Next Article in Journal
Battery-Type Lithium-Ion Hybrid Capacitors: Current Status and Future Perspectives
Next Article in Special Issue
Impact of Surface Structure on SEI for Carbon Materials in Alkali Ion Batteries: A Review
Previous Article in Journal
Genetic Algorithm and Taguchi Method: An Approach for Better Li-Ion Cell Model Parameter Identification
Previous Article in Special Issue
Recent Development of Electrolyte Engineering for Sodium Metal Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mini-Review on the Regulation of Electrolyte Solvation Structure for Aqueous Zinc Ion Batteries

1
College of Materials Science and Engineering, Changsha University of Science and Technology, Changsha 410114, China
2
School of Chemistry, South China Normal University, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Batteries 2023, 9(2), 73; https://doi.org/10.3390/batteries9020073
Submission received: 23 November 2022 / Revised: 12 January 2023 / Accepted: 18 January 2023 / Published: 21 January 2023
(This article belongs to the Special Issue Review of Electrode Materials and Electrolyte for Batteries)

Abstract

:
Zinc as an anode, with low potential (−0.762 V vs. SHE) and high theoretical capacity (820 mAh g−1 or 5854 mAh L−1), shows great promise for energy storage devices. The aqueous zinc ion battery (ZIB) is known as a prospective candidate for large-scale application in the future due to its high safety, environmental friendliness, abundant zinc resources on earth, and low-cost advantages. However, the existence of zinc dendrites and side reactions limit the practical application of ZIBs. Therefore, a lot of effort has been made to improve the performance from aspects including the structure design and surface modification of zinc anodes, regulation of the electrolyte solvation structure, and design of the functional separator. In this review, we attempt to summarize recent advances on the regulation of the electrolyte solvation structure through a number of selected representative works from two aspects: high-concentration salt strategy and electrolyte additives. At the end of this review, the challenges and future development prospects are briefly outlined.

1. Introduction

The depletion of fossil energy and deterioration of the environment have led to the rapid development of sustainable clean energy, such as wind and solar energy. Thus, their proportion in the energy consumption structure has increased yearly. Due to the discontinuous and regional characteristics of clean energy, the development of energy storage devices has been driven [1,2,3]. The aqueous zinc ion battery (ZIB) with metal zinc as the anode exhibiting a high theoretical capacity and low reduction potential has attracted much attention. Meanwhile, ZIBs show the advantages of low cost with rich and cheap zinc resources, high energy density, and high safety, which has been regarded as one of the most favorable candidate systems for broad application prospects [4,5,6].
As early as 1986, Yamamoto et al. [7] made a preliminary exploration of the neutral secondary zinc manganese battery. In 2012, Kang et al. [8] proposed an environment-friendly and safe power battery structure consisting of MnO2 as the cathode with zinc metal as the anode, and ZnSO4 as the neutral electrolyte solution, which is called a “zinc ion battery” with the working principle of reversible intercalation/stripping of Zn2+ in MnO2. These works opened the prelude to the research of ZIBs. Since then, researchers have invested a lot of effort in studying the energy storage mechanism of ZIBs and have proposed many feasible strategies, for example, Zn2+ insertion/extraction, H+/Zn2+ insertion/extraction, and chemical conversion reaction.

1.1. Zn2+ Insertion/Extraction Mechanism

The cathode materials of ZIBs are generally selected from nanomaterials with a layered or tunneled structure. The Zn ions are inserted/extracted in the cathode material during the working process of ZIBs. Kang et al. [8] found that the energy storage in Zn2+||α−MnO2 cells can be ascribed to the reversible inserted/extracted reaction of Zn2+ in/from the lattice tunnel. The Zn anode can be successfully oxidized to Zn2+ during the discharge procedures and rapidly inserted in the cathode material with a laminar or tunnel structure after being transported through the electrolyte. Some of the Mn4+ is quickly reduced to Mn3+, but Mn3+ is prone to disproportionation reactions and converts to Mn4+ and Mn2+ due to its instability nature. Zn2+ would be removed from the cathode material during the charge process, pass through the electrolyte solution, and return to the anode for reduction to Zn. The Mn2+ in the electrolyte returns to the manganese vacancy and is oxidized to Mn3+ and Mn4+, and the original tunneling structure is restored [9]. Furthermore, Mai et al. [10] also confirmed the highly reversible single-phase reaction of Zn2+ insertion in vanadium-based materials using the operando technique and the corresponding qualitative analysis. The reaction equation of manganese-based materials can be described as follows:
Cathode   process :   Zn 2 + + 2 e + 2 MnO 2 ZnMn 2 O 4
Anode   process :   Zn Zn 2 + + 2 e

1.2. H+ and Zn2+ Co-Insertion/Extraction Mechanism

The energy storage mechanism for another Zn2+||α-MnO2 cell was proved to be the H+/Zn2+ co-insertion/extraction process. Wang et al. [11] characterized the diffusion kinetics of Zn2+ and H+ by the constant current intermittent titration technique (GITT). The H+/Zn2+ co-insertion mechanism was confirmed in a Mn-based cathode. In particular, H+ will first embed into the α-MnO2 material in the discharge stage, and Zn2+ with relatively large radius will be vulnerable to electrostatic effects when entering the lattice gap. The embedding process can be slightly affected and will slowly embed into α-MnO2 after H+ embedding, finally achieving reversible ion transport. During the discharge process, MnOOH and OH will generate from the reaction between MnO2 and H2O, while Zn2+ will embed in MnO2 to provide capacity. The same energy storage mechanism was demonstrated by Chen et al. [12] for vanadium-based cathode materials.
Cathode   process :   MnO 2 + H 2 O + e MnOOH + OH
Zn 2 + + 2 e + 2 MnO 2 2 ZnMn 2 O 4
Anode   process :   Zn Zn 2 + + 2 e

1.3. Chemical Conversion Reaction Mechanism

Except for the aforementioned mechanisms, the chemical conversion reaction of α-MnO2 and MnOOH was reported by Liu et al. [13] in a Zn||α-MnO2 cell. MnOOH will form after α-MnO2 reacts with H+ in water during the discharge process, while the remaining OH in water will take a couple with ZnSO4 and H2O, finally leading to the formation of ZnSO4(Zn(OH)2)3·xH2O, during which the manganese ions in α-MnO2 will be converted from Mn4+ to Mn3+. MnOOH could be characterized by ex situ HRTEM with XRD tests, and zinc hydroxide sulfate in the discharge electrode was observed by NMR and XRD, presumably undergoing Equation (1.8) reactions [14]. The negative electrode undergoes the stripping and deposition of metallic zinc, and the reaction processes are as follows:
Cathode   process :   H 2 O H + + OH -
MnO 2 + H + + e - MnOOH
1 2 Zn 2 + + OH - + 1 6 ZnSO 4 + 6 x H 2 O 1 6 ZnSO 4 [ Zn ( OH ) 2 ] 3 . xH 2 O
Anode   process :   Zn Zn 2 + + 2 e
Due to different crystal structures of the cathode materials, the energy storage mechanisms are different from each other, and zinc ions enter the cathode through a complex reaction. However, the energy storage mechanism of some other materials is even controversial, and the detailed investigation of the mechanism still needs to be further explored, which is essential for the development of high-performance materials and practical application for ZIBs. Apart from reaction mechanism, another research hotspot is the Zn anode, which deeply affects the electrochemical behaviors of ZIBs as well as the commercial use. Compared with other metal ion batteries (Li, Na, and K) [15], zinc is more stable in an aqueous environment. Nevertheless, there are more severe problems in the cycling process, and the solution condition to the anode problem is of great significance to developing ZIBs.
Over the past decade, research on ZIBs continues to pile up, which helps to stimulate the development of ZIBs. However, many bottlenecks still remain to be solved. During the charge and discharge process of ZIBs, the uneven deposition caused by the “tip effect”, lattice matching degree, and ion concentration leads to zinc dendrites on the Zn anode surface, and “dead zinc” forms when it is de-embedded [16,17,18,19]. Meanwhile, side-effects are inevitable due to the inherent disadvantages of aqueous batteries, consuming the electrolyte and anode material and seriously affecting the cycle life of ZIBs [20,21]. The above two serious issues are the main reasons that limit the electrochemical performance for aqueous ZIBs. The common ideas to suppress zinc dendrites in ZIBs can be divided into the following aspects: structural design of zinc anode [22,23], surface modification [24,25], electrolyte optimization [26,27], and research on the multi-functional design of separators [28,29]. For example, a 3D fiber network was prepared by constructing zincophilic Sn nanoparticles with nitrogen-enriched hollow carbon spheres via the electrospinning method, which can inhibit hydrogen evolution reactions (HERs) and promote uniform zinc deposition by regulating the flux of zinc ions [30]. Fabrication of an artificial SEI layer at the negative zinc electrode can control Zn2+ deposition homogeneously and physically block penetration of zinc dendrites into the diaphragm. Wang et al. carbonized polymer coatings containing F to construct inorganic CF layers with an electronic insulation feature and high ionic conductivity on the collector copper foil, capable of guiding dendrite-free Zn deposition [31]. Kang et al. [32] prepared inorganic nano-CaCO3 coatings on the surface of the Zn anode to tune the uniform deposition of Zn2+ through nanopores [32]. Due to the complexity of structural design and interfacial modification and the difficulty of flexible application in realistic environments, researchers found that regulation of the solvation structure is more easily achieved. The structure design of the solvent can change the coordination environment of Zn2+ and the activity of water molecules to promote the stability as well as stabilize the electrochemical window of aqueous ZIBs. Herein, we attempt to summarize the current status of research from both a high-concentration strategy and additive engineering. Particularly, the high-concentration electrolyte strategy can alleviate the electrochemical window limitation of the ZIBs and inhibit the activity of the water molecules, which makes the solvation structure of Zn2+ reconfigured and is an effective way to adjust the solvation structure [33]. The electrolyte additive approach is facile, reliable, and highly efficient [34]. The existence of additives exhibits different interaction forces with Zn2+ and water molecules, which have significant effects on the tuning of the solvation structure. We believe that this review will give guidance for further investigations for high-performance ZIBs.

