and will only interact with the H2O molecules in the solvation structures by H-bonds. It is different from the case of common aqueous solvent, where MESP value of O atom in hydroxyl is often more negative than the case of H2O molecules (SI Appendix, Fig. S6). They tend to coordinate with Zn2+ ions preferentially in 30 m ZnCl2 electrolyte (SI Appendix, Fig. S7), resulting in the separation of H2O molecules from solvation structures and the generation of free H2O in the electrolyte (SI Appendix, Fig. S8). The binding energies between Zn2+ ions and solvents can be compared by the density functional theory (DFT) calculation (Fig. 3B and SI Appendix, Figs. S9–S13). The binding energies of Zn2+-TFE with different coordination numbers are lower than the case of Zn2+-H2O. It further confirms that TFE is difficult to coordinate with Zn2+ ion in LIAE-20%, ensuring that the H2O molecules will be not separated from solvation structures and Zn2+ ion coordination in LIAE-20% is similar to the case in 30 m ZnCl2 even the aggregate clusters are divided into smaller parts in LIAE-20%.
Fig. 3.
(A) MESP mappings of H2O and TFE molecules. (B) Binding energies of Zn2+-H2O and Zn2+-TFE with different coordination numbers. MD simulation snapshots of (C) 30 m ZnCl2 and (D) LIAE-20%. (E) RDFs and corresponding average coordination number of 30 m ZnCl2 and LIAE-20%. (F) Ionization energies of H2O molecules in different Zn2+ solvation structures.
The Zn2+ solvation structures and the distributions of their neighboring molecules can be further understood by molecular dynamics (MD) simulations (Fig. 3 C and D). The MD simulations suggest that there are some aggregates of [ZnxCl2x+2]2− ions in 30 m ZnCl2 electrolyte due to the strong electrostatic interaction between Zn2+ ions and Cl− ions. After adding TFE, the Zn2+ solvation structure [ZnxCl2x+2]2− converts to the [Zn(H2O)2Cl4]2− and [Zn(H2O)Cl3]− in LIAEs-20%. Furthermore, TFE molecules will not coordinate with Zn2+ ions and only interact with the H2O molecules in these solvation structures by H-bonds. The corresponding radial distribution functions (RDFs) and coordination number distribution functions were further achieved (Fig. 3E). The LIAE-20% displays a broad Zn-OTFE peak at 0.53 nm, indicating that TFE molecules construct H-bonds with H2O instead of coordinating with Zn2+ ions, which inhibits the ionization of H2O (32). Different from LIAE-20%, 30 m ZnCl2 electrolyte shows the sharp Zn-OH2O peak at 0.21 nm, suggesting the coordination structure of Zn2+-H2O. These results are consistent with the discussion of the above MESP calculation. Moreover, the ionization energies of H2O in different electrolytes were also calculated to understand the activity of H2O in them (Fig. 3F). It is noted that the ionization energies of [Zn(H2O)2Cl4]2−-TFE (330.46 Kcal mol−1) and [Zn(H2O)Cl3]−-TFE (507.17 Kcal mol−1) are much higher than the cases of [Zn(H2O)2Cl4]2− (241.89 Kcal mol−1) and [Zn(H2O)Cl3]− (426.43 Kcal mol−1), respectively. Therefore, the H-bonds between TFE and H2O molecules will suppress the ionization of H2O in the Zn2+ solvation structure in LIAE-20%. As shown in linear sweep voltammetry (LSV) curves (Fig. 2G), H2 evolution potential in LIAE-20% is delayed and its decomposition current is smaller than the case of 30 m ZnCl2 electrolyte. Moreover, the chemical corrosion of Zn anodes is also inhibited in LIAE-20% (SI Appendix, Fig. S14). As a result, LIAE-20% was selected for further electrochemical analysis.
The symmetric cells with Zn foils (denoted as Zn||Zn-LIAE-20% and Zn||Zn-30 m ZnCl2) were utilized to illustrate the plating/stripping behavior of Zn electrodes in electrolytes. Their Zn2+ ion plating curves show that the nucleation overpotential of Zn2+ ions in LIAE-20% is 6.1 mV (Fig. 4A), which is lower than the case (10.1 mV) of 30 m ZnCl2 electrolyte. Moreover, the CV curve of Zn||Ti cells also displays a small nucleation overpotential in LIAE-20% (SI Appendix, Fig. S15). It is attributed to the enhanced wettability of LIAE-20% for Zn metal and the lower charge-transfer resistance at the interface between Zn metal and electrolyte (SI Appendix, Figs. S16 and S17). Such low nucleation overpotential will be beneficial to reduce the energy barrier of Zn2+ ion plating (33), which can be further reflected by the active energy (Ea) of desolvation during Zn2+ ions plating at Zn–electrolyte interface (Fig. 4C). The Ea is evaluated by electrochemical impedance spectroscopy (EIS) and can be calculated based on the Arrhenius equation (34):
| [1] |
Fig. 4.