2. Solvation Structure

The concept of SEI film, first introduced in 1979, is a passivation layer covering the surface of the electrode material by a reaction between the electrode material and the solid–liquid phase interface of the electrolyte [35]. As an interfacial layer, this passivation layer has the characteristics of a solid electrolyte. However, with the in-depth study of SEI membranes, it is found that some phenomena cannot be fully explained by SEI theory. For example, when dimethyl sulfoxide (DMSO) was added to the Zn(TFSI)2 electrolyte solution [36], DMSO could significantly increase the proportion of TFSI anions in the inner solvation sheath of Zn2+. TFSI anions would be preferentially reduced before Zn deposition to form a fluorine-containing SEI layer in situ on the Zn anode. In this case, the synergistic mechanism of SEI theory and solvation theory work together to adjust the flux of Zn2+ to achieve dendrite-free Zn deposition. The addition of tetramethylurea (TMU) to Zn(OTf)2 was able to adjust the solvation structure of Zn2+ while forming a unique inorganic–organic bilayer SEI to improve the transport kinetics of Zn2+ and promote the uniform deposition of Zn2+ [37]. Due to this consideration, the introduction of the solvation structure theory can be well explained. The solvation theory can be traced back to 1981. In the past decades, the study of the solvation structure of electrolytes with solvent–solute interactions has gradually intensified [38,39,40,41,42]. Interaction between the cation/anion and the additives and solvent in aqueous electrolyte will form a relatively stable sheath-like structure, which is called the solvation structure [43]. In aqueous electrolytes, water molecules exist in two forms: solvation and free water. Compared with anions, metal cations are more easily combined with water molecules to form larger hydrated metal ions with a more muscular solvent sheath layer [43,44]. As can be noted from Figure 1b,c, compared to other metal ions, Zn2+ will form a larger radius of hydrated zinc ions than other metal ions and non-metal carriers given the smaller ionic radius (Li+, Na+, K+, Mg2+, Ca2+, and NH4+), and generate the solvation form of Zn[(H2O)6]2+ [45]. Significantly, the deposition process of Zn involves three steps and is presented in Figure 1a: (1) liquid-phase mass transfer, (2) Zn[(H2O)6]2+ desolvation, and (3) substrate nucleation. Particularly, the solvation H2O will not form hydrogen bonds at the electrode and electrolyte interfaces and can easily shed protons, which will be reduced more preferentially than Zn2+, thus inducing HER. Subsequently, the local pH value will increase in the electrolyte solution, causing side reactions and leading to anode corrosion (Figure 1d) [46,47,48].
The first two steps of the Zn deposition process mentioned above are related to the applied electrolyte, which are the critical factors affecting the zinc dendrites. Additionally, the solvent structure would have a significant effect on the performance of the ZIBs, which has been confirmed by more and more research. By adjusting the solvation structure of Zn2+, generation of Zn dendrites and side reactions as well as anodic corrosion can be feasibly suppressed so that it can diffuse in two dimensions on the electrode surface and electrolyte to realize homogeneous deposition of Zn2+. As illustrated in Figure 2, commonly used methods for optimizing electrolytes through structure design can be divided into two aspects including the high-salt-concentration strategy and electrolyte additives.

2.1. High-Concentration “Water-in-Salt” Strategy

Conventional aqueous electrolytes, influenced by the voltage of water decomposition, show a narrow electrochemical window (~1.23 V) [49]. The water activity will decrease and the electrochemical stability window will enlarge in aqueous electrolytes with the increase in salt concentration [50,51]. Meanwhile, the coordination environment of metal ions can be tuned and the number of solvation water molecules of metal cations can be reduced by employing high-concentration salt. The solvent sheath layer of metal ions is occupied by anions, which reduces the required energy for the desolvation of metal ions. The active water molecules reaching the electrode and electrolyte surface are reduced, which is able to prohibit the evolution of H2 and the corrosion phenomenon, and inhibit the occurrence of side reactions of the metal anode, effectively promoting the cycle life of ZIBs [52,53].
Given the advantages of low cost, non-toxicity, good solubility, and stable electrochemical properties of zinc sulfate (ZnSO4) in an aqueous solution, it is regarded as one of the common zinc salts at this stage and widely used in aqueous ZIBs. In the mixed electrolyte (ZnSO4+MnSO4), Raman spectra show that the broad peaks indicating free water molecules are gradually suppressed (Figure 3a), and Zn+-coordinated ambient water molecules are reduced as the electrolyte concentration of ZnSO4 increases from 2 M to 4.2 M (Figure 3b,c). It is able to restrain side reactions and improve the cyclic stability of zinc metal anodes. A half-cell of Zn||Cu showed low polarization in dilute electrolyte (2 M ZnSO4 + 0.1 M MnSO4) with Zn plating/stripping cycles over 1000 h at 0.2 mA cm−2 current density (Figure 3d) [54]; the average coulombic efficiency (ACE) for cycling >120 was 97.54%; while in 4.2 M ZnSO4 and 0.1 M MnSO4 mixed electrolyte, it was able to reach 99.21% (Figure 3g). Another full cell of Zn||MnO2 showed advanced long-cycle behavior under a measurement of more than 1200 cycles at 938 mA g−1 (Figure 3h), with an ACE close to 100% and 88.7% capacity retention. The dissolution of the cathode material would be inhibited via addition of MnSO4 and a synergistic effect with its solvation structure was created to achieve a stable, reversible, and high-performing Zn||MnO2 full cell [55]. The Zn anode in 3 M ZnSO4 electrolyte exhibited an impressively advanced coulombic efficiency (>99%) [56], excellent reversibility, as well as good rate capability, achieving >8000 cycles, which significantly prolonged the cycle life of Zn with increasing current density (Figure 3e,f).
Another example of zinc salt is Zn(TFSI)2 with high electrochemical stability and ionic conductivity. Wang and coworkers [15] introduced the concept of “water-in-salt” in ZIBs for the first time by using the electrolyte with high concentration (20 m LiTFSI + 1 m Zn(TFSI)2), resulting in an electrolyte pH of 7 (Figure 4a). According to the FTIR and 17O−NMR spectra (Figure 4b,c), it can be found that with increase in the concentration of Li ions, the water molecules tend to form solvent sheaths with Li ions more. When the concentration of LiTFSI ≥20 m (see in Figure 4d), the water molecules are strictly confined to the Li ions’ solvation structure, and such a solvent sheath layer of Zn ions is changed from the original partition to be occupied by anions, and TFSI surrounds Zn2+. The altered solvation structure of zinc ions avoids direct contact between the zinc metal anode and the water molecules of the solvent sheath layer, thus significantly inhibiting the HER phenomenon and the formation of by-products. The metallic Zn cathode exhibits superb electrochemical stability and deposition/stripping reversibility. The highly concentrated electrolyte results in dendrite-free zinc plating/stripping of about 100% coulombic efficiency with no observed by-product formation throughout the application. As for another highly concentrated solution with 20 m LiTFSI+ 1 m Zn(OTf)2 [57], Zn2+ prefers to combine with O(TFSI) compared to O(water). Compared with low-concentration electrolytes, it shows good reversibility of galvanizing/zinc stripping, less polarization in long cycles, and long cycle life (Figure 4e–h). It can form a stable SEI film (Figure 4i–j), which makes the galvanized Cu surface uniform and flat. The capacity retention of the Zn||LiMn2O4 full cell is 92% with an average CE about 99.62% after 300 cycles. A Zn||P(4VC86-stat-SS14) full cell exhibited an ultra-long lifetime in 4 M Zn(TFSI)2 electrolyte [58]. A high reversible capacity about 184 mAh g−1 was maintained with a capacity retention rate of 83% after 48,000 continuous cycles.
ZnCl2 with significant solubility is often chosen as a high-concentration electrolyte. A highly reversible full cell of V2O5||Zn was fabricated by Tang et al. [59] with a ZnCl2 “water-in-salt” electrolyte (WISE) to extend its life at high-temperature storage. V2O5 cathodes in dilute electrolytes experience severe lattice swelling from 4.4 to 13.9 Å (319%) due to the co-insertion of hydrated zinc ions, with reduced active material, resulting in capacity loss and reduced cycling performance. An ultrahigh capacity of about 302 mAh g−1 can be achieved by a V2O5||Zn full cell with conventional electrolyte (1 M ZnSO4), and the capacity decreases rapidly. In contrast, cycling in 30 m ZnCl2 exhibits a maximum capacity of 341 mAh g−1, which shows about a 9.5% decay after 300 cycles (Figure 5a,b), demonstrating ultra-high cycling stability within 1000 cycles (Figure 5c). The ZnCl2 WISE (30 m) can prevent co-insertion of water molecules, and inhibit dissolution of vanadium-based materials, resulting in an expanded electrochemical window, which inhibits the protrusion of zinc dendrites, and improves electrochemical performance. (Figure 5d). A mixed electrolyte was formed by adding 5 m LiCl to 30 m ZnCl2, in which hydrogen bonds between H2O can be broken by Li salt, reduce the free water molecules around Zn2+, and limit the solvation process of Zn2+ ions to achieve a higher CE of the mixed double-ion (5 m LiCl + 30 m ZnCl2) WISE than that of a single salt (30 m ZnCl2) [60]. Zhang and coworkers [61] found that the solubility of ZnCl2 was able to reach 31 m at room temperature and proposed using a 30 m ZnCl2 solution as an electrolyte for ZIBs to enhance the reversibility of the Zn anode. Distinct from 5 m ZnCl2 (Figure 5e), a symmetric cell of Zn||Zn exhibited excellent cycle stability when plating/stripping at 0.2 mA cm−2 for 10 min in a 30 m ZnCl2 electrolyte solution. The viscosity of the electrolyte solution increases with concentration increases and electrical conductivity decreases. However, the ion migration number of Zn2+ is higher in the highly concentrated electrolyte than that in the lowly concentrated one. A lower degree of hydrolysis slows down Zn(OH)2/ZnO formation (Figure 5f), and the electrochemical window and pH also increase with concentration, indicating that the hydrolysis of Zn2+ is inhibited (Figure 5g) and inhibits the HER, which brings higher reversibility of the zinc metal anode as well as tunes the uniform deposition of Zn2+ (Figure 5h,i).
Table 1 summarizes the electrochemical properties of zinc anode using highly concentrated electrolytes. With the development of ZIBs, a range of zinc salts used in ZIBs were investigated, including ZnCl2, ZnSO4, Zn(Ac)2, ZnF2, Zn(TFSI)2, and Zn(CF3SO3)2 [62,63,64,65]. In summary, the addition of highly concentrated salt solutions can, on the one hand, directly reduce the number of water molecules, disrupt the coordination equilibrium of Zn2+ in aqueous solvent, and replace the H2O molecules in the Zn2+ solvation structure. On the other hand, interaction between anionic or cationic water molecules of some highly concentrated salt solutions is greater than that between water molecules and Zn2+, and Zn2+ shares the solvation, reducing the energy required for the desolvation of Zn2+. Alternatively, a new solvation sheath can be formed directly to suppress the HER phenomenon and regulate the deposition of Zn2+ homogeneously. However, although highly concentrated electrolytes can promote the coulombic efficiency of Zn plating/stripping and are beneficial for the deposition of dendrite-free Zn, some unavoidable problems still remain to be solved. The concentrated electrolyte is highly polarized even at 0.2 mA cm−2 and 0.03 mAh [66,67]. Its viscosity increases with the concentration [61,64], resulting in reduced ionic conductivity of the electrolyte. In addition, the high cost of the concentrated electrolyte can raise the cost of ZIBs, losing the advantage of the low cost of aqueous batteries. All these issues can limit the practical application of highly concentrated electrolytes in ZIBs. Another neglected point is the effect of the high concentration strategy on the cathode material. ZIB cathode materials suffer from the issues of material dissolution and structural collapse. However, based on the complexity of the energy storage mechanism of cathode materials, the current focus of research hotspots is the modification of the anode, but not much research is focused on the impact of high concentration on the cathode materials, so more in-depth research is in urgent need.