Electrochemical performance of Zn anodes. (A) Voltage profiles during Zn2+ ion plating at 0.2 mA cm−2. (B) Rate performance of Zn||Zn-LIAE-20% and Zn||Zn-30 m ZnCl2. (C) Fitted Arrhenius curves and corresponding activation energies of Zn2+ desolvation in LIAE-20% and 30 m ZnCl2. (D) Tafel polarization curves in LIAE-20% and 30 m ZnCl2. (E) Coulombic efficiency of Zn||Zn-LIAE-20% and Zn||Zn-30 m ZnCl2. (F) X-ray diffraction (XRD) patterns after cycling in LIAE-20% and 30 m ZnCl2 for 200 cycles. (G) Cycling performance of Zn||Zn-LIAE-20% and Zn||Zn-30 m ZnCl2.
where Rct, T, R, and A represent charge-transfer resistance, absolute temperature, standard gas constant, and preexponential constant, respectively. The Ea of Zn2+ desolvation in LIAE-20% is 41.5 kJ mol−1, which is smaller than that (63.7 kJ mol−1) in 30 m ZnCl2 electrolyte. The H-bonds between TFE and H2O molecules in Zn2+ solvation structures result in the reduced Ea and simultaneously lead to the weakened coordination interaction between Zn2+ ions and H2O molecules, which benefits the desolvation of solvated Zn2+ ions at the Zn–electrolyte interface and accelerates the Zn2+ ion-transfer kinetics.
The fast Zn2+ ion-transfer kinetics at electrode–electrolyte interface endow the Zn||Zn-LIAE-20% with excellent rate capability, as shown in Fig. 4B. The voltage hysteresis of Zn||Zn-LIAE-20% and Zn||Zn-30 m ZnCl2 are comparable under low current density (0.2 mA cm−2) (SI Appendix, Fig. S18). However, at large current densities, the Zn||Zn-LIAE-20% exhibits lower voltage hysteresis in comparison with Zn||Zn-30 m ZnCl2, and even Zn||Zn-30 m ZnCl2 will suddenly break down at 3 mA cm−2, demonstrating the excellent electrochemical stability of Zn||Zn-LIAE-20% under high current density. In addition, the corrosion of Zn electrodes is also suppressed in LIAE-20%, as reflected by Tafel polarization curves (Fig. 4D). Generally, the corrosion current determines the corrosion kinetics of electrodes. The Zn anode in LIAE-20% displays a lower corrosion current density (0.13 mA cm−2) in comparison with the case of 30 m ZnCl2 electrolyte, indicating the diminished corrosion rate of the Zn anode in LIAE-20%. Therefore, the reversibility of Zn2+ plating/stripping process is enhanced in LIAE-20%. The average CE of Zn2+ plating/stripping in LIAE-20% can reach 99.74% after 100 cycles at 1 mAh cm−2 (Fig. 4E). Furthermore, at larger current density, average CE of Zn2+ plating/stripping is still high (SI Appendix, Fig. S19). In addition, there is nearly no dendrite on the Zn surface after cycling in LIAE-20%, which is different from the case of 30 m ZnCl2 electrolyte (Fig. 4F and SI Appendix, Figs. S20 and S21). As a result, the Zn||Zn-LIAE-20% displays a long cycle life with stable voltage hysteresis over 4,000 h (Fig. 4G). In contrast, the surface of Zn anode in Zn||Zn-30 m ZnCl2 is tarnished after cycling, and many irregular flake-like ZnCl4·4Zn(OH)2·H2O by-products are piled up on the Zn surface, resulting in the uneven Zn2+ ions plating and eventually causing the internal short circuit in cell.