2.2. Functional Additives

As mentioned above, side reactions and Zn dendrites are the main issues that need to be addressed in ZIBs, which can result in corrosion, HER, and passivation. Different types of functional electrolyte additives can inhibit side reactions and growth of zinc dendrites by different mechanisms. In recent years, the role of additive engineering has been extensively studied in ZIBs.
Tetrabutylammonium sulfate (TBA2SO4) is an effective, low-cost, non-toxic cationic surface-activator-type electrolyte additive. Zhu et al. [71] proposed to add TBA2SO4 to the 2 M ZnSO4 electrolyte with a small amount about 0.029 g L−1, and the positively charged TBA+ gathered at the negatively charged zinc “bulge” by electrostatic adsorption, forming an electrostatic shield and limiting the tip effect. TBA+ causes Zn2+ deposition in adjacent flat areas (Figure 6a,b), inducing uniform zinc deposition through a unique zincophobic repulsion mechanism. A full cell of Zn||MnO2 with TBA2SO4 exhibits a cycle stability of about 300 times at 1 A g−1 with a 94% capacity retention. Wang et al. [62] found that by adding 2 v.% of diethyl ether (Et2O), Zn2+ will shift to other regions during zinc deposition. Zn||MnO2 cells in 0.1 M Mn(CF3SO3)2 and 3 M Zn(CF3SO3)2 mixed aqueous solution with Et2O as an additive were able to cycle 4000 times at 5 A g−1 with a 97.7% capacity retention rate (Figure 6c,d), showing excellent cycling performance. Because Et2O molecules can preferentially adsorb on the prominent tip, eliminating the “tip effect” and flattening the zinc anode surface (Figure 6e), it can largely restrain growth of zinc dendrites. Guo et al. [63] added a small amount (25 × 10−3 M) of Zn(H2PO4)2 salt to 1 M Zn(CF3SO3)2 electrolyte. From the Raman spectrum, it can be seen that SEI−Zn has bonds between Zn and O, and P and O, compared with bare−Zn, which demonstrates the formation of the SEI (Figure 6f). As shown in Figure 6g, Zn(H2PO4)2 can generate a dense, stable, and highly Zn2+−conducting SEI film on the negative surface, ensuring uniform and rapid Zn2+ migration and achieving dendrite-free growth of zinc metal, which also is confirmed by the XRD pattern (Figure 6h). Even if the coating breaks down, the Zn(H2PO4)2 in electrolyte will still form a new interfacial layer, thus solving the coating failure problem. Under actual test conditions with Zn(H2PO4)2 addition, the Zn||V2O5 full cell exhibits a capacity retention of about 94.4% after 500 cycles.
The abovementioned additives can be broadly classified as ionic and nonionic, capable of achieving dendrite-free zinc deposition by electrostatic shielding or by forming SEI films. Improving the performance of ZIBs also requires avoiding side reactions. In the aqueous electrolyte, the stable form of Zn2+ is [Zn(H2O)6]2+ requires high energy for desolvation during the deposition process. Researchers have found that additives can adjust the solvation structure. Replacing H2O molecules in [Zn(H2O)6]2+ with other ions or molecules with weaker solvation structure effects can reduce the desolvation energy as well as prevent the side reactions [46].
Much research related to organic substances as common additives already exists. Li et al. [72] added ethylene glycol (EG) to the ZnSO4 electrolyte and found that it is effective in improving Zn2+ reversible deposition (Figure 7a). The red-shift in the OH stretches H2O (3200~3400 cm−1), and OH stretched bending of C−OH (1460~1465 cm−1) shows a blue shift and can be noted from the Raman spectrum (Figure 7b), indicating that the electron density of OH in H2O is lower and the electron density of C−OH in EG is higher, and the hydroxyl group of EG has a solid force to generate hydrogen bonds with H2O, which can effectively weaken the force between Zn2+ and H2O. Experiments and theoretical calculations show that with the increase in EG content, the OH stretching vibration range from 3000 to 3500 cm−1 can lead to a significant redshift, and H2O bending vibration in the range of 1600~1700 cm−1 is slightly blueshifted in the FTIR spectrum (Figure 7c) [73,74]. Pan et al. [75] used ab initio calculations to analyze the solvation energy of different Zn2+ solvation sheaths in ethylene glycol/water solutions. They found that the solvation energy was ranked as [Zn(H2O)m (EG)n]2+ < [Zn(H2O)6 ]2+ < [Zn(EG)3 ]2+, indicating that the interaction force between EG and zinc ions was greater than that between Zn2+ and water, disrupting the solvation of Zn2+ and effectively inhibiting the occurrence of the HER (Figure 7d). As shown in Figure 7e, the Zn||Zn cell with a 3 M ZnSO4/H2O/68 v.% EG electrolyte achieves an extremely long cycle life of 2668 h at 0.5 mA cm−2. It is attributed that the solvation structure of Zn2+ can be controlled by the EG molecule. Figure 7f shows that the radius of [Zn(EG)3 ]2+ is larger than that of [Zn(H2O)6 ]2+, which increases the spatial site resistance of Zn2+ transfer, limits the diffusion of Zn2+, and increases the over potential of zinc deposition, thus refining the grain size of zinc as well as avoiding rapid growth of zinc to produce dendrites [76,77]. A study found the addition of dimethyl sulfoxide (DMSO) to ZnCl2 solution [78] (Figure 7g). H2O in a zinc ion solvation sheath can be replaced by DMSO because of its (29.8) higher Gutmann donor number and larger dielectric constant than that of H2O (18), and O had a smaller electron density in DMSO [79,80]. With the addition of DMSO, interaction between solvation water and Zn2+ can be weakened, which facilitates the desolvation process of zinc ions and can suppress the decomposition of active H2O, improving the zinc plating and stripping coulombic efficiency. The solvation DMSO can lead to SEI layer formation in situ on the anode surface, which allows passage of zinc ions and hinders the passage of water molecules (Figure 7h). The Zn||Zn symmetric cell in such a mixed electrolyte exhibited a cycle life of more than 1000 h. The cycle life was improved by a factor of 2.5 compared to that of the ZnCl2−H2O symmetric cell. Zn||Ti half-cells increased the CE to 99.0% and finally 100% within 30 cycles, exhibiting a stable over potential of ~24 mV, confirming the ability of the DMSO additive in improving the CE of plating/stripping. Qiao et al. [81] added 50 v.% methanol to the ZnSO4 electrolyte. They found that methanol can insert into the internal Zn2+ solvation sheath owing to the high dielectric constant and small size (37.2) [82,83] and can disrupt the coordination equilibrium of Zn2+ by adjusting the amount of methanol addition (Figure 7i). The methanol molecules will first interact with H2O in the first layer of the Zn2+ solvated sheath layer and form hydrogen bonds. In addition, the high wettability between methanol and the Zn metal anode makes it easier for methanol molecules to adsorb on the surface of Zn metal than solvent water, inducing the realization of dendrite-free Zn2+ uniform deposition (Figure 7j). When the size of methanol molecules becomes larger, they will eventually insert into the inner layer of the Zn2+ solvent sheath, the activity of water will decrease, and both hydrogen precipitation reactions and side reactions will be inhibited (Figure 7k,l). Yang et al. [84] demonstrated by introducing N-methyl-2-pyrrolidone (NMP) polar additives to ZnSO4 that the addition of organic solvents containing carbonyl groups contributes to stabilizing the hydrogen bonding network of water and the structural remodeling of Zn2+−solvation, and such a synergistic effect helps to inhibit dendrite formation and water-induced parasitic reactions.
By investigating the influence of different organic additives on the Zn2+ solvation structure, researchers have provided an electrolyte regulation approach to achieve highly reversible Zn anodes and cells. Except for organic additives, salt additives are considered as another effective way to promote the electrochemical performance of Zn anode. Zhang et al. [90] used trifluoromethanesulfonate (OTf) anions and ethylenediaminetetraacetic acid (Y4−) anions to form a mixed electrolyte (BE + 100 mM) consisting of 1 M Zn(OTf)2 + 100 mM Na4Y, and the double anion can change the Zn2+ solvation sheath to stabilize the Zn anode (Figure 8a–c). As shown in Figure 8d–f, this coordination environment can reduce the solvation water content and activity, inhibiting HER, and homogenizing the Zn2+ plating process. In addition, this solvation structure allows decomposition of ligand anions to generate an in situ organic–inorganic SEI on the Zn surface. The resulting Zn2+ conducting SEI layer can lead to separation between Zn and the electrolyte, thereby inhibiting side reactions derived from H2O. As a result, the Zn||Zn cell has a coulomb efficiency of 99.7% at 1.0 mA cm−2 and 1.0 mAh cm−2 for more than 1600 h, demonstrating excellent long-term stability (Figure 8g). The Zn||Cu cell is highly reversible, possessing a coulomb efficiency of 99.7% at more than 400 cycles at 1.0 mA cm−2 (Figure 8h). Chen et al. [91] demonstrated that via adding halogen ions to the Zn2+ solvation structure, the issues of dendrite growth and hydrogen precipitation reactions could be overcome. Through electrolyte regulation consisting of ammonium halide and zinc acetate, I can combine with Zn2+ to convert the conventional Zn(H2O)62+ to ZnI(H2O)5+, where I can transfer electrons to H2O, and water molecules in the solvation structure would be reduced by inhibiting the reduction in the lowest unoccupied molecular orbital (LUMO) energy level and reducing the electron loss, thus improving the reduction stability and, thus, inhibiting HER (Figure 8i). The formation of a dynamic electrostatic shielding layer accompanied with NH4+ can inhibit the growth of dendrites (Figure 8j–k).
The electrochemical performances of zinc anode with various electrolyte additives are summarized in Table 2. In summary, electrolyte additives have obvious effects inhibiting side reactions and dendrites of ZIBs. Significantly, the selection of additives should consider three aspects: (1) Gutmann donor number of solvent additives, which can replace H2O in the solvation sheath of Zn2+ when the donor number is larger than H2O and can maintain Zn2+ ion transfer kinetics to some extent [92]. (2) Interaction with H2O, which can form hydrogen bonds with H2O, reduce H2O activity, and inhibit H2O reduction. (3) In situ formation of SEI film, which should form a dense, self-healing Zn2+-conducting SEI film to prevent water penetration into the zinc anode. Furthermore, the electrolyte additives also have a vital effect on the cathode material. For example, the addition of triethyl phosphate (TEP) in the Zn(OTF)2-H2O electrolyte in Zn||V2O5 batteries can keep the pH of the electrolyte solution stable and hinder the dissolution of V2O5 to some extent, which has strong interactions with Zn2+ and H2O molecules. The TEP will enter the inner layer of the solvation sheath of Zn2+ and break the hydrogen bonding network between water molecules, thus reducing the activity of H2O and avoiding the reaction between V2O5 and H2O [93]. Another report mentioned that introducing polyethylene glycol (PEG) into the electrolyte can reduce the content of free water molecules and inhibit the H+ insertion to the LiV2(PO4)3 (LVP) cathode through the strong interaction between PEG and H2O, which is beneficial for suppressing the phase change of LVP. Significantly, much attention should also be paid to the cathode materials for the development of water-based ZIBs.

3. Challenges and Future Development Prospects

This paper reviews two approaches to optimize the electrolyte from a solvation structure perspective, and summarizes the effects on the anode in terms of both the inhibition of zinc dendrites and reduction in side reactions. The current research on high-concentration strategies and electrolyte additives has achieved some results, but they are far from sufficient. For the high-concentration strategy of ZIBs, the mechanism affecting the evolutionary transformation of the solvation structure needs further insight through some in situ characterization and theoretical calculations. In addition, based on the practicality of the high-concentration strategy, it is necessary to explore more suitable salt solutions, including high ionic conductivity, good thermodynamic stability, low cost, ability to achieve compatibility with both poles of the battery system, and a good balance between high concentration and viscosity. ZIBs work differently with various electrolyte systems (alkaline, neutral, and weak acid) with different cathode materials for electrolyte additives. The working mechanism of electrolyte additives and their effects on cathode materials need to be further disclosed. In addition, some of the electrolyte additive types studied so far are costly and toxic, increasing the battery system’s flammability. The type of electrolyte and its mechanism of action need to be considered when adjusting the solvation structure of Zn2+.
The research on the electrolyte engineering for ZIBs has yet to mature and is more in the laboratory stage. In view of the existing problems, the future research directions of electrolyte engineering can include the following: (1) The influence of the electrolyte system on the cathode material. The working mechanism of the electrolyte is based on the energy storage mechanism of the battery system because the energy storage mechanism is different for various cathode materials. Research targeting cathode materials will exhibit great potential. (2) Development of quasi-solid or all-solid electrolytes. In order to avoid the hazards caused by active water molecules and with the development of flexible devices, it is promising to explore solid or gel electrolytes with environmental friendliness, low cost, high ionic conductivity, and non-toxicity. (3) Exploring new multifunctional electrolyte additives. Current research on electrolyte additives shows the great potential of electrolyte additives. Research to explore more different additives is necessary, depending on the demand. Despite many difficulties visible to the naked eye, we firmly believe that water-based ZIBs will continue to evolve and provide a viable path for future energy storage methods.