To illustrate the practical feasibility of LIAE-20%, aqueous ZIBs were assembled based on polyaniline (PANI) cathodes (SI Appendix, Fig. S22), Zn anodes, and LIAE-20%. The resultant aqueous ZIBs can deliver a discharge capacity of 128 mAh g−1 at 0.1 A g−1 (Fig. 5A), which is comparable to those of previously reported PANI-based ZIBs (35, 36). Impressively, even at 5 A g−1, the discharge capacity of aqueous ZIBs can be still up to 49 mAh g−1 (Fig. 5B). Moreover, when the current density goes back to 0.1 A g−1, its capacity recovers to 127 mAh g−1, demonstrating an excellent rate performance. The reaction kinetics of ZIBs can be further reflected by their cyclic voltammetry (CV) curves with different scan rates (Fig. 5C). The CV curves show two pairs of redox peaks and the relationship between the peak currents (i) and scan rates (v) can be described as the following equation (37–39):
| [2] |
Fig. 5.
Electrochemical performance of Zn||PANI batteries. (A) Galvanostatic charge/discharge curves. (B) Rate performance. (C) CV curves at different scan rates. (D) The log (peak current) vs. log (scan rate) plots at each redox peak according to the CV curves. (E) Capacitive contributions at various scan rates. (F) The ion diffusion coefficient during discharging process. (G) Nyquist plots. (H) Cycling performance at 2 A g−1.
where a and b are the variable parameters. When b value approaches 1, the pseudocapacitance will determine the electrochemical reaction. It is noted that the b values of peaks 1, 2, 3, and 4 in CV curves are 0.88, 0.98, 0.95, and 0.99, respectively (Fig. 5D), suggesting that the redox reaction of the PANI cathode is mainly dominated by pseudocapacitance-controlled behavior (40). Moreover, with the increase of scan rates, the capacitive contribution of the PANI cathode in ZIBs rises gradually and can reach to 77% at 1 mV s−1 (Fig. 5E), which is higher than the case of ZIBs based on 30 m ZnCl2 electrolyte (SI Appendix, Fig. S23). In addition, the fast kinetics can be further understood by the diffusion coefficient (D) of ions in PANI cathodes, which was calculated by the galvanostatic intermittent titration technique measurement (41–43). As shown in Fig. 5F, the D values of PANI cathodes in ZIBs are 10−11 to 10−9 cm2 s−1, demonstrating a fast reaction kinetics of PANI cathodes in LIAE-20%. This is closely related to the small charge transfer resistance at the electrode/electrolyte interface (Fig. 5G). Furthermore, the ZIBs can still maintain 87.6% of its initial capacity even after charge/discharge for 2,000 cycles at 2 A g−1 (Fig. 5H), exhibiting a superior cycling behavior. The enhanced cycling performance is ascribed to that the introduction of TFE into 30 m ZnCl2 electrolyte will effectively suppress the dendrite and HER of Zn anodes in batteries (SI Appendix, Fig. S24). Therefore, the LIAE-20% endows the ZIBs with enhanced electrochemical performance.
Discussion
An LIAE was designed by introducing TFE into 30 m ZnCl2 electrolyte. The introduction of TFE induces the solvation structure conversion from [ZnxCl2x+2]2− to [Zn(H2O)2Cl4]2− and [Zn(H2O)Cl3]− in LIAE and the formation of H-bonds between TFE and H2O molecules in the solvation structures due to the electron-withdrawing effect of -CF3 groups in TFE molecules. As a result, LIAE displays higher ionic transport rate and the lower charge transfer resistance at the interface between Zn anodes and LIAE, leading to a faster plating/stripping kinetics. Simultaneously, the H-bonds between TFE and H2O molecules will suppress the ionization of H2O in Zn2+ solvation structure in LIAE. Therefore, in LIAE, the reversibility of Zn2+ plating/stripping behaviors is improved and the formation of by-products is inhibited, which endows Zn anodes with high Coulombic efficiency of 99.74%. The aqueous ZIBs with LIAE exhibit an excellent rate performance and a high capacity retention of 87.5% after 2,000 cycles. This work will pave a route to design high-rate and stable aqueous ZIBs.
Materials and Methods
Materials.
ZnCl2 (99.95%) and Zn foil (99.9%) were purchased from Alfa Aesar. PANI nanorods were prepared according to the previous literature (44). TFE (99.8%) and N-methyl-2-pyrrolidone (99.5%) were purchased from Aladdin. Single-walled carbon nanotubes (P3) were purchased from Carbon Solutions Inc. Polyvinylidene fluoride was purchased from SinopharmChemical Reagent Co., Ltd.