Author Contributions

Conceptualization, W.Q. and C.W.; methodology, B.W.; software, B.W.; investigation, B.W.; resources, J.H. and J.D.; data curation, W.Q., C.W. and Z.M.; writing—original draft preparation, B.W. and H.X.; writing—review and editing, W.Q., C.W. and Z.M.; supervision, J.H. and J.D.; project administration, W.Q., C.W.; funding acquisition, W.Q., C.W. and Z.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Natural Science Foundation of Hunan Province (Project Nos. 2020JJ5580; 2020JJ5563); the Scientific Research Foundation of Hunan Provincial Education Department (Contract No. 21B0328); the Scientific and Technological Plan Projects of Guangzhou City (No. 202103040001); the Fellowship of China Postdoctoral Science Foundation (No. 2020M672670).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Wu, C.; Du, J.; Zhu, Y.; Qin, W.; Wang, X.; Jia, C.; Zhang, K. Highly stable 3D hierarchical manganese sulfide multi-layer nanoflakes with excellent electrochemical performances for supercapacitor electrodes. J. Alloys Compd. 2022, 894, 162390. [Google Scholar] [CrossRef]
  2. Xing, T.; Ouyang, Y.; Chen, Y.; Zheng, L.; Wu, C.; Wang, X. P-doped ternary transition metal oxide as electrode material of asymmetric supercapacitor. J. Energy Storage 2020, 28, 101248. [Google Scholar] [CrossRef]
  3. Wu, C.; Zhu, Y.; Guan, C.; Jia, C.; Qin, W.; Wang, X.; Zhang, K. Mesoporous aluminium manganese cobalt oxide with pentahedron structures for energy storage devices. J. Mater. Chem. A 2019, 7, 18417–18427. [Google Scholar] [CrossRef]
  4. Fang, G.; Zhou, J.; Pan, A.; Liang, S. Recent Advances in Aqueous Zinc-Ion Batteries. ACS Energy Lett. 2018, 3, 2480–2501. [Google Scholar] [CrossRef]
  5. Liu, N.; Li, B.; He, Z.; Dai, L.; Wang, H.; Wang, L. Recent advances and perspectives on vanadium- and manganese-based cathode materials for aqueous zinc ion batteries. J. Energy Chem. 2021, 59, 134–159. [Google Scholar] [CrossRef]
  6. Wu, Y.; Song, T.-Y.; Chen, L.-N. A review on recent developments of vanadium-based cathode for rechargeable zinc-ion batteries. Tungsten 2021, 3, 289–304. [Google Scholar] [CrossRef]
  7. Yamamoto, T.; Shoji, T. Rechargeable Zn∣ZnSO4∣MnO2-type cells. Inorg. Chim. Acta 1986, 117, L27–L28. [Google Scholar] [CrossRef]
  8. Xu, C.; Li, B.; Du, H.; Kang, F. Energetic Zinc Ion Chemistry: The Rechargeable Zinc Ion Battery. Angew. Chem. Int. Ed. 2012, 51, 933–935. [Google Scholar] [CrossRef]
  9. Chen, D.; Lu, M.; Cai, D.; Yang, H.; Han, W. Recent advances in energy storage mechanism of aqueous zinc-ion batteries. J. Energy Chem. 2021, 54, 712–726. [Google Scholar] [CrossRef]
  10. Chen, L.; Ruan, Y.; Zhang, G.; Wei, Q.; Jiang, Y.; Xiong, T.; He, P.; Yang, W.; Yan, M.; An, Q.; et al. Ultrastable and High-Performance Zn/VO2 Battery Based on a Reversible Single-Phase Reaction. Chem. Mater. 2019, 31, 699–706. [Google Scholar] [CrossRef]
  11. Sun, W.; Wang, F.; Hou, S.; Yang, C.; Fan, X.; Ma, Z.; Gao, T.; Han, F.; Hu, R.; Zhu, M.; et al. Zn/MnO2 Battery Chemistry With H+ and Zn2+ Coinsertion. J. Am. Chem. Soc. 2017, 139, 9775–9778. [Google Scholar] [CrossRef] [PubMed]
  12. Wan, F.; Zhang, L.; Dai, X.; Wang, X.; Niu, Z.; Chen, J. Aqueous rechargeable zinc/sodium vanadate batteries with enhanced performance from simultaneous insertion of dual carriers. Nat. Commun. 2018, 9, 1656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Liu, Z.; Pulletikurthi, G.; Endres, F. A Prussian Blue/Zinc Secondary Battery with a Bio-Ionic Liquid–Water Mixture as Electrolyte. ACS Appl. Mater. Interfaces 2016, 8, 12158–12164. [Google Scholar] [CrossRef] [PubMed]
  14. Song, M.; Tan, H.; Chao, D.; Fan, H.J. Recent Advances in Zn-Ion Batteries. Adv. Funct. Mater. 2018, 28, 1802564. [Google Scholar] [CrossRef]
  15. Wang, F.; Borodin, O.; Gao, T.; Fan, X.; Sun, W.; Han, F.; Faraone, A.; Dura, J.A.; Xu, K.; Wang, C. Highly reversible zinc metal anode for aqueous batteries. Nat. Mater. 2018, 17, 543–549. [Google Scholar] [CrossRef] [PubMed]
  16. Barton, J.L.; Bockris, J.O.M. The Electrolytic Growth of Dendrites from Ionic Solutions. Proc. R. Soc. Lond. A Math. Phys. Sci. 1962, 268, 485–505. [Google Scholar]
  17. Yang, Q.; Liang, G.; Guo, Y.; Liu, Z.; Yan, B.; Wang, D.; Huang, Z.; Li, X.; Fan, J.; Zhi, C. Do Zinc Dendrites Exist in Neutral Zinc Batteries: A Developed Electrohealing Strategy to In Situ Rescue In-Service Batteries. Adv. Mater. 2019, 31, 1903778. [Google Scholar] [CrossRef]
  18. Zhang, W.; Zhuang, H.L.; Fan, L.; Gao, L.; Lu, Y. A “cation-anion regulation” synergistic anode host for dendrite-free lithium metal batteries. Sci. Adv. 2018, 4, eaar4410. [Google Scholar] [CrossRef] [Green Version]
  19. Xie, C.; Li, Y.; Wang, Q.; Sun, D.; Tang, Y.; Wang, H. Issues and solutions toward zinc anode in aqueous zinc-ion batteries: A mini review. Carbon Energy 2020, 2, 540–560. [Google Scholar] [CrossRef]
  20. Tang, B.; Shan, L.; Liang, S.; Zhou, J. Issues and opportunities facing aqueous zinc-ion batteries. Energy Environ. Sci. 2019, 12, 3288–3304. [Google Scholar] [CrossRef]
  21. Pan, H.; Shao, Y.; Yan, P.; Cheng, Y.; Han, K.S.; Nie, Z.; Wang, C.; Yang, J.; Li, X.; Bhattacharya, P.; et al. Reversible aqueous zinc/manganese oxide energy storage from conversion reactions. Nat. Energy 2016, 1, 16039. [Google Scholar] [CrossRef]
  22. Wang, W.; Huang, G.; Wang, Y.; Cao, Z.; Cavallo, L.; Hedhili, M.N.; Alshareef, H.N. Organic Acid Etching Strategy for Dendrite Suppression in Aqueous Zinc-Ion Batteries. Adv. Energy Mater. 2022, 12, 2102797. [Google Scholar] [CrossRef]
  23. Liu, P.; Liu, W.; Liu, K. Rational modulation of emerging MXene materials for zinc-ion storage. Carbon Energy 2022, 4, 60–76. [Google Scholar] [CrossRef]
  24. Zhang, Z.; Said, S.; Smith, K.; Zhang, Y.S.; He, G.; Jervis, R.; Shearing, P.R.; Miller, T.S.; Brett, D.J.L. Dendrite suppression by anode polishing in zinc-ion batteries. J. Mater. Chem. A 2021, 9, 15355–15362. [Google Scholar] [CrossRef]
  25. Yang, S.; Li, Y.; Du, H.; Liu, Y.; Xiang, Y.; Xiong, L.; Wu, X.; Wu, X. Copper Nanoparticle-Modified Carbon Nanofiber for Seeded Zinc Deposition Enables Stable Zn Metal Anode. ACS Sustain. Chem. Eng. 2022, 10, 12630–12641. [Google Scholar] [CrossRef]
  26. Vijayakumar, V.; Ghosh, M.; Kurian, M.; Torris, A.; Dilwale, S.; Badiger, M.V.; Winter, M.; Nair, J.R.; Kurungot, S. An In Situ Cross-Linked Nonaqueous Polymer Electrolyte for Zinc-Metal Polymer Batteries and Hybrid Supercapacitors. Small 2020, 16, 2002528. [Google Scholar] [CrossRef] [PubMed]
  27. Han, L.; Huang, H.; Li, J.; Zhang, X.; Yang, Z.; Xu, M.; Pan, L. A novel redox bromide-ion additive hydrogel electrolyte for flexible Zn-ion hybrid supercapacitors with boosted energy density and controllable zinc deposition. J. Mater. Chem. A 2020, 8, 15042–15050. [Google Scholar] [CrossRef]
  28. Fang, Y.; Xie, X.; Zhang, B.; Chai, Y.; Lu, B.; Liu, M.; Zhou, J.; Liang, S. Regulating Zinc Deposition Behaviors by the Conditioner of PAN Separator for Zinc-Ion Batteries. Adv. Funct. Mater. 2022, 32, 2109671. [Google Scholar] [CrossRef]
  29. Yang, X.; Wu, W.; Liu, Y.; Lin, Z.; Sun, X. Chitosan modified filter paper separators with specific ion adsorption to inhibit side reactions and induce uniform Zn deposition for aqueous Zn batteries. Chem. Eng. J. 2022, 450, 137902. [Google Scholar] [CrossRef]
  30. Yu, H.; Zeng, Y.; Li, N.W.; Luan, D.; Yu, L.; Lou, X.W. Confining Sn nanoparticles in interconnected N-doped hollow carbon spheres as hierarchical zincophilic fibers for dendrite-free Zn metal anodes. Sci. Adv. 2022, 8, eabm5766. [Google Scholar] [CrossRef]
  31. Wang, H.; Chen, Y.; Yu, H.; Liu, W.; Kuang, G.; Mei, L.; Wu, Z.; Wei, W.; Ji, X.; Qu, B.; et al. A Multifunctional Artificial Interphase with Fluorine-Doped Amorphous Carbon layer for Ultra-Stable Zn Anode. Adv. Funct. Mater. 2022, 32, 2205600. [Google Scholar] [CrossRef]
  32. Kang, L.; Cui, M.; Jiang, F.; Gao, Y.; Luo, H.; Liu, J.; Liang, W.; Zhi, C. Nanoporous CaCO3 Coatings Enabled Uniform Zn Stripping/Plating for Long-Life Zinc Rechargeable Aqueous Batteries. Adv. Energy Mater. 2018, 8, 1801090. [Google Scholar] [CrossRef]
  33. Wang, Y.; Meng, X.; Sun, J.; Liu, Y.; Hou, L. Recent Progress in “Water-in-Salt” Electrolytes Toward Non-lithium Based Rechargeable Batteries. Front. Chem. 2020, 8, 595. [Google Scholar] [CrossRef]
  34. Zhai, C.; Zhao, D.; He, Y.; Huang, H.; Chen, B.; Wang, X.; Guo, Z. Electrolyte Additive Strategies for Suppression of Zinc Dendrites in Aqueous Zinc-Ion Batteries. Batteries 2022, 8, 153. [Google Scholar] [CrossRef]
  35. Peled, E. The Electrochemical Behavior of Alkali and Alkaline Earth Metals in Nonaqueous Battery Systems—The Solid Electrolyte Interphase Model. J. Electrochem. Soc. 1979, 126, 2047–2051. [Google Scholar] [CrossRef]
  36. Jian, Q.; Wang, T.; Sun, J.; Wu, M.; Zhao, T. In-situ construction of fluorinated solid-electrolyte interphase for highly reversible zinc anodes. Energy Storage Mater. 2022, 53, 559–568. [Google Scholar] [CrossRef]
  37. Yang, J.; Zhang, Y.; Li, Z.; Xu, X.; Su, X.; Lai, J.; Liu, Y.; Ding, K.; Chen, L.; Cai, Y.-P.; et al. Three Birds with One Stone: Tetramethylurea as Electrolyte Additive for Highly Reversible Zn-Metal Anode. Adv. Funct. Mater. 2022, 32, 2209642. [Google Scholar] [CrossRef]
  38. Cheng, H.; Sun, Q.; Li, L.; Zou, Y.; Wang, Y.; Cai, T.; Zhao, F.; Liu, G.; Ma, Z.; Wahyudi, W.; et al. Emerging Era of Electrolyte Solvation Structure and Interfacial Model in Batteries. ACS Energy Lett. 2022, 7, 490–513. [Google Scholar] [CrossRef]