Preparation of LIAEs and PANI Cathode.
ZnCl2 (4.089 g) and TFE were dispersed in deionized H2O (1 g) under stirring to obtain LIAEs at different molar ratios. PANI, single-walled carbon nanotubes, and polyvinylidene fluoride were mixed with a weight ratio of 7: 2: 1 in N-methyl-2-pyrrolidone to achieve a slurry, which was then coated on a Ti foil. After vacuum drying at 60 °C for 24 h, the PANI cathode was obtained.
Characterizations and Electrochemical Measurements.
The size distribution of LIAEs was characterized via DLS (Malvern, ZetasizerNanoZS90). The viscosity and shear stress of LIAEs were measured by a rheometer (MCR302). The thermal properties of LIAEs were characterized using a differential scanning calorimeter (NETZSCH). The FTIR and Raman spectra of LIAEs were collected through Bruker Tensor II and a confocal Raman microscope (InVia Reflex) with a wavelength of 532 nm, respectively. The morphologies of materials were characterized via a scanning electron microscope (Phenom XL). CV curves of ZIBs were tested on an electrochemical workstation (CHI Instruments, CHI660E). Galvanostatic charge/discharge curves were recorded on a battery test system (LAND CT2001A). EIS curves were measured in the frequency range from 100 kHz to 100 mHz on an electrochemical workstation (Zahner, IM6ex). The sizes of PANI cathodes and Zn anodes are 1 × 1 cm2. The loading and thickness of SWCNTs/PANI cathodes are 1 mg cm−2 and 50 μm, respectively. The amount of electrolyte in a ZIB is 30 μL. The specific capacity of PANI cathodes was calculated based on the mass of the PANI.
Computational Methods.
The MESP mappings, binding energy and ionization energy calculations were performed using the DFT program DMol3 in Materials Studio. The physical wave functions were expanded in terms of numerical basis sets, DMol3/generalized gradient approximation (GGA)-Perdew–Burke–Ernzerhof (PBE)/Dual Numerical Polarization (DNP) (3.5) basis set. The core electrons were treated with DFT semicore pseudopotentials. The exchange-correlation energy was calculated with PBE GGA. A Fermi smearing of 0.005 Ha and a global orbital cutoff of 5.2 Å were employed. The convergence criteria for the geometric optimization and energy calculation were set as follows: a) a self-consistent field tolerance of 1.0 × 10−6 Ha atom−1; b) an energy tolerance of 1.0 × 10−5 Ha atom−1; c) a maximum force tolerance of 0.002 Ha Å−1; and d) a maximum displacement tolerance of 0.005 Å. The MD simulations were made by gromacs 2019.1 using the OPLS-AA force field. Nose–Hoover and Berendsen method were used as the temperature and pressure coupling method, respectively. The electrostatic interactions used PME methods. A cutoff length of 1.0 nm was used in the calculation of electrostatic interactions. To make the density of the system reach a stable value, the NPT process was used to equilibrate the system under the conditions of 298.15 K and 1 atm for 30 ns. Then, 10-ns simulation in the NVT process at 298.15 K was used for data collection.
Supplementary Material
Appendix 01 (PDF)
Acknowledgments
This work was supported by Ministry of Science and Technology of China (2019YFA0705600), National Natural Science Foundation of China (21875121), and “Frontiers Science Center for New Organic Matter”, Nankai University (63181206). We thank Professor Z. Zhou (Nankai University) for supporting Materials Studio calculations.
Author contributions
R.W., J.C., and Z.N. designed research; R.W. and Z.N. performed research; R.W. contributed new reagents/analytic tools; R.W., M. Yao, M. Yang, J.Z., J.C., and Z.N. analyzed data; and R.W. and Z.N. wrote the paper.
Competing interests
The authors declare no competing interest.
Footnotes
This article is a PNAS Direct Submission.
Data, Materials, and Software Availability
All study data are included in the article and/or SI Appendix.
Supporting Information
References
- 1.Pan H., et al. , Reversible aqueous zinc/manganese oxide energy storage from conversion reactions. Nat. Energy 1, 16039 (2016). [Google Scholar]
- 2.Cai S., et al. , Water-salt oligomers enable supersoluble electrolytes for high-performance aqueous batteries. Adv. Mater. 33, 2007470 (2021). [DOI] [PubMed] [Google Scholar]
- 3.Zhong C., et al. , A review of electrolyte materials and compositions for electrochemical supercapacitors. Chem. Soc. Rev. 44, 7484–7539 (2015). [DOI] [PubMed] [Google Scholar]
- 4.Xiao Y., et al. , Understanding interface stability in solid-state batteries. Nat. Rev. Mater. 5, 105–126 (2019). [Google Scholar]
- 5.Cheng X.-B., et al. , A review of solid electrolyte interphases on lithium metal anode. Adv. Sci. 3, 1500213 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Huang S., Zhu J., Tian J., Niu Z., Recent progress in the electrolytes of aqueous zinc-ion batteries. Chem. Eur. J. 25, 14480–14494 (2019). [DOI] [PubMed] [Google Scholar]
- 7.Huang J., Zhou J., Liang S., Guest pre-intercalation strategy to boost the electrochemical performance of aqueous zinc-ion battery cathodes. Acta Phys. -Chim. Sin. 37, 2005020 (2021). [Google Scholar]
- 8.Jin S., et al. , Production of fast-charge Zn-based aqueous batteries via interfacial adsorption of ion-oligomer complexes. Nat. Commun. 13, 2283 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Li H., et al. , Ultrastretchable and superior healable supercapacitors based on a double cross-linked hydrogel electrolyte. Nat. Commun. 10, 536 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Geng L., et al. , Eutectic electrolyte with unique solvation structure for high-performance zinc-ion batteries. Angew. Chem. Int. Ed. 61, e202206717 (2022). [DOI] [PubMed] [Google Scholar]
- 11.Hao J., et al. , Boosting Zn electrode reversibility in aqueous electrolyte using low-cost antisolvents. Angew. Chem. Int. Ed. 60, 7366–7375 (2021). [DOI] [PubMed] [Google Scholar]
- 12.Wei J., et al. , Dissolution-crystallization transition within a polymer hydrogel for a processable ultratough electrolyte. Adv. Mater. 31, 1900248 (2019). [DOI] [PubMed] [Google Scholar]
- 13.Qiu H., et al. , Zinc anode-compatible in-situ solid electrolyte interphase via cation solvation modulation. Nat. Commun. 10, 5374 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Li Q., et al. , Tailoring the metal electrode morphology via electrochemical protocol optimization for long-lasting aqueous zinc batteries. Nat. Commun. 13, 3699 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Wang F., et al. , Highly reversible zinc metal anode for aqueous batteries. Nat. Mater. 17, 543–549 (2018). [DOI] [PubMed] [Google Scholar]
- 16.Zhang C., et al. , A ZnCl2 water-in-salt electrolyte for a reversible Zn metal anode. Chem. Commun. 54, 14097–14099 (2018). [DOI] [PubMed] [Google Scholar]
- 17.Xiao D., Zhang L., Li Z., Dou H., Zhang X., Design strategies and research progress for water-in-salt electrolytes. Energy Storage Mater. 44, 10–28 (2022). [Google Scholar]
- 18.Guo Q., et al. , A high-potential anion-insertion carbon cathode for aqueous zinc dual-ion battery. Adv. Funct. Mater. 30, 2002825 (2020). [Google Scholar]
- 19.Zhang L., et al. , ZnCl2 “water-in-salt” electrolyte transforms the performance of vanadium oxide as a Zn battery cathode. Adv. Funct. Mater. 29, 1902653 (2019). [Google Scholar]
- 20.Wu X., et al. , Reverse dual-ion battery via a ZnCl2 water-in-salt electrolyte. J. Am. Chem. Soc. 141, 6338–6344 (2019). [DOI] [PubMed] [Google Scholar]
- 21.Yang M., Zhu J., Bi S., Wang R., Niu Z., A binary hydrate-melt electrolyte with acetate-oriented cross-linking solvation shells for stable zinc anodes. Adv. Mater. 34, 2201744 (2022). [DOI] [PubMed] [Google Scholar]
- 22.Tang X., et al. , Unveiling the reversibility and stability origin of the aqueous V2O5-Zn batteries with a ZnCl2 “water-in-salt” electrolyte. Adv. Sci. 8, 2102053 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Ji X., A perspective of ZnCl2 electrolytes: The physical and electrochemical properties. eScience 1, 99–107 (2021). [Google Scholar]