  39. Alvarado, J.; Schroeder, M.A.; Pollard, T.P.; Wang, X.; Lee, J.Z.; Zhang, M.; Wynn, T.; Ding, M.; Borodin, O.; Meng, Y.S.; et al. Bisalt ether electrolytes: A pathway towards lithium metal batteries with Ni-rich cathodes. Energy Environ. Sci. 2019, 12, 780–794. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, A.; Kadam, S.; Li, H.; Shi, S.; Qi, Y. Review on modeling of the anode solid electrolyte interphase (SEI) for lithium-ion batteries. Npj Comput. Mater. 2018, 4, 15. [Google Scholar] [CrossRef] [Green Version]
  41. Ming, J.; Cao, Z.; Wahyudi, W.; Li, M.; Kumar, P.; Wu, Y.; Hwang, J.-Y.; Hedhili, M.N.; Cavallo, L.; Sun, Y.-K.; et al. New Insights on Graphite Anode Stability in Rechargeable Batteries: Li Ion Coordination Structures Prevail Over Solid Electrolyte Interphases. ACS Energy Lett. 2018, 3, 335–340. [Google Scholar] [CrossRef]
  42. Miertuš, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilizaion of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  43. Li, X.; Wang, X.; Ma, L.; Huang, W. Solvation Structures in Aqueous Metal-Ion Batteries. Adv. Energy Mater. 2022, 12, 2202068. [Google Scholar] [CrossRef]
  44. von Wald Cresce, A.; Gobet, M.; Borodin, O.; Peng, J.; Russell, S.M.; Wikner, E.; Fu, A.; Hu, L.; Lee, H.-S.; Zhang, z.; et al. Anion Solvation in Carbonate-Based Electrolytes. J. Phys. Chem. C 2015, 119, 27255–27264. [Google Scholar] [CrossRef]
  45. Jia, X.; Liu, C.; Neale, Z.G.; Yang, J.; Cao, G. Active Materials for Aqueous Zinc Ion Batteries: Synthesis, Crystal Structure, Morphology, and Electrochemistry. Chem. Rev. 2020, 120, 7795–7866. [Google Scholar] [CrossRef]
  46. Li, M.; Li, Z.; Wang, X.; Meng, J.; Liu, X.; Wu, B.; Han, C.; Mai, L. Comprehensive understanding of the roles of water molecules in aqueous Zn-ion batteries: From electrolytes to electrode materials. Energy Environ. Sci. 2021, 14, 3796–3839. [Google Scholar] [CrossRef]
  47. Cao, J.; Zhang, D.; Zhang, X.; Zeng, Z.; Qin, J.; Huang, Y. Strategies of regulating Zn2+ solvation structures for dendrite-free and side reaction-suppressed zinc-ion batteries. Energy Environ. Sci. 2022, 15, 499–528. [Google Scholar] [CrossRef]
  48. Dou, Q.; Yao, N.; Pang, W.K.; Park, Y.; Xiong, P.; Han, X.; Rana, H.H.; Chen, X.; Fu, Z.-H.; Thomsen, L.; et al. Unveiling solvation structure and desolvation dynamics of hybrid electrolytes for ultralong cyclability and facile kinetics of Zn–Al alloy anodes. Energy Environ. Sci. 2022, 15, 4572–4583. [Google Scholar] [CrossRef]
  49. Smith, L.; Dunn, B. Opening the window for aqueous electrolytes. Science 2015, 350, 918. [Google Scholar] [CrossRef]
  50. Suo, L.; Borodin, O.; Gao, T.; Olguin, M.; Ho, J.; Fan, X.; Luo, C.; Wang, C.; Xu, K. “Water-in-salt” electrolyte enables high-voltage aqueous lithium-ion chemistries. Science 2015, 350, 938–943. [Google Scholar] [CrossRef]
  51. Jiang, L.; Liu, L.; Yue, J.; Zhang, Q.; Zhou, A.; Borodin, O.; Suo, L.; Li, H.; Chen, L.; Xu, K.; et al. High-Voltage Aqueous Na-Ion Battery Enabled by Inert-Cation-Assisted Water-in-Salt Electrolyte. Adv. Mater. 2020, 32, 1904427. [Google Scholar] [CrossRef] [PubMed]
  52. Yuan, L.; Hao, J.; Kao, C.-C.; Wu, C.; Liu, H.-K.; Dou, S.-X.; Qiao, S.-Z. Regulation methods for the Zn/electrolyte interphase and the effectiveness evaluation in aqueous Zn-ion batteries. Energy Environ. Sci. 2021, 14, 5669–5689. [Google Scholar] [CrossRef]
  53. Liu, S.; Liu, J. Reshaping Electrolyte Solvation Structure for High-Energy Aqueous Batteries. Energy Environ. Mater. 2022, 5, 686–687. [Google Scholar] [CrossRef]
  54. Olbasa, B.W.; Fenta, F.W.; Chiu, S.-F.; Tsai, M.-C.; Huang, C.-J.; Jote, B.A.; Beyene, T.T.; Liao, Y.-F.; Wang, C.-H.; Su, W.-N.; et al. High-Rate and Long-Cycle Stability with a Dendrite-Free Zinc Anode in an Aqueous Zn-Ion Battery Using Concentrated Electrolytes. ACS Appl. Energy Mater. 2020, 3, 4499–4508. [Google Scholar] [CrossRef]
  55. Zhang, N.; Cheng, F.; Liu, J.; Wang, L.; Long, X.; Liu, X.; Li, F.; Chen, J. Rechargeable aqueous zinc-manganese dioxide batteries with high energy and power densities. Nat. Commun. 2017, 8, 405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Glatz, H.; Tervoort, E.; Kundu, D. Unveiling Critical Insight into the Zn Metal Anode Cyclability in Mildly Acidic Aqueous Electrolytes: Implications for Aqueous Zinc Batteries. ACS Appl. Mater. Interfaces 2020, 12, 3522–3530. [Google Scholar] [CrossRef]
  57. Clarisza, A.; Bezabh, H.K.; Jiang, S.-K.; Huang, C.-J.; Olbasa, B.W.; Wu, S.-H.; Su, W.-N.; Hwang, B.J. Highly Concentrated Salt Electrolyte for a Highly Stable Aqueous Dual-Ion Zinc Battery. ACS Appl. Mater. Interfaces 2022, 14, 36644–36655. [Google Scholar] [CrossRef]
  58. Patil, N.; de la Cruz, C.; Ciurduc, D.; Mavrandonakis, A.; Palma, J.; Marcilla, R. An Ultrahigh Performance Zinc-Organic Battery using Poly(catechol) Cathode in Zn(TFSI)2-Based Concentrated Aqueous Electrolytes. Adv. Energy Mater. 2021, 11, 2100939. [Google Scholar] [CrossRef]
  59. Tang, X.; Wang, P.; Bai, M.; Wang, Z.; Wang, H.; Zhang, M.; Ma, Y. Unveiling the Reversibility and Stability Origin of the Aqueous V2O5–Zn Batteries with a ZnCl2 “Water-in-Salt” Electrolyte. Adv. Sci. 2021, 8, 2102053. [Google Scholar] [CrossRef]
  60. Zhang, C.; Shin, W.; Zhu, L.; Chen, C.; Neuefeind, J.C.; Xu, Y.; Allec, S.I.; Liu, C.; Wei, Z.; Daniyar, A.; et al. The electrolyte comprising more robust water and superhalides transforms Zn-metal anode reversibly and dendrite-free. Carbon Energy 2021, 3, 339–348. [Google Scholar] [CrossRef]
  61. Zhang, C.; Holoubek, J.; Wu, X.; Daniyar, A.; Zhu, L.; Chen, C.; Leonard, D.P.; Rodríguez-Pérez, I.A.; Jiang, J.-X.; Fang, C.; et al. A ZnCl2 water-in-salt electrolyte for a reversible Zn metal anode. Chem. Commun. 2018, 54, 14097–14099. [Google Scholar] [CrossRef] [PubMed]
  62. Xu, W.; Zhao, K.; Huo, W.; Wang, Y.; Yao, G.; Gu, X.; Cheng, H.; Mai, L.; Hu, C.; Wang, X. Diethyl ether as self-healing electrolyte additive enabled long-life rechargeable aqueous zinc ion batteries. Nano Energy 2019, 62, 275–281. [Google Scholar] [CrossRef]
  63. Zeng, X.; Mao, J.; Hao, J.; Liu, J.; Liu, S.; Wang, Z.; Wang, Y.; Zhang, S.; Zheng, T.; Liu, J.; et al. Electrolyte Design for In Situ Construction of Highly Zn2+-Conductive Solid Electrolyte Interphase to Enable High-Performance Aqueous Zn-Ion Batteries under Practical Conditions. Adv. Mater. 2021, 33, 2007416. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, N.; Cheng, F.; Liu, Y.; Zhao, Q.; Lei, K.; Chen, C.; Liu, X.; Chen, J. Cation-Deficient Spinel ZnMn2O4 Cathode in Zn(CF3SO3)2 Electrolyte for Rechargeable Aqueous Zn-Ion Battery. J. Am. Chem. Soc. 2016, 138, 12894–12901. [Google Scholar] [CrossRef]
  65. Zhang, L.; Liu, Z.; Wang, G.; Feng, J.; Ma, Q. Developing high voltage Zn(TFSI)2/Pyr14TFSI/AN hybrid electrolyte for a carbon-based Zn-ion hybrid capacitor. Nanoscale 2021, 13, 17068–17076. [Google Scholar] [CrossRef]
  66. Shah, D.; Mjalli, F.S. Effect of water on the thermo-physical properties of Reline: An experimental and molecular simulation based approach. Phys. Chem. Chem. Phys. 2014, 16, 23900–23907. [Google Scholar] [CrossRef]
  67. Blanc, L.E.; Kundu, D.; Nazar, L.F. Scientific Challenges for the Implementation of Zn-Ion Batteries. Joule 2020, 4, 771–799. [Google Scholar] [CrossRef]
  68. Owusu, K.A.; Pan, X.; Yu, R.; Qu, L.; Liu, Z.; Wang, Z.; Tahir, M.; Haider, W.A.; Zhou, L.; Mai, L. Introducing Na2SO4 in aqueous ZnSO4 electrolyte realizes superior electrochemical performance in zinc-ion hybrid capacitor. Mater. Today Energy 2020, 18, 100529. [Google Scholar] [CrossRef]
  69. Wang, L.; Zhang, Y.; Hu, H.; Shi, H.-Y.; Song, Y.; Guo, D.; Liu, X.-X.; Sun, X. A Zn(ClO4)2 Electrolyte Enabling Long-Life Zinc Metal Electrodes for Rechargeable Aqueous Zinc Batteries. ACS Appl. Mater. Interfaces 2019, 11, 42000–42005. [Google Scholar] [CrossRef]
  70. Li, W.; Wang, K.; Zhou, M.; Zhan, H.; Cheng, S.; Jiang, K. Advanced Low-Cost, High-Voltage, Long-Life Aqueous Hybrid Sodium/Zinc Batteries Enabled by a Dendrite-Free Zinc Anode and Concentrated Electrolyte. ACS Appl. Mater. Interfaces 2018, 10, 22059–22066. [Google Scholar] [CrossRef]
  71. Bayaguud, A.; Luo, X.; Fu, Y.; Zhu, C. Cationic Surfactant-Type Electrolyte Additive Enables Three-Dimensional Dendrite-Free Zinc Anode for Stable Zinc-Ion Batteries. ACS Energy Lett. 2020, 5, 3012–3020. [Google Scholar] [CrossRef]
  72. Chang, N.; Li, T.; Li, R.; Wang, S.; Yin, Y.; Zhang, H.; Li, X. An aqueous hybrid electrolyte for low-temperature zinc-based energy storage devices. Energy Environ. Sci. 2020, 13, 3527–3535. [Google Scholar] [CrossRef]
  73. Kumar, R.M.; Baskar, P.; Balamurugan, K.; Das, S.; Subramanian, V. On the Perturbation of the H-Bonding Interaction in Ethylene Glycol Clusters upon Hydration. J. Phys. Chem. A 2012, 116, 4239–4247. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, J.; Zhang, P.; Ma, K.; Han, F.; Chen, G.; Wei, X. Hydrogen bonding interactions between ethylene glycol and water: Density, excess molar volume, and spectral study. Sci. China Ser. B Chem. 2008, 51, 420–426. [Google Scholar] [CrossRef]
  75. Qin, R.; Wang, Y.; Zhang, M.; Wang, Y.; Ding, S.; Song, A.; Yi, H.; Yang, L.; Song, Y.; Cui, Y.; et al. Tuning Zn2+ coordination environment to suppress dendrite formation for high-performance Zn-ion batteries. Nano Energy 2021, 80, 105478. [Google Scholar] [CrossRef]
  76. Zhao, Z.; Zhao, J.; Hu, Z.; Li, J.; Li, J.; Zhang, Y.; Wang, C.; Cui, G. Long-life and deeply rechargeable aqueous Zn anodes enabled by a multifunctional brightener-inspired interphase. Energy Environ. Sci. 2019, 12, 1938–1949. [Google Scholar] [CrossRef]
  77. Chi, S.-S.; Wang, Q.; Han, B.; Luo, C.; Jiang, Y.; Wang, J.; Wang, C.; Yu, Y.; Deng, Y. Lithiophilic Zn Sites in Porous CuZn Alloy Induced Uniform Li Nucleation and Dendrite-free Li Metal Deposition. Nano Lett. 2020, 20, 2724–2732. [Google Scholar] [CrossRef]
  78. Cao, L.; Li, D.; Hu, E.; Xu, J.; Deng, T.; Ma, L.; Wang, Y.; Yang, X.-Q.; Wang, C. Solvation Structure Design for Aqueous Zn Metal Batteries. J. Am. Chem. Soc. 2020, 142, 21404–21409. [Google Scholar] [CrossRef]
  79. Kamieńska-Piotrowicz, E. Solvation of cobalt(II) and perchlorate ions in binary mixtures of donor solvents. J. Chem. Soc. Faraday Trans. 1995, 91, 71–75. [Google Scholar] [CrossRef]
  80. Senanayake, G.; Muir, D.M. Competitive solvation and complexation of Cu(I), Cu(II), Pb(II), Zn(II), and Ag(I) in aqueous ethanol, acetonitrile, and dimethylsulfoxide solutions containing chloride ion with applications to hydrometallurgy. Metall. Trans. B 1990, 21, 439–448. [Google Scholar] [CrossRef]
  81. Hao, J.; Yuan, L.; Ye, C.; Chao, D.; Davey, K.; Guo, Z.; Qiao, S.-Z. Boosting Zinc Electrode Reversibility in Aqueous Electrolytes by Using Low-Cost Antisolvents. Angew. Chem. Int. Ed. 2021, 60, 7366–7375. [Google Scholar] [CrossRef] [PubMed]
  82. Mohsen-Nia, M.; Amiri, H.; Jazi, B. Dielectric Constants of Water, Methanol, Ethanol, Butanol and Acetone: Measurement and Computational Study. J. Solution Chem. 2010, 39, 701–708. [Google Scholar] [CrossRef]
  83. Aguayo, A.T.; Gayubo, A.G.; Vivanco, R.; Olazar, M.; Bilbao, J. Role of acidity and microporous structure in alternative catalysts for the transformation of methanol into olefins. Appl. Catal. A 2005, 283, 197–207. [Google Scholar] [CrossRef]
  84. Li, T.C.; Lim, Y.; Li, X.L.; Luo, S.; Lin, C.; Fang, D.; Xia, S.; Wang, Y.; Yang, H.Y. A Universal Additive Strategy to Reshape Electrolyte Solvation Structure toward Reversible Zn Storage. Adv. Energy Mater. 2022, 12, 2103231. [Google Scholar] [CrossRef]
  85. Zhou, X.; Lu, Y.; Zhang, Q.; Miao, L.; Zhang, K.; Yan, Z.; Li, F.; Chen, J. Exploring the Interfacial Chemistry between Zinc Anodes and Aqueous Electrolytes via an In Situ Visualized Characterization System. ACS Appl. Mater. Interfaces 2020, 12, 55476–55482. [Google Scholar] [CrossRef] [PubMed]
  86. Feng, D.; Cao, F.; Hou, L.; Li, T.; Jiao, Y.; Wu, P. Immunizing Aqueous Zn Batteries against Dendrite Formation and Side Reactions at Various Temperatures via Electrolyte Additives. Small 2021, 17, 2103195. [Google Scholar] [CrossRef]
  87. Yan, M.; Xu, C.; Sun, Y.; Pan, H.; Li, H. Manipulating Zn anode reactions through salt anion involving hydrogen bonding network in aqueous electrolytes with PEO additive. Nano Energy 2021, 82, 105739. [Google Scholar] [CrossRef]
  88. Wang, A.; Zhou, W.; Huang, A.; Chen, M.; Tian, Q.; Chen, J. Developing improved electrolytes for aqueous zinc-ion batteries to achieve excellent cyclability and antifreezing ability. J. Colloid Interface Sci. 2021, 586, 362–370. [Google Scholar] [CrossRef]
  89. Wei, T.; Peng, Y.; Mo, L.E.; Chen, S.; Ghadari, R.; Li, Z.; Hu, L. Modulated bonding interaction in propanediol electrolytes toward stable aqueous zinc-ion batteries. Sci. China Mater. 2022, 65, 1156–1164. [Google Scholar] [CrossRef]
  90. Di, S.; Miao, L.; Wang, Y.; Ma, G.; Wang, Y.; Yuan, W.; Qiu, K.; Nie, X.; Zhang, N. Dual-anion-coordinated solvation sheath for stable aqueous zinc batteries. J. Power Sources 2022, 535, 231452. [Google Scholar] [CrossRef]
  91. Zhang, Q.; Ma, Y.; Lu, Y.; Ni, Y.; Lin, L.; Hao, Z.; Yan, Z.; Zhao, Q.; Chen, J. Halogenated Zn2+ Solvation Structure for Reversible Zn Metal Batteries. J. Am. Chem. Soc. 2022, 144, 18435–18443. [Google Scholar] [CrossRef] [PubMed]
  92. Deng, W.; Xu, Z.; Wang, X. High-donor electrolyte additive enabling stable aqueous zinc-ion batteries. Energy Storage Mater. 2022, 52, 52–60. [Google Scholar] [CrossRef]
  93. Liu, S.; Mao, J.; Pang, W.K.; Vongsvivut, J.; Zeng, X.; Thomsen, L.; Wang, Y.; Liu, J.; Li, D.; Guo, Z. Tuning the Electrolyte Solvation Structure to Suppress Cathode Dissolution, Water Reactivity, and Zn Dendrite Growth in Zinc-Ion Batteries. Adv. Funct. Mater. 2021, 31, 2104281. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic illustration of the zinc ion deposition process. Adapted with permission from Ref. [47]. Copyright 2022 Energy & Environmental Science. Comparison of the ionic radius, hydrated radius, and hydration enthalpy of (b) metal cation charge carriers and (c) non-metal charge carriers. Adapted with permission from Ref. [46]. Copyright 2022 Energy & Environmental Science. (d) Pourbaix diagram of the Zn metal in aqueous solutions. Adapted with permission from Ref. [45]. Copyright 2020 American Chemical Society.
Figure 1. (a) Schematic illustration of the zinc ion deposition process. Adapted with permission from Ref. [47]. Copyright 2022 Energy & Environmental Science. Comparison of the ionic radius, hydrated radius, and hydration enthalpy of (b) metal cation charge carriers and (c) non-metal charge carriers. Adapted with permission from Ref. [46]. Copyright 2022 Energy & Environmental Science. (d) Pourbaix diagram of the Zn metal in aqueous solutions. Adapted with permission from Ref. [45]. Copyright 2020 American Chemical Society.
Batteries 09 00073 g001
Figure 2. Schematic illustration of solvation structure regulation strategies.
Figure 2. Schematic illustration of solvation structure regulation strategies.
Batteries 09 00073 g002
Figure 3. (a) Raman spectra for different concentrations of zinc-salt electrolytes. Solution structure during Zn plating on the Cu substrate using (b) 2 M ZnSO4 and (c) 4.2 M ZnSO4. (d) Zn plating/stripping on/from the bare Cu current collector in 4.2 M ZnSO4 + 0.1 M MnSO4 at 0.2 mA cm−2. Adapted with permission from Ref. [54]. Copyright 2020 American Chemical Society. (e) Cycling stability of zinc anodes and (f) cycling stability of full cell in 3 M ZnSO4 electrolyte at various current densities. Adapted with permission from Ref. [56]. Copyright 2020 American Chemical Society. (g) Stripping capacity and CE of bare Cu current collector in 4.2 M ZnSO4 + 0.1 M MnSO4 at 0.2 mA cm−2. (h) Long-cycle capacity and CE of Zn||MnO2 at a 938 mA g−1 current density. Adapted with permission from Ref. [54]. Copyright 2020 American Chemical Society.
Figure 3. (a) Raman spectra for different concentrations of zinc-salt electrolytes. Solution structure during Zn plating on the Cu substrate using (b) 2 M ZnSO4 and (c) 4.2 M ZnSO4. (d) Zn plating/stripping on/from the bare Cu current collector in 4.2 M ZnSO4 + 0.1 M MnSO4 at 0.2 mA cm−2. Adapted with permission from Ref. [54]. Copyright 2020 American Chemical Society. (e) Cycling stability of zinc anodes and (f) cycling stability of full cell in 3 M ZnSO4 electrolyte at various current densities. Adapted with permission from Ref. [56]. Copyright 2020 American Chemical Society. (g) Stripping capacity and CE of bare Cu current collector in 4.2 M ZnSO4 + 0.1 M MnSO4 at 0.2 mA cm−2. (h) Long-cycle capacity and CE of Zn||MnO2 at a 938 mA g−1 current density. Adapted with permission from Ref. [54]. Copyright 2020 American Chemical Society.
Batteries 09 00073 g003
Figure 4. Effect of LiTFSI concentration on the structure of cationic solvation sheaths. (a) The pH values of the electrolytes with varying LiTFSI concentrations. (b) FTIR spectra with varying LiTFSI concentration between 3800 and 3100 cm−1. (c) The change with salt concentration of chemical shifts for 17O nuclei in solvent (water). Adapted with permission from Ref. [15]. Copyright 2018 Nature Materials. (d) Schematic of Zn2+-solvation structure. Cycling performance of the Zn||Cu half-cell at a current density of 0.2 mA cm−2, 1 mAh cm−2. (e) Stripping capacity and coulombic efficiency. (f) Galvanostatic plating/stripping of zinc, and voltage profile at the (g) 5th cycle and (h) 75th cycle. Snapshots of in operando OM in the Zn||Cu half-cell at a current density of 1 mA cm−2, 1 mAh cm−2 using (i) LCE and (j) HCE. Adapted with permission from Ref. [57]. Copyright 2022 American Chemical Society.
Figure 4. Effect of LiTFSI concentration on the structure of cationic solvation sheaths. (a) The pH values of the electrolytes with varying LiTFSI concentrations. (b) FTIR spectra with varying LiTFSI concentration between 3800 and 3100 cm−1. (c) The change with salt concentration of chemical shifts for 17O nuclei in solvent (water). Adapted with permission from Ref. [15]. Copyright 2018 Nature Materials. (d) Schematic of Zn2+-solvation structure. Cycling performance of the Zn||Cu half-cell at a current density of 0.2 mA cm−2, 1 mAh cm−2. (e) Stripping capacity and coulombic efficiency. (f) Galvanostatic plating/stripping of zinc, and voltage profile at the (g) 5th cycle and (h) 75th cycle. Snapshots of in operando OM in the Zn||Cu half-cell at a current density of 1 mA cm−2, 1 mAh cm−2 using (i) LCE and (j) HCE. Adapted with permission from Ref. [57]. Copyright 2022 American Chemical Society.
Batteries 09 00073 g004
Figure 5. Cycling performance of the bulk V2O5 cathode in (a) 1 m ZnSO4 and 30 m ZnCl2. (b) at a current density of 50 mA g−1. (c) Long-cycling performance of V2O5 in WISE at 1 A g−1. (d) Schematic illustration of the dynamic structural evolution for V2O5 cathode in different electrolytes. Adapted with permission from Ref. [59]. Copyright 2021 Advanced Science. (e) The plating/stripping and (f) XRD patterns of Zn in 5 and 30 m ZnCl2 in a Zn||Zn symmetric cell at 0.2 mA cm−2 with 10 min. (g) Electrochemical stability window of the ZnCl2 electrolyte at different concentrations. SEM images of the Zn electrodes cycled for 75 times at 1 mA cm−2 with 1 h in (h) 5 m and (i) 30 m ZnCl2. Adapted with permission from Ref. [61]. Copyright 2018 Chemical Communication.