- 24.Cheng H., et al. , Emerging era of electrolyte solvation structure and interfacial model in batteries. ACS Energy Lett. 7, 490–513 (2022). [Google Scholar]
- 25.Qin R., et al. , Progress in interface structure and modification of zinc anode for aqueous batteries. Nano Energy 98, 107333 (2022). [Google Scholar]
- 26.Liu C., Xie X., Lu B., Zhou J., Liang S., Electrolyte strategies toward better zinc-ion batteries. ACS Energy Lett. 6, 1015–1033 (2021). [Google Scholar]
- 27.Zhang Q., et al. , Modulating electrolyte structure for ultralow temperature aqueous zinc batteries. Nat. Commun. 11, 4463 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Wei J., et al. , Metal-ion oligomerization inside electrified carbon micropores and its effect on capacitive charge storage. Adv. Mater. 34, 2107439 (2021). [DOI] [PubMed] [Google Scholar]
- 29.Wang M., et al. , A colloidal aqueous electrolyte modulated by oleic acid for durable zinc metal anode. Chem. Eng. J. 451, 138589 (2023). [Google Scholar]
- 30.Zhang C., et al. , The electrolyte comprising more robust water and superhalides transforms Zn-metal anode reversibly and dendrite-free. Carbon Energy 3, 339–348 (2021). [Google Scholar]
- 31.Kadyan A., Gandhi Y., Pandey S., Probing interactions within liquid media via a model H-bond donor–acceptor mixture. Phys. Chem. Chem. Phys. 21, 4791–4801 (2019). [DOI] [PubMed] [Google Scholar]
- 32.Yang W., et al. , Hydrated eutectic electrolytes with ligand-oriented solvation shells for long-cycling zinc-organic batteries. Joule 4, 1557–1574 (2020). [Google Scholar]
- 33.Zhao K., et al. , Boosting the kinetics and stability of Zn anodes in aqueous electrolytes with supramolecular cyclodextrin additives. J. Am. Chem. Soc. 144, 11129–11137 (2022). [DOI] [PubMed] [Google Scholar]
- 34.Zhou X., et al. , Anion reinforced solvation for gradient inorganic-rich interphase enables high-rate and stable sodium batteries. Angew. Chem. Int. Ed. 61, e202205045 (2022). [DOI] [PubMed] [Google Scholar]
- 35.Wan F., et al. , An aqueous rechargeable zinc-organic battery with hybrid mechanism. Adv. Funct. Mater. 28, 1804975 (2018). [Google Scholar]
- 36.Zhang Q., et al. , Chaotropic anion and fast-kinetics cathode enabling low-temperature aqueous Zn batteries. ACS Energy Lett. 6, 2704–2712 (2021). [Google Scholar]
- 37.Tie Z., et al. , A symmetric all-organic proton battery in mild electrolyte. Angew. Chem. Int. Ed. 61, e202115180 (2022). [DOI] [PubMed] [Google Scholar]
- 38.Deng S., et al. , Rational design of ZnMn2O4 quantum dots in a carbon framework for durable aqueous zinc-ion batteries. Angew. Chem. Int. Ed. 61, e202115877 (2022). [DOI] [PubMed] [Google Scholar]
- 39.Wang D., et al. , A superior δ-MnO2 cathode and a self-healing Zn-δ-MnO2 battery. ACS Nano 13, 10643–10652 (2019). [DOI] [PubMed] [Google Scholar]
- 40.Hu Z., et al. , A self-regulated electrostatic shielding layer toward dendrite-free Zn batteries. Adv. Mater. 34, 2203104 (2022). [DOI] [PubMed] [Google Scholar]
- 41.Zhang T., et al. , Fundamentals and perspectives in developing zinc-ion battery electrolytes: A comprehensive review. Energy Environ. Sci. 13, 4625–4665 (2020). [Google Scholar]
- 42.Chao D., et al. , An electrolytic Zn-MnO2 battery for high-voltage and scalable energy storage. Angew. Chem. Int. Ed. 58, 7823–7828 (2019). [DOI] [PubMed] [Google Scholar]
- 43.Wang M., et al. , Toward dendrite-free and anti-corrosion Zn anodes by regulating a bismuth-based energizer. eScience 2, 509–517 (2022). [Google Scholar]
- 44.Wang R., Yao M., Huang S., Tian J., Niu Z., Sustainable dough-based gel electrolytes for aqueous energy storage devices. Adv. Funct. Mater. 31, 2009209 (2021). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Appendix 01 (PDF)
Data Availability Statement
All study data are included in the article and/or SI Appendix.