Figure 5. Cycling performance of the bulk V2O5 cathode in (a) 1 m ZnSO4 and 30 m ZnCl2. (b) at a current density of 50 mA g−1. (c) Long-cycling performance of V2O5 in WISE at 1 A g−1. (d) Schematic illustration of the dynamic structural evolution for V2O5 cathode in different electrolytes. Adapted with permission from Ref. [59]. Copyright 2021 Advanced Science. (e) The plating/stripping and (f) XRD patterns of Zn in 5 and 30 m ZnCl2 in a Zn||Zn symmetric cell at 0.2 mA cm−2 with 10 min. (g) Electrochemical stability window of the ZnCl2 electrolyte at different concentrations. SEM images of the Zn electrodes cycled for 75 times at 1 mA cm−2 with 1 h in (h) 5 m and (i) 30 m ZnCl2. Adapted with permission from Ref. [61]. Copyright 2018 Chemical Communication.
Batteries 09 00073 g005
Figure 6. (a) Schematics of the Zn2+ ion diffusion and reduction processes on electrodes in ZnSO4 electrolyte (upper part) and ZnSO4 electrolyte with TBA2SO4 (lower part). (b) Model of DFT calculations, showing a hydrated Zn2+ ion passing through the TBA+ cation adsorption layer on a Zn surface. Adapted with permission from Ref. [71]. Copyright 2020 American Chemical Society. (c) Long-term cycling performance of Zn-MnO2 battery with Et2O additive at 5 A g−1. (d) Cycling performances of Zn||MnO2 battery with Et2O additive at 0.3 A g−1 and 1 A g−1. (e) Schematics of morphology evolution for Zn anodes with and without Et2O additive. Adapted with permission from Ref. [62]. Copyright 2019 Nano Energy. (f) Raman spectra of the SEI-Zn electrode and bare Zn. (g) Schematic illustration of the Zn surface evolution and (h) XRD patterns of Zn electrode before and after 20 cycles in the Zn(H2PO4)2 + Zn(CF3SO3)2. Adapted with permission from Ref. [63]. Copyright 2021 Advanced Materials.
Figure 6. (a) Schematics of the Zn2+ ion diffusion and reduction processes on electrodes in ZnSO4 electrolyte (upper part) and ZnSO4 electrolyte with TBA2SO4 (lower part). (b) Model of DFT calculations, showing a hydrated Zn2+ ion passing through the TBA+ cation adsorption layer on a Zn surface. Adapted with permission from Ref. [71]. Copyright 2020 American Chemical Society. (c) Long-term cycling performance of Zn-MnO2 battery with Et2O additive at 5 A g−1. (d) Cycling performances of Zn||MnO2 battery with Et2O additive at 0.3 A g−1 and 1 A g−1. (e) Schematics of morphology evolution for Zn anodes with and without Et2O additive. Adapted with permission from Ref. [62]. Copyright 2019 Nano Energy. (f) Raman spectra of the SEI-Zn electrode and bare Zn. (g) Schematic illustration of the Zn surface evolution and (h) XRD patterns of Zn electrode before and after 20 cycles in the Zn(H2PO4)2 + Zn(CF3SO3)2. Adapted with permission from Ref. [63]. Copyright 2021 Advanced Materials.
Batteries 09 00073 g006
Figure 7. (a) Schematic illustration of a possible mechanism of how Zn2+-EG solvation interaction impacts the chemistry of the hybrid electrolyte. (b) Raman spectra of a series of hybrid electrolytes. (c) FTIR spectrum of the hybrid electrolytes. Adapted with permission from Ref. [72]. Copyright 2020 Energy & Environmental Science. (d) HER curves in different electrolytes. (e) Schematic illustration of coordination state of Zn2+ in H2O/EG hybrid electrolytes, and the effect of EG on the cycling stability of Zn||Zn symmetric cells. (f) Coordination complexes of Zn[(H2O)6]2+ and Zn[(EG)3]2+. Adapted with permission from Ref. [75]. Copyright 2021 Nano Energy. (g) Scheme of Zn2+ solvation structure and zinc surface passivation in H2O (left) and H2O-DMSO (right) solvents. (h) XRD patterns of Zn anodes after plating/stripping cycles in ZnCl2–H2O–DMSO and ZnCl2–H2O electrolytes. Adapted with permission from Ref. [78]. Copyright 2020 American Chemical Society. (i) Schematic of changes in the Zn2+ solvent sheath with methanol addition. (j) Contact angle measurement on Zn electrode. (k) LSV response curves for different electrolytes at 0.1 mV s−1. (l) Cycling stability of Zn||PANI coin cells at 5 A g−1 with different electrolytes. Adapted with permission from Ref. [81]. Copyright 2021 Angewandte Chemie International Edition.
Figure 7. (a) Schematic illustration of a possible mechanism of how Zn2+-EG solvation interaction impacts the chemistry of the hybrid electrolyte. (b) Raman spectra of a series of hybrid electrolytes. (c) FTIR spectrum of the hybrid electrolytes. Adapted with permission from Ref. [72]. Copyright 2020 Energy & Environmental Science. (d) HER curves in different electrolytes. (e) Schematic illustration of coordination state of Zn2+ in H2O/EG hybrid electrolytes, and the effect of EG on the cycling stability of Zn||Zn symmetric cells. (f) Coordination complexes of Zn[(H2O)6]2+ and Zn[(EG)3]2+. Adapted with permission from Ref. [75]. Copyright 2021 Nano Energy. (g) Scheme of Zn2+ solvation structure and zinc surface passivation in H2O (left) and H2O-DMSO (right) solvents. (h) XRD patterns of Zn anodes after plating/stripping cycles in ZnCl2–H2O–DMSO and ZnCl2–H2O electrolytes. Adapted with permission from Ref. [78]. Copyright 2020 American Chemical Society. (i) Schematic of changes in the Zn2+ solvent sheath with methanol addition. (j) Contact angle measurement on Zn electrode. (k) LSV response curves for different electrolytes at 0.1 mV s−1. (l) Cycling stability of Zn||PANI coin cells at 5 A g−1 with different electrolytes. Adapted with permission from Ref. [81]. Copyright 2021 Angewandte Chemie International Edition.
Batteries 09 00073 g007
Figure 8. (a) The binding energy of Zn2+-H2O, Zn2+- OTf, and Zn2+-Y4- complexes. (b) Schematic of dual-anion electrolyte and (c) base-line electrolyte. (d) and (e) SEM images after 50 cycles in different electrolytes. (f) LSV of different electrolytes. (g) Galvanostatic cycling performances of symmetric Zn||Zn cells in different electrolytes at 1.0 mA cm−2 with 1.0 mAh cm−2. Adapted with permission from Ref. [90]. Copyright 2022 Journals of Power Sources. (h) Coulombic efficiencies of Zn2+ plating/stripping in asymmetric Zn||Cu cells with BE and BE + 100 mM at 1.0 mA cm−2 with 1 mAh cm−2. (i) The NH4+ electrostatic shielding layer inhibits the dendritic growth. SEM of Zn deposition in (j) 1 M ZnAc2 and (k) ISE. Adapted with permission from Ref. [91]. Copyright 2022 American Chemical Society.
Figure 8. (a) The binding energy of Zn2+-H2O, Zn2+- OTf, and Zn2+-Y4- complexes. (b) Schematic of dual-anion electrolyte and (c) base-line electrolyte. (d) and (e) SEM images after 50 cycles in different electrolytes. (f) LSV of different electrolytes. (g) Galvanostatic cycling performances of symmetric Zn||Zn cells in different electrolytes at 1.0 mA cm−2 with 1.0 mAh cm−2. Adapted with permission from Ref. [90]. Copyright 2022 Journals of Power Sources. (h) Coulombic efficiencies of Zn2+ plating/stripping in asymmetric Zn||Cu cells with BE and BE + 100 mM at 1.0 mA cm−2 with 1 mAh cm−2. (i) The NH4+ electrostatic shielding layer inhibits the dendritic growth. SEM of Zn deposition in (j) 1 M ZnAc2 and (k) ISE. Adapted with permission from Ref. [91]. Copyright 2022 American Chemical Society.
Batteries 09 00073 g008
Table 1. Recent publications focused on high−concentration electrolytes in ZIBs.
Table 1. Recent publications focused on high−concentration electrolytes in ZIBs.
ElectrolyteCurrent DensityCELifespanRef.
3 M ZnSO420 mA cm−2, 1 mAh cm−2100%800 h[56]
3 M ZnSO4 + 2 M LiCl 0.2 mA cm−2, 2 mAh cm−2__170 h[68]
4.2 M ZnSO4 + 0.1 M MnSO40.5 mA cm−2, 1 mAh cm−299.21%1000 h[54]
3 M Zn(CF3SO3)20.1 mA cm−2, 0.1 mAh cm−2100%800 h[64]
1 m Zn(TFSI)2 + 20 m LiTFSI 0.2 mA cm−2, 0.033 mAh cm−2100%170 h[15]
30 m ZnCl20.2 mA cm−2,0.035 mAh cm−295.4%600 h[61]
30 m ZnCl2 + 5 m LiCl 2 mA cm−2, 4 mAh cm−299.7%4000 h[60]
2.4 m Zn(ClO4)21 mA cm−2, 1 mAh cm−299%3000 h[69]
8 M NaClO4 + 0.4 M Zn(CF3SO3)21 mA cm−2, 1 mAh cm−2__200 h[70]
Table 2. Recent publications focused on organic additives in ZIBs.
Table 2. Recent publications focused on organic additives in ZIBs.
Electrolyte AddictiveCurrent DensityCELifespanRef.
68 vol.% Ethylene glycol 0.5 mA cm−2, 0.5 mAh cm−2__2668 h[75]
50 vol.% Methanol 1 mA cm−2, 0.5 mAh cm−299.7%__[81]
Polyethylene glycol 1 mA cm−2, 1 mAh cm−299.6%650 h[85]
Dimethyl sulfoxide 1 mA cm−2, 1 mAh cm−299.73%2100 h[86]
PEO 1 mA cm−2, 1 mAh cm−298.7700 h[87]
Diethyl ether + Ethylene glycol 20 mA cm−2, 1 mAh cm−298%700 h[88]
Propanediol 0.2 mA cm−2, 0.2 mAh cm−298.9%1000 h[89]
Diethyl ether 20 mA cm−2, 1 mAh cm−2__250 h[62]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, B.; Xu, H.; Hao, J.; Du, J.; Wu, C.; Ma, Z.; Qin, W. Mini-Review on the Regulation of Electrolyte Solvation Structure for Aqueous Zinc Ion Batteries. Batteries 2023, 9, 73. https://doi.org/10.3390/batteries9020073

AMA Style

Wang B, Xu H, Hao J, Du J, Wu C, Ma Z, Qin W. Mini-Review on the Regulation of Electrolyte Solvation Structure for Aqueous Zinc Ion Batteries. Batteries. 2023; 9(2):73. https://doi.org/10.3390/batteries9020073

Chicago/Turabian Style

Wang, Bixia, Hui Xu, Jiayi Hao, Jinchao Du, Chun Wu, Zhen Ma, and Wei Qin. 2023. "Mini-Review on the Regulation of Electrolyte Solvation Structure for Aqueous Zinc Ion Batteries" Batteries 9, no. 2: 73. https://doi.org/10.3390/batteries9020073

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop