Jiayi
Liu
abc,
Zhongrong
Shen
abc and
Can-Zhong
Lu
*abc
aState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, P. R. China. E-mail: czlu@fjirsm.ac.cn
bUniversity of the Chinese Academy of Sciences, Beijing 100049, P. R. China
cXiamen Key Laboratory of Rare Earth Photoelectric Functional Materials, Xiamen Institute of Rare Earth Materials, Haixi Institutes, Chinese Academy of Sciences, Xiamen 361021, P. R. China
First published on 1st January 2024
Aqueous zinc-ion batteries (AZIBs) have garnered significant attention for large-scale applications due to their low cost, safety, and eco-friendliness. Among all cathode materials, Prussian blue and its analogues (PBAs) have garnered significant attention and been subjected to extensive research due to their diverse property-effectiveness, simple synthesis process, reversible Zn2+ insertion/removal capability, and other notable characteristics. However, numerous studies have revealed that the utilization of PBAs as a cathode material in AZIBs still presents certain challenges, including limited energy density and prolonged structural instability during cycling. In light of the aforementioned opportunities and challenges for PBAs, we conducted a comprehensive literature review encompassing synthesis, analysis of energy storage mechanisms, enhancement of electrolyte–electrode interfaces, and battery applications. The objective of this review is to encourage active engagement from researchers interested in AZIBs and PBA cathode materials, thereby fostering research and industrialization efforts that collectively contribute towards the advancement of safe and environmentally friendly energy storage.
Metal ions | Li+ | Na+ | K+ | Mg2+ | Zn2+ | Al3+ |
---|---|---|---|---|---|---|
Ionic radius [Å] | 0.76 | 1.02 | 1.38 | 0.72 | 0.740 | 0.535 |
Theoretical specific capacity [mA g g−1] | 3862 | 1166 | 685 | 2205 | 820 | 2980 |
Volumetric energy density [mA h cm−3] | 2061 | 1129 | 610 | 3834 | 5851 | 8064 |
Standard potential (V vs. SHE) | −3.04 | −2.71 | −2.93 | −2.37 | −0.76 | −1.66 |
Crustal abundance [ppm] | 18 | 23000 | 21000 | 23000 | 79 | 82000 |
Metal cost [$ per ton] | 165000 | 200 | 1000 | 4600 | 2570 | 2390 |
The conventional aqueous zinc ion battery consists of a zinc metal anode, a near-neutral (or mildly acidic) aqueous electrolyte, and a cathode that facilitates the reversible intercalation (deintercalation) of Zn2+ ions. Among these, the design and modification of cathode materials is an essential stage in the development of AZIBs. Currently, Zn2+ can be reversibly inserted or removed from manganese-based oxides, vanadium-based oxides, Prussian blue analogues (PBAs), organic compounds, polyanionic compounds, and two-dimensional layered materials (Mxene, sulfides, and layered selenides), among others.11,15–23 We illustrate the performance characteristics of various materials with radar diagrams (Fig. 1c). Manganese-based materials often face challenges of rapid capacity decay due to structural instability caused by manganese dissolution during cycling, parasitic side reactions, and structural transitions.24–27 The stability and capacity of compounds based on the element V are enhanced compared to those of Mn-based cathodes due to their superior capacitance contribution and more robust crystal structure.
However, the average operating voltage of most V-based compounds compared to Zn/Zn2+ is approximately 0.75 V, which prevents the attainment of high energy density.28–31 Prussian blue and its analogues (PBAs) stand out as a class of high-voltage resistant materials. Their operating voltage range can reach 1.2 V, in addition to their simple synthesis process, low cost, and environmentally friendly characteristics.32,33
Among the 20000 species of MOFs, PBAs belong to a distinct subclass characterized by unique structural and electrochemical properties.34 The applications of PBAs span across various disciplines, encompassing hydrogen storage, biosensing, cancer therapy, and sea water desalination.35 In recent years, PBAs have been employed as precursors in the synthesis of a variety of functional materials. In the framework of PBA crystals, the substantial gap (∼4.6 Å) can effectively accommodate a wide range of alkali metal ions as well as multivalent metal ions. The success of monovalent ions in sodium ion and potassium ion batteries has prompted scientists to investigate the feasibility of incorporating bivalent ions into such materials (Fig. 1b).36–39 The original Prussian blue is ferric hexacyanoferrate (Fe4[Fe(CN)6]3). Early research focused primarily on the available Zn2+ storage host, where Fe atoms connected to N atoms can be replaced and interstitially modified with Cr, Zn, Mn, Co, Ni, Cu, and other elements to produce a series of metallic hexacyanoferrates (MHCFs) with a similar composition and crystal structure. These salts are known as Prussian blue analogues (PBAs), which have become an important electrode material in ZIBs due to their inexpensive raw materials, straightforward synthesis process, and working voltage of up to 1.5–1.8 V (vs. Zn2+/Zn) in aqueous electrolyte.
The investigation of PB and PBAs has led to an increasing number of papers reporting their utilization as cathode materials for zinc ion batteries in recent years. In this review, the author focuses on the current status and challenges associated with the use of PBAs in ZIB cathode materials (Fig. 2), by focusing on the structure and energy storage mechanism of Prussian blue and its analogues, as well as by referencing current academic research on the electrochemical energy storage of Prussian blue analogues. Here, various varieties of PBAs applied to ZIB cathode materials are described, along with their respective characteristics and current challenges.
Fig. 2 A summary of this review (reprinted with permission from ref. 88. Copyright 2022, Elsevier; reprinted with permission from ref. 50. Copyright 2021, John Wiley and Sons). |
Fig. 3 (a) Ideal Prussian blue analogue crystal structure. (b) Prussian blue ideal framework and its coordination form. (c) Prussian blue frame containing structural water. (d) Schematic diagram of a zinc ion battery based on a ZnHCF positive electrode. (Reprinted with permission from ref. 49. Copyright 2014, John Wiley and Sons). (e) Schematic of key factors in suppressing metal-ion dissolution of A-PBA in the concentrated electrolyte. (Reprinted with permission from ref. 54. Copyright 2022, American Chemical Society). |
K0.86Ni[Fe(CN)6]0.954(H2O)0.766 + xZn2+ + 2xe− ↔ K0.86ZnxNi[Fe(CN))6]0.954(H2O)0.766 |
In 2014, Zhang et al.49 introduced zinc hexacyanoferrate (Zn3[Fe(CN)6]2, ZnHCF) as the positive electrode of an aqueous zinc ion battery, which demonstrates the removal of cations(Fig. 3d). The corresponding positive electrode electrochemical reaction shown in Fig. 3d can be expressed as:
Cathode:
Zn2+ + 2xe− + Zn3[Fe(CN)6]2 ↔ Zn3+x[Fe(CN)6]2 |
The investigation of energy storage mechanisms necessitates careful attention to the structural transformation of PBAs during charge and discharge. Deng investigated the XRD analysis of a KMnHCF electrode under various cycling conditions and observed a monoclinic to cubic phase transition during the charging process, with both phases coexisting within a specific voltage range. Subsequently, the cubic phase persisted until complete charging of the electrode, followed by a reversion to the monoclinic phase during discharge engineering (Fig. 4a).50 Aili et al. observed a positive shift in the peak of KFCHCF (K1.66Fe0.25Co0.75[Fe(CN)6]·0.83H2O) in the X-ray diffraction pattern during the charging process, indicating contraction of the crystal structure after ion extraction, with a subsequent return to its original position upon discharge. This subtle structural change accounts for the high speed and recyclability exhibited by PBA materials (Fig. 4b and c).51
Fig. 4 (a) In situ XRD patterns of a KMnHCF electrode. First two cycles to reveal the structure evolution during Zn2+ (de)insertion. (Reprinted with permission from ref. 50. Copyright 2021, John Wiley and Sons). (b) The ex situ XRD patterns and (c) the enlarged area of the KFCHCF sample at different charged and discharged states. (Reprinted with permission from ref. 51. Copyright 2023, John Wiley and Sons.). (d) and (e) In situ Raman spectroscopic analysis of the energy storage mechanism of VOHCF. (Reprinted with permission from ref. 53. Copyright 2023, John Wiley and Sons). (f) Schematic illustration of reversible Zn2+ intercalation/deintercalation in CoFe(CN)6 frameworks during the electrochemical process. (Reprinted with permission from ref. 60. Copyright 2019, John Wiley and Sons). |
The MB site in PBAs is the representative redox active site in this structure. In addition, the substitution of a different transition metal MA influences the reaction potential and capacity of PBAs. The MA and MB sites are occupied by distinct or identical metal atoms, resulting in distinct combination and reaction potentials.42,52 Fe, Mn, Co, and V are electrochemically active within the stable window of the organic electrolyte and aqueous system in PBAs. When these metal ions occupy the MA site, a double-electron redox reaction can occur. In situ analysis techniques were employed to investigate the PBAs at two REDOX sites. For the VOHCF electrode, the intensity of the VO at 864–921 cm−1 gradually diminishes as the charge progresses, indicating a gradual ion insertion process. Upon discharge, there is a slow increase in the intensity of the VO peak, suggesting reversible ion ejection and embedding within VOHCF (Fig. 4d). In addition, the [Fe(CN)6]4− group gradually shifts to higher wavenumbers at around 2150 cm−1, followed by a reverse migration that corresponds to Fe2+ to Fe3+, thereby highlighting the pivotal role of [Fe(CN)6]4− as an additional REDOX active site (Fig. 4e).53 In addition, many PBAs have a reversible zinc ion intercalation/deintercalation reaction and high charge–discharge performance due to the coupling of multiple redox reactions and an ideal crystal structure.
In the process of a charge–discharge cycle, besides the crystal structure change induced by Zn2+ intercalation/deintercalation, PBAs as cathodes also show the dissolution of metal ions in the bulk.54 The bulk metal ion dissolution process undergoes the following main steps (Fig. 3e):(i) the metal ions in the material bulk diffuse to the surface of the solid electrode; (ii) metal ions are solvated by free solvation (free H2O molecules) at the solid–liquid interface; (iii) diffusion of solvation metal ions to electrolytic liquid phases.55–57 The dissolution of the electrode material in PBAs and its compatibility with the electrolyte will be comprehensively discussed in the section dedicated to electrolytes.
Fig. 5 (a) Schematic illustration of the redox mechanism of HQ-NaFeHCF. (Reprinted with permission from ref. 59. Copyright 2014, Royal Society of Chemistry). (b) Schematic illustration of the nano-impact method for the kinetic study of K+ insertion/de-insertion in single PB particles. The ion channels at the interface of Fe[Fe(CN)6] (c) and KFe[Fe(CN)6] (d). (Reprinted with permission from ref. 63. Copyright 2019, John Wiley and Sons). (e) Concentration of the oxidation states of Fe3+ and Mn3+ upon electrochemical potentials. The percentage of oxidation states in hydrated and anhydrated samples reflects the REDOX sequence. (Reprinted with permission from ref. 69. Copyright 2017, American Chemical Society). |
In 2016, Ojwang et al.61 inserted the K+ portion into the CuHCF structure by reducing Fe(III) to Fe(II) in 0.1 M K2S2O3 aqueous electrolyte.61 The parameters of the CuHCF unit cell decrease as Fe(III) is converted to Fe(II) and as the quantity of K inserted increases. In their later work, which focused on the electrochemical-structural relationship of CuHCF cycling in 1 M ZnSO4, and it was demonstrated that the average cell voltage increased and a two-step plateau was observed during the initial CuHCF cycle.62 Consequently, the initiation (deinsertion) of zinc is associated with the nonlinear modification of CuHCF unit cell parameters during material operation. These results indicate that the reduction of CuHCF unit cells is greater in the presence of zinc ions compared to potassium ions.
Zhou et al.63 employed nano-impact electrochemistry (NIE), a novel technique, to reveal the intrinsic properties of the electrochemical processes that occur in PB nanoparticles. It was found that the KFHCF particle is not solely composed of KFe[Fe(CN)6] but also consists of K2Fe[Fe(CN)6] and Fe[Fe(CN)6]. Consequently, the electrochemical behavior of an individual PB particle enables the measurement of two distinct pairs of REDOX reactions within the same particle, irrespective of variations among different particles in the set. The feedback current signal exhibited by a single particle exhibits variation with respect to the applied potential, as depicted in Fig. 5b. When an electric potential is applied between two REDOX centers, both the oxidation and reduction current signals can be observed; however, the signal difference is related to the aggregation mode of K+.
K2FeII[FeII(CN)6] ⇌ KFeIII[FeII(CN)6] ⇌ FeIII[FeIII(CN)6] |
The findings of this study demonstrate the kinetic influence of K+ on the ion embedding process. Specifically, the presence of K+ obstructs a portion of the ion transport channel, thereby restricting the reduction of KFe[Fe(CN)6] through slower interfacial transfer with K+, which is comparatively slower than the diffusion rate of K+ within the particle. In contrast, the reduction of Fe[Fe(CN)6] is constrained by the K+ within the particles, which is considered to exhibit slower kinetics compared to the transfer of K+ at the electrolyte composite particle interface (Fig. 5c and d).
By means of neutron diffraction and X-ray single crystal diffraction, it is determined that there are two kinds of water molecules with different structural positions in the crystal structure of PBA.65,66 The first type is zeolite water, or interstitial water, which competes with inserted ions for site A, or vacancy in the framework. Fig. 3c depicts the second form of water, which is coordination water coordinated with the MA site ion at the hexacyanoferrate vacancy. The negatively charged oxygen atoms in these water molecules assist in shielding a large number of positive transition metal ion charges at the vacancies. There is a correlation between the quantity of vacancies and coordinated water in the lattice, with vacancy sites providing space for water molecules coordinated with MA ions and possibly providing sites for coordinated water. Doping MA sites in organic systems with nickel or electrochemically inert metals to preserve the structural stability advantages of vacancies and water is an active area of research. However, the interfacial charge transfer of an aqueous electrolyte is significantly superior to that of an organic electrolyte due to the fact that ion diffusion in an organic electrolyte requires solvation/desolvation and ion diffusion to the electrode surface. In an aqueous electrolyte, the activation energy required for only partial shedding of ion solvation spheres is less than that in an organic system; consequently, the kinetics are faster.67
However, the effect of water on ion conduction in a lattice is extremely complex and poorly understood. According to one theory, the movement of ion groups through crystal structure vacancies involves exchange with nearby water molecules and rotation around the surrounding ion hydration sheath. This ion migration assisted by rotation is called a “gear” or “paddle wheel” mechanism, which is used to move ions through channels and vacancies.68 It has been found that the presence of interstitial water in a Prussian blue lattice has an effect on the redox platform of transition metals.69 In particular, the traditional view holds that the redox potential is dependent on the ionization energy of a transition metal in a particular oxidation state. In this study, the standard ionization energy of Fe2+ is less than that of Mn2+, whereas the ligand field stability energy (LFSE) of Fe2+ (LS) is greater than that of Mn2+ (HS), resulting in an increase in the redox potential of Fe2+/3+. If the ionization energy and the competitive effect of LFSE are in equilibrium, the Fe2+ (LS) and Mn2+ (HS) redox potentials will overlap to form a single plateau (Fig. 5e). In a hydrated system, the interstitial water molecules dilute the ligand field in the FeC6 and MnN6 octahedra and disrupt the structure that defines the spin states. When the LFSE effect is unable to compete with the ionization energy, the redox potential gap between Fe2+/3+ and Mn2+/3+ reappears, resulting in a double plateau structure in the electrochemical morphology of the hydrated electrode.
Cappe et al. studied the effect of water on cation mobility in an A2Zn3[Fe(CN)6]2·xH2O skeleton by using impedance spectroscopy data, where A = Na, K, Rb, Cs.70 It was discovered that porous hexacyanometallates have low surface polarity and no partially exposed transition metal coordination centers, which reduces the possibility of H+ and OH− analogue formation. The presence of weakly bonded water molecules in the lattice increases the mobility impedance resistance. This inference can be explained by the fact that increased hydration of the solid leads to the formation of a denser network of water molecules within the porous framework, thereby reducing the number of available pathways for cation migration. When the number of water molecules in a cationic coordination environment increases, the water molecules become more closely associated with the solid, and the barrier to diffusion increases in energy. Therefore, further theoretical and targeted studies, such as nuclear magnetic resonance spectroscopy or isotope labeling studies, can more accurately distinguish the effects of zeolite water and coordination water on ion conduction.
Fig. 6 (a) TEM images of Prussian blue nanocubes prepared at 40 °C using different concentrations of K4Fe(CN)6. (Reprinted with permission from ref. 75. Copyright 2006, American Chemical Society). (b) Schematic diagram of the nucleation and growth process of the ZnHCF particle shape at different drop rates of reactants. (c) SEM images of cubic octahedron ZnHCF, truncated angular octahedron ZnHCF and octahedron ZnHCF obtained at room temperature. (Reprinted with permission from ref. 83. Copyright 2015, Springer Nature). |
Materials | Precursor | Chelating agent | T | Reaction time | Size | Ref. |
---|---|---|---|---|---|---|
K2Zn3[Fe(CN)6]2 | 0.1 M ZnSO4 + 0.05 M K3Fe(CN)6 | — | 60 °C | Several hours | 200 nm | 49 |
K0.08Zn1.53[Fe(CN)6]·0.26H2O | 0.02 M ZnSO4 + 0.02 M K3Fe(CN)6 | — | Room temperature | 1 h | 500–800 nm | 81 |
K1.88Zn2.88[Fe(CN)6]2(H2O)5 | 50 mM ZnSO4·7H2O + 50 mM K4Fe(CN)6·3H2O | — | 70 °C | 30 min | 70–120 nm | 82 |
K0.07Zn[Fe(CN)6]0.69 | 0.1 M ZnSO4 + 0.05 M K3Fe(CN)6 | — | Room temperature | Immediately | 1.7 μm | 83 |
K0.08Zn[Fe(CN)6]0.67 | 0.1 M ZnSO4 + 0.05 M K3Fe(CN)6 | — | Room temperature | 20 min | ∼3.2 μm | 83 |
K0.07Zn[Fe(CN)6]0.68 | 0.1 M ZnSO4 + 0.05 M K3Fe(CN)6 | — | Room temperature | 2 h | ∼2.6 μm | 83 |
Zn3[Fe(CN)6]2 | ZnSO4 + K3Fe(CN)6 | — | 70 °C | 1 h | 300 nm | 84 |
CuHCF | Cu(NO3)2·3H2O + K3Fe(CN)6 | — | Room temperature | 30 min | 100 nm | 85 |
K2x/3Cu[Fe(CN)6]2/3⋅nH2O | Cu(NO3)2·3H2O + K3Fe(CN)6 | — | Room temperature | 30 min | 40–100 nm | 86 |
CuHCF (nano cube) | CuSO4 + K3Fe(CN)6 | — | Room temperature | 30 min | 50–100 nm | 43 |
CuZnHCF | Cu(NO3)2·3H2O + Zn(NO3)2·6H2O + K3Fe(CN)6 | — | Room temperature | 30 min | 2–12 μm | 87 |
NiHCF (nanocubes) | Ni(NO3)2·6H2O + K3Fe(CN)6 | Na3C6H5O7 | Room temperature | 10 min | 100 nm | 88 |
NiFe(CN)6 | Ni(NO3)2 + Na4Fe(CN)6 | — | — | 2 h | 89 | |
K1.51Ni[Fe(CN)6]0.954(H2O)0.766 | Ni(NO3)2·6H2O + K4[Fe(CN)6]·3H2O | — | Room temperature | 30 min | 5–25 nm | 48 |
K1.6Mn[Fe(CN)6]0.94·0.63H2O | MnSO4 + K4Fe(CN)6·3H2O | Potassium citrate | Room temperature | 24 h | 500–600 nm | 50 |
ZnMnFe-PBA | MnSO4·H2O + K3Fe(CN)6 + ZnSO4·7H2O | — | 80 °C | 1 h | 0.5–1 μm | 90 |
K2MnFe(CN)6 | MnCl2·4H2O + K4Fe(CN)6 | 2 g PVP(K30) | 80 °C | 50 min | 2–4 μm | 91 |
KCoFe(CN)6⋅5.5H2O | K3[Fe(CN)6] + Co(CH3COO)2·4H2O | — | Room temperature | 24 h | 500 nm | 60 |
CoMn-PBA HSs | K4[Fe(CN)6]·3H2O + Mn(NO3)2·4H2O + Co(NO3)2·6H2O | — | 90 °C | 4 h | 1.2 μm | 92 |
FeFe(CN)6 | K3[Fe(CN)6] + FeCl3 +30% HCl | — | 100 °C | 30 min | 300 μm | 33 |
VHCF | VOPO4·2H2O + Na4Fe(CN)6⋅10H2O | — | 80 °C | 4 h | 23.6 nm | 93 |
VHCF | VOSO4·xH2O + K3Fe(CN)6 | — | 60 °C | 4 h | 30 nm | 94 |
VO-PBAs | VOSO4 + K3Fe(CN)6 | — | Room temperature | Immediately | 30–50 nm | 95 |
Fig. 7 SEM images of PB particles prepared with different amounts of PVP added (a) 3 g, (b) 5 g, and (c) 10 g. (Reprinted with permission from ref. 106. Copyright 2012, Royal Society of Chemistry). TEM images of chitosan-PB nanoparticles prepared at [Fe2+], [Fe(CN)6]3− = 1 mM, [chitosan] = 3 mg mL−1 (d), 2 mg mL−1 (e), and 1 mg mL−1 (f). (Reprinted with permission from ref. 98. Copyright 2008, Elsevier). (g) Morphology evolution of Prussian blue from a cube to a star-like hexapod with increasing concentrations of HNO3. (Reprinted with permission from ref. 99. Copyright 2012, Korean Chemical Society). Field emission scanning electron microscope (FE-SEM) images of (h) PB-1 mL HCl, (i) PB-2 mL HCl, and (j) PB-3 mL HCl. (Reprinted with permission from ref. 100. Copyright 2017, John Wiley and Sons). |
Lou's Group controlled the amount of PVP and citrate to control the crystal growth kinetics of PBA. Frame-like nanostructures were found in the absence of additives, and nanocubes were formed with the concentration of additives.105 In another attempt, the influence of citrate and PVP concentration on the growth of a PBA crystal was discussed. High concentrations of PVP and citrate produce a surface closed cube with a truncated cube shape, while low concentrations of citrate result in an open cage structure with cavities at corners. Hu et al.106 developed a simple method to control the size and morphology of PB nanoparticles by systematically determining the pH value of the solution and the amount of K3[Fe(CN)6]. They developed three different PB nanoparticles: small (about 20 nm), medium (about 100 nm) and large (about 200 nm) (Fig. 7a–c). Currently, PVP is crucial in determining the final shape, crystallization process, size, and oxidation state of the metal skeleton. The pH of the solution, the concentration of K3[Fe(CN)6], and the amount of PVP control the size and morphology of the final product via a multistep mechanism. PVP is utilized primarily as a capping agent to stabilize the crystal nucleus in the initial stage. When the concentration of PVP is low, direct particle precipitation cannot expand. However, when the amount of PVP is sufficient, the nuclei first stabilize and grow in an orderly manner. The acid concentration regulates K3[Fe(CN)6] and produces trivalent Fe ions, which, along with the ions in [Fe(CN)6]4−, result in the formation of PB nanocubes. Therefore, a small amount of PVP is advantageous for products containing small nanoparticles, while a large amount of PVP is advantageous for products containing medium and large nanoparticles. It can be concluded that chelating agents and surfactants have a controllable effect on the crystallinity and morphology of synthesized PB/PBAs. The chelating agent regulates the release of precursor ions to facilitate crystal growth and the formation of large particles from the released ions. Surfactants inhibit the growth of particle size by acting directly on the surface of nanoparticles.
Fig. 8 (a) Schematic illustration of the synthetic process of NiHCF/RGO. (b) TEM image of NiHCF/RGO. (c) SEM image of NiHCF/RGO. (Reprinted with permission from ref. 88. Copyright 2022, Elsevier). (d) Schematic diagram for the fabrication procedures of the ZnHCF@MnO2 composite, and (e) corresponding TEM morphology. (Reprinted with permission from ref. 107. Copyright 2017, Royal Society of Chemistry). SEM images of (f) CoHCF-Cit/CNT and (g) CoHCF. (h) Galvanostatic discharge/charge profiles. (Reprinted with permission from ref. 108. Copyright 2018, Elsevier). |
In 2013, Zhang et al.107 developed an in situ coprecipitation method to prepare manganese oxide-coated zinc hexocyanoferrate (ZnHCF) nanocubes (ZnHCF@MnO2). Zn2+ ions were adsorbed on the surface of two-dimensional MnO2 nanosheets via electrostatic interaction in an aqueous electrolyte, and then K3Fe(CN)6 was added dropwise during aging to form ZnHCF@MnO2 composites. The intercalated ZnHCF core is surrounded by a pseudocapacitive manganese oxide shell, and the 2D MnO2 nano-sheet serves as a buffer layer (Fig. 8d and e). ZnHCF was used as the central core, and 2D MnO2 nanosheets with a certain pseudocapacity served as the buffer layer. The re-embedded Zn2+ from ZnHCF after charging enters the peripheral MnO2 nanosheets, and Zn2+ from MnO2 remains embedded in ZnHCF during discharge, thereby reducing the diffusion control from the electrolyte to the electrode. In addition, the Zn2+ in the electrolyte contributes additional capacity to MnO2, thereby enhancing the cell's performance.
CNT and PB can be recombined to enhance PB's intrinsic structural vacancies and poor electronic conductivity. It is simple for carbon materials such as CNTs to aggregate and challenging for them to disseminate uniformly in PBA. In addition, the high surface energy of PBA facilitates aggregation during growth, preventing carbon materials from forming complete 3D conducting channels. Qian et al. synthesized a cobalt hexocyanoferrite truncated nanocube (CoHCF-Cit/CNT) with citrate and glycerol to address this problem (Fig. 8f and g).108 CoHCF-Cit/CNT was used as the cathode of a Na+/Zn2+ double ion battery for the first time. CNTs not only show effective internal resistance for CoHCF particles, but also show significant resistance between adjacent CoHCF particles, which significantly improves the electrochemical performance and cycle capacity of CoHCF-Cit/CNT (Fig. 8h). Additionally, research has produced PB-CNT composites that prevent PB particle agglomeration by penetrating single PB particles into carbon nanotubes.109 CNTs significantly enhance the composite's electron transport, resulting in four orders of magnitude greater electron conductivity at room temperature for PB/CNT than for PB alone. Due to the “metallic nature” of CNTs, the conductivity of the PB/CNT composite increased by 47.6% when cooled to −25 °C compared to the conductivity of the material without a CNT composite, indicating its potential in the construction of low-temperature batteries. The obtained materials exhibit outstanding electrochemical properties at low temperatures, and the synthesis strategy is applicable to the development of additional low temperature cathode materials.
All of the aforementioned composites improve the electrochemical performance of PBAs as a positive electrode material to some degree, and although the MnO2 coating introduces a new energy storage mechanism, it still has poor conductivity and excessive internal resistance, which will reduce the battery's lifespan. However, the CNT film coating had no negative impact on the internal resistance, but the capacity increase was insufficient. Moreover, during cycling, irreversible phase transition, transition metal dissolution, and structure collapse occur. Huang et al. coated PANI in ZnHCF by simple in situ polymerization to prepare ZnHCF/PANI positive electrodes.110 As the coating material of ZnHCF, PANI can effectively avoid the problems of high internal resistance and large cushioning stress in large deformation. PANI coating provides a new voltage platform of about 1.2 V for ZnHCF at 1.8 V. It can reach a high capacity of 150 mA h g−1 at 100 mA g−1, and the capacity rate is 75% after 350 cycles.
Particle size also has an effect on water adsorption, and smaller crystals may lead to stronger water adsorption.116 The sodium-rich Na1+xFeFe(CN)6 produced by Li et al. in a single step can increase the number of sodium ion skeletons with effective vacancies and coordination water.117 The amount of water in the MNHCF structure can be obtained by dewatering at high temperature under vacuum conditions.118,119 The vacancies and defect-rich PBA produced by dehydration are incapable of reaching their theoretical capacity due to the absence of Fe[(CN)6]3−/4− as a structural support during cation intercalation and disintercalation, resulting in a crystal structure that is fragile. By suppressing lattice defects, it is possible to increase the amount of water in the PB framework; the key to this strategy is to improve crystallinity.115,120,121 Reducing the crystallization rate is the key to improving PBA crystallization, and numerous researchers have proposed various solutions: (1) the chelator/surfactant-assisted coprecipitation method, where the crystallization rate can be slowed by using sodium citrate as a chelator to increase the crystallinity of PBA.114,122 (2) Decomposition of hexocyanate metal salts in acid: Guo and his colleagues59 fabricated high-quality PB nanocubic crystalline cubic crystal Na0.61Fe[Fe(CN)6]0.94 (HQ-NaFe) by a simple synthesis procedure using Na4Fe(CN)6 as the sole iron source. Due to the low water content of HQ-NaFe zeolite crystal growth, there is a small amount of [Fe(CN)6] vacancies in the crystal skeleton, which further enhances the ion storage capacity, enriches the transport path, and effectively preserves the crystal structure throughout the cycle.
The addition of a chelating agent has an significant effect on inhibiting the formation of a good crystal structure by adding Mn2+ in a K/Na-citrate solution (citric acid k/Na) chelating reactant to slow down the nucleation of KMHCF to achieve uniform and nearly chemical crystal growth.123 Compared with citrate as the chelating agent, EDTA-2K has stronger complexing ability to Mn2+ (Kstable[Mn(EDTA)]2− = 1013.8 » Kstable[Mn(citrate)]− = 103.67). The nucleation and growth of K2Mn[Fe(CN)6] can be greatly inhibited, and the defects and moisture of the synthesized sample (KMF-EDTA) are significantly reduced.124 In addition to the use of a chelating agent, the acid-assisted hydrothermal method can slow down the reaction rate of FeFe(CN)6 nanocrystals with low defects, which may be due to the acid-assisted dynamic balance between corrosion and growth. The precursor K3Fe(CN)6 slowly decomposes into Fe3+/Fe2+ in acid solution, and then reacts with residual [Fe(CN)6]3− to form FeFe(CN)6 nanocrystals with low defects.59,115 Therefore, some strategies to reduce water content in PBA can be summarized as follows: (a) coarsening PBA particles to reduce surface adsorption; (b) dehydration of PBA samples at higher temperatures under high vacuum; (c) introducing more basic ions into the skeleton to reduce the adsorption sites of zeolite water; (d) reduce the coordination sites of coordinating water by reducing the number of MB(CN)6 vacancies.
Fig. 10 (a) Unit cell structure of NZH. (Reprinted with permission from ref. 36. Copyright 2012, Royal Society of Chemistry). (b) Cycling performance of ZnHCF and MZHCF at 250 mA g−1. (Reprinted with permission from ref. 81. Copyright 2021, American Chemical Society). XPS spectra of the original (c) and Zn-inserted CuHCF electrode (d). (Reprinted with permission from ref. 43. Copyright 2014, Elsevier). (e) The schematic diagram of the phase transformation for the KMnHCF electrode. (Reprinted with permission from ref. 50. Copyright 2021, John Wiley and Sons). (f) Synthesis process of CoMn-PBA HSs. (Reprinted with permission from ref. 92. Copyright 2021, John Wiley and Sons). (g) Ex situ XPS spectra of Fe 2p of FeHCF cathodes in the discharge state. (Reprinted with permission from ref. 138. Copyright 2019, John Wiley and Sons). |
In 2014, Zhang et al.49 reported for the first time that ZnHCF was used as a host material for divalent ion insertion on the basis of previous research studies. The experiments show the first use of ZnHCF as a cathode material for an aqueous zinc ion battery. When it is assembled with a zinc anode, the average operating voltage of ZnHCF can reach 1.7 V, which is a record operating voltage of aqueous zinc ion batteries. Based on the total active electrode materials, it also provides a specific energy density of 100 W h Kg−1 with a capacity retention rate of over 76% after 100 cycles. The electrode has longer cycle stability under ZnSO4 electrolyte than K2SO4 and Na2SO4 solutions. Although ZnHCF as a cathode material can significantly improve the working voltage of ZIBs, compared with Mn oxide cathode material, the lack of cycle performance of PBA analogues has always been a difficult problem that limits its large-scale maturity. Ni et al. introduced Mn ions into ZnHCF compounds by simple precipitation to study the effect of metals connected with N bonds on Zn intercalation chemistry and to improve electrochemical stability.81 A series of manganese-substituted ZnHCF materials (MZHCFs) were used as cathode materials for aqueous zinc ion batteries (ARZIBs), where an Mn content of 7% exhibited optimal cycling performance. Because the substitution of Mn ions inhibits the cuboidal to rhomboid phase transition, the solid solution mechanism becomes dominant, thus improving the structural stability during the insertion and extraction of Zn ions. Thus MZHCF exhibits a significantly improved capacity retention rate during the constant current cycle compared to the MnHCF and ZnHCF materials in ARZIBs (Fig. 10b).
In order to better understand the aging mechanism of CuHCF, Kasiri et al. studied the influence of electrolyte concentration and properties on the electrochemical performance of CuHCF.129 The results showed that the electrolyte had great influence on the degradation mechanism of CuHCF. In addition to the influence of the properties of different ions in the electrolyte, the ion concentration also affects the properties of the material (the higher the electrolyte concentration, the faster the aging of CuHCF). From the potential curves of CuHCF cycled in 100 mM ZnSO4, we can observe the formation of a two-step plateau and the change in average cell voltage at higher cycling concentrations. In particular, it is observed that higher electrolyte concentration and lower current rate lead to faster degradation of a positive electrode. The results show that the phase transition of CuHCF occurs when it is cycled in high concentration ZnSO4 and 100 mM Zn(ClO4)2 electrolyte, which negatively affects the aging of the active material. Kasiri et al. delayed the degradation mechanism by optimizing the structure of CuHCF by using Zn2+ in the synthesis.87 The experimental results show that the CuZnHCF mixture with a Cu:Zn ratio of 93:7 exhibits excellent specific charge and stability, and two different phases are produced during the charge–discharge cycle of CuHCF. The first phase (cubic) and the second phase (non-cubic) nanoparticles are found to increase the zinc-rich composition of the CuZnHCF (93:7) mixture as expected.
In 2014, Trócoli et al.85 synthesized Cu-HCF from Cu(NO3)2·3H2O and K3Fe(CN)6 with similar stoichiometric ratios, based on a 20 mM ZnSO4 aqueous solution (pH 6.0) electrolyte. The charge retention rate of the CuHCF–Zn cell after 100 cycles was 96.3% of the maximum charge (the 15th cycle). However, the stability of CuHCF in ZnSO4 is slightly worse than that of classical electrolytes KNO3 and HNO3 (pH = 2), and a phase transition occurs during cycling at higher electrolyte concentrations (100 mM). In order to improve the stability of CuHCF, the group added sodium salt 2 M NaClO4 into zinc-based electrolyte Zn(ClO4)2 to obtain a mixed ion battery.130 The results showed that after 500 cycles, the initial charge retention rate of the battery with NaClO4 was 73.3% higher than that without NaClO4 (58.1%). It is inferred that sodium can improve the electrochemical performance of the system by slowing down the phase transition. At the same time, the content of zinc in CuHCF is calculated to be about 1.5–2% that of sodium by mathematical treatment.
As mentioned previously, the addition of a surfactant to the synthesis process can effectively reduce crystal defects. Cao et al. controlled the reaction process by adding PVP to the synthesis process and obtained a cubic granular KMHCF material with low water content and low defect uniformity.91 In this study, PVP can not only induce anisotropic growth through preferential adsorption to a single crystal plane, but also provide metastability through hydration-based spatial repulsion. The results show that the KMHCF-PVP-80 electrode also exhibits excellent long-cycle performance at other higher current densities. It has been reported that anorganic conductive polymer and Mn-PBA coated materials can also remedy this issue. Chen et al. coated polypyrrole (ppy) on the external surface of a KMnHCF cube, which not only effectively prolonged its cycle life but also led to excellent cycle performance.135 This is attributed to the limitation of Mn dissolution during electrochemical cycling by using ppy as a protective layer and its excellent electronic conductivity.
A recent study reported that the original K2MnFe(CN)6 cathode was gradually transformed into rhomboidic K2Zn3[Fe(CN)6]2 by the introduction of Zn2+ during the electrochemical cycle, and the introduction of Zn2+ caused a strong John-Teller effect on Mn3+, leading to strong lattice distortion (Fig. 10e).50 Together with the disproportionation reaction of manganese, the MnN6 octahedron is replaced by the ZnN4 tetrahedron, which eventually generates a new K2Zn3[Fe(CN)6]2 phase. The resulting solid structure of the K2Zn3[Fe(CN)6]2 phase contains wider channels for accommodating divalent ions, thereby enabling highly stable and reversible storage of Zn2+ ions. They also theoretically calculated different MnHCFs, in which the MnN6 octahedral surface exhibits a modest Jahn–Teller effect during K+ deintercalation from K2MnFe(CN)6 to MnFe(CN)6, and the crystal structure maintains a monoclinic skeleton without phase transition. In contrast, with Zn2+ insertion, some of the octahedral MnN6 are highly distorted and even become tetrahedral MnN4, indicating that Zn2+ induces a strong Jahn–Teller effect.
Previous research has demonstrated that hollow structure construction is an effective method to reduce structural strain during ion insertion/ejection and enhance the stability of PBA.139,140 Based on this, Zeng et al.92 proposed an effective self-template strategy (Fig. 10f) to construct cobalt-rich substituted manganese-rich PBA hollow spheres (represented as CoMn-PBA HSs) by simple anion exchange. Due to the higher specific surface area, sufficient active sites, and low coordinated water content of CoMnPBA HSs, the electrochemical performance is better than that of Co-PBA HSs and Mn-PBA HS. In contrast, the Co-PBA HS electrode has low current density and poor electrochemical activity. The CoMn-PBA HS electrode has high current density and has a similar cyclic voltammetry (CV) curve to that of the Mn-PBA HS electrode. Capacitance recovery shows excellent magnification performance as the current density decreases back to 0.1 A g−1. In addition, after 1000 cycles, it exhibits a coulombic efficiency close to 100%.
Despite the fact that numerous studies have focused on enhancing the capacity and cycling performance of V-PBA, the inherent low conductivity of V-PBA remains a limitation. Combining PBA with conductive agents to improve the conductivity of electrons is a viable remedy to this problem. Xue et al.94 designed a VHCF/CNT hybrid material in which vanadium hexanoate nanoparticles were grown on carbon nanotubes by in situ coprecipitation. On the one hand, the inherent 3D open framework of VHCF provides sufficient ion diffusion paths. On the other hand, since VHCF nanoparticles are connected through cross-linked CNTs, the conductive network formed by CNTs can accelerate the electron transfer between VHCF nanoparticles and promote the full utilization of active materials. However, considering that a single conductive carbon framework cannot provide adequate electrical conductivity, a double conductive carbon framework composed of internal and external conductive carbon components can provide higher electronic conductivity and better electrochemical properties. Wang et al. synthesized K1.14(VO)3.33[Fe(CN)6]2·6.8H2O cathode material, through the electron reaction of three different REDOX pairs V5+/4+, V4+/3+ and Fe3+/2+, and a dual conductive framework constructed from CNTs and Super P (SP).143 Benefiting from its structural and morphology engineering, the KVHCF@DCCF cathode material has a high specific capacity of 180 mA h g−1 at 400 mA g−1.
There is a strong chemical/electrochemical interaction between polar water molecules and PB/PBAs in aqueous electrolyte, which has a negative impact on the stability of the electrode, and the operation will lead to inadequate dissolution of active substances. Due to the low water decomposition voltage of 1.23 V, the effective potential window of a water medium is also restricted. On the one hand, the minimal operating voltage increases the device's energy density. On the other hand, the narrow potential window does not permit the high operating voltage of the PB/PBA cathode, which may result in the hazard of water-induced side reactions (such as the hydrogen/oxygen evolution reaction (HER/OER)) and consequently the device performance. Approaches to expanding the electrochemical stability window include adjusting the pH of the electrolyte and utilizing a “water in salt” (WIS) electrolyte system. The most frequent reason for adjusting pH is to regulate the decomposition of aqueous electrolytes.146–148 Xia et al. inhibited hydrogen evolution by adjusting the pH value of the electrolyte to 13, and shifted the potential of hydrogen evolution from 0.43 V to 0.8 V. In this way, the electrochemical window will be shifted as a whole, while the entire electrochemical stability window remains the same width in the aqueous electrolytic liquid system.146
A new type of “water in salt” (WIS) electrolyte system can effectively inhibit the dissolution and side reactions of PB/PBAs due to its extremely low water content.149–151 Liu et al. developed an acetonitrile/water-in-salt (AWIS) mixed electrolyte. Compared with the aqueous electrolyte, the AWIS mixed electrolyte prolonged the life of a ZnZn battery from 150 hours to 2500 hours and increased the upper limit cut-off voltage of a Zn–MnO2 battery from 1.8 V to 2.2 V.150 In addition, it has been reported that the Zn-PB battery capacity, rate ability, and cycle stability are enhanced after high voltage scanning at 2.3 V (vs. Zn/Zn2+) in the HCZLE salt-covered water electrolyte of 21 M lithium trifluoromethane sulfonyl and 1 M zinc bis (trifluoromethane sulfonyl) imide. This improvement is a result of the activation of C-linked low FeHCF.138
The strong solvation effect of 30 M KFSI + 1 M Zn(CF3SO3)2 salt solution in water can significantly reduce the activity of water in aqueous solutions. This pronounced solvation effect is supported by Raman spectroscopy results (Fig. 11a). With increasing electrolyte concentration, the corresponding H–O stretching peak shifts to 3557 cm−1, indicating a robust coordination between water molecules and potassium ions. The XRD analysis of a single charge and discharge cycle also provides evidence that a high concentration electrolyte can effectively maintain the structural stability of the positive electrode KZnHCF (Fig. 11c). Specifically, no significant changes in the structure were observed at both the initial stage of charging and the final stage of discharging. Furthermore, the atomic ratio of the KZnHCF electrode pre- and post-cycling substantiates the incorporation of Zn2+ ions and underscores the structural stability of [Fe(CN)6]4− (Fig. 11b).153 Other studies have demonstrated a significant reduction in the concentration of Mn within the KMnHCF electrode when exposed to diluted electrolyte, with no detectable presence of manganese observed in the electrode after 100 cycles using a 30 M KFSI + 1 M Zn(CF3SO3)2 electrolyte (Fig. 11d).50 Li et al. demonstrated that diluted hydrolysates exhibit limited reversibility towards Zn2+ ions, and the presence of NiHCF in diluted electrolytes leads to the formation of novel phases. However, this phenomenon is effectively suppressed in highly concentrated 1 M Zn(TFSI)2 + 21 M LiTFSI electrolytes (Fig. 11e).154
Fig. 11 (a) Raman spectra of the two mixed electrolytes and pure water (H2O) observed in the range of 2500–4000 cm−1 corresponding to the O–H stretching modes of water molecules. (b) Atomic ratio of the KZnHCF cathode before and after cycles. (c) Structure stability analysis of KZnHCF. In situ XRD test of the KZnHCF/Zn cell in the MC electrolyte and the corresponding charge–discharge curve. (Reprinted with permission from ref. 153. Copyright 2022, American Chemical Society). (d) The EDS results corresponding to the KMnHCF electrode after 100 cycles in diluted electrolyte and WIS electrolyte at 0.3 A g−1. (Reprinted with permission from ref. 50. Copyright 2021, John Wiley and Sons). (e) XRD patterns for structure evolution of the NiHCF electrode over 50 cycles in different electrolytes. (Reprinted with permission from ref. 154. Copyright 2020, John Wiley and Sons). (f) XRD results of MnHCF after the 1st, 100th, and 500th cycles using gel-0.3. (g) The morphology of MnHCF after 200th cycles using gel-0.3. (Reprinted with permission from ref. 162. Copyright 2023, American Chemical Society). |
Despite the fact that the water-salt system can reduce the water content to a certain extent, batteries with a specific mechanical strength and shape deformation are required in certain exceptional circumstances. With the continuous miniaturization of electronic chips, the development of integrated electronic devices, such as implantable medical devices, wearable health monitoring systems, flexible displays, and smart clothing, has attracted the attention of scientists around the globe.155–157 The gel electrolyte has some mechanical strength and deformation capability, and the cross-linked polymer network can interact powerfully with polar water molecules and PB/PBAs. Typically, quasi-solid gel electrolytes are prepared by dissolving inorganic salts in a fluid polymer backbone, such as polyvinyl alcohol (PVA),158 polyacrylamide (PAM),138 sodium alginate (SA),159 and sodium carboxymethyl cellulose (CMC).160 After the solid/gel interface has been assembled into a device, it inhibits the dissolution of PB/PBAs and the interface water decomposition reaction effectively and has a lengthy service life. For instance, WIG electrolytes can attain robust properties due to SA's strong affinity for water. The formation of hydrogen bonds between water molecules and polar groups on the SA chain (including OH− and COO−) permits the storage of unbound water within the WIG electrolyte. The electrolyte enlarges the electrochemical window and reduces decomposition-induced water loss. NaCl/ZnSO4/sodium alginate electrolyte (NaCl/ZnSO4/SA) serves as the electrolyte to match the CuHCF cathode in order to obtain high performance in a Na–Zn hybrid battery.159
Zhang and his colleagues107 showed the practical application of zinc ion batteries in flexible wearable electronics. The flexible quasi-solid-state battery was prepared by coupling ZnHCF@MnO2 with Zn foil in ZnSO4/PVA gel electrolyte. The current change of the battery can be ignored under different bending angles, showing excellent flexibility. In addition, the discharge capacities of the flexible device at 100, 200, 400, 500 and 800 mA g−1 were 89, 78, 67, 58 and 53 mA h g−1, respectively. In addition to the higher discharge capacity, the quasi-solid state battery also has a higher rate capacity. When the current density was increased from 100 mA g−1 to 800 mA g−1, the discharge capacity could still exceed 49 mA h g−1 with a capacity retention of 55%. At the same time, the flexible battery can stably cycle for more than 500 times and still maintain 71% capacity. Even folding the flexible battery can power the LED bulb (drive voltage of 1.8 V).
Zhi et al. studied the effect of different factors on the electrochemical performance of the Zn/MnHCF battery with PVHF/MXene-g-PMA as the electrolyte.161 This solid state Zn/MnHCF cell can operate normally at −35, and the capacity retention rate is 36.4% compared with that at 25 °C. When the operating temperature was increased to 100 °C, a significant capacity of up to 143.2 mA h g−1 was achieved. After high or low temperature testing, the capacity of all-solid Zn/MnHCF full cells can be well recovered when the temperature is restored to 25 °C. These results demonstrate the excellent environmental adaptability of an all-solid Zn/MnHCF whole-cell and confirm the remarkable thermal and freezing resistance of this type of solid-state cell. Although PBA as a flexible zinc ion cathode material has high voltage and good zinc storage performance, its development is still limited by the generally low lifetime caused by dissolution in the electrochemical cycle. The hydrogel electrolyte can act as a physical barrier to the dissolution of the anode material, especially for Mn-PBAs and PBA-based cathode materials. Luo et al. employed gelatin as a quasi-solid electrolyte to effectively suppress the dissolution of Mn in MnHCF. Throughout the cycling process, XRD analysis revealed that the cathode of MnHCF maintained its monoclinic phase (Fig. 11f), and SEM imaging demonstrated the preserved cubic morphology of the overall block structure, thereby indicating excellent cycling stability of MnHCF in gel-0.3 electrolyte (Fig. 11g).162 Zhi's group60 introduced a sol–gel transition strategy by introducing a hydrogel as a flexible electrolyte. The assembled Zn/CoFe(CN)6 cell achieves 2200 cycles and an excellent performance of nearly 100% coulombic efficiency.
Adding inorganic163 and organic164 additives to aqueous electrolytes is considered to be an effective strategy to modify the interfacial properties and inhibit the dissolution of PB/PBAs. Generally speaking, an electrolyte additive is a substance that improves the electrode/electrolyte interface to improve the electrochemical performance of the battery without participating in the electrode reaction, for example, the common organic surfactant sodium dodecylsulfonate (SDS). SDS molecules face the electrode and the electrolyte with their hydrophilic groups and hydrophobic groups, respectively, and are adsorbed on the electrode surface. The adsorbed SDS layer is on the outer hydrophobic interface, which effectively prevents the passage of water molecules and expands the electrochemical stability window of the electrolyte to about 2.5 V. In addition, based on the density functional theory experiments and calculation results, it is shown that SDS can effectively inhibit the dissolution of Mn in the Na2MnFe(CN)6 cathode, and effectively improve the cycle life of the Na–Zn hybrid battery.165
Material | Electrolyte | Working window [V] | Capacity [mA h g−1@mA g−1] | Cycle life [cycles, retention%@ mA g−1] | Ref. |
---|---|---|---|---|---|
Zn3[Fe(CN)6]2 | 3 M ZnSO4 | 0.8–1.9 | 66.5@60 | 200, 81@300 | 83 |
Zn3[Fe(CN)6]2 | 1 M ZnSO4 | 0.8–1.9 | 65.4@60 | 200, 80@300 | 49 |
K0.14Mn0.11Zn1.49[Fe(CN)6] | 1 M ZnSO4 | 1.0–2.0 | 41.6@50 | 500, 85.5@250 | 81 |
CuFe(CN)6 | 0.02 M ZnSO4 | 0.5–1.4 | 53@60 | 100, 81@600 | 84 |
KCuFe(CN)6 | 1 M ZnSO4 | 0.2–1.2 | 56@20 | 20, 77@20 | 43 |
CuZnHCF | 0.02 M ZnSO4 | 0.2–1.1 | 50@85 | 1000, 88.74@85 | 87 |
NiHCF/RGO | 2 M ZnSO4 | 0.7–1.8 | 50.1@200 | 1000, 80.3@200 | 88 |
K1.51Ni[Fe(CN)6]0.954(H2O)0.766 | 0.5 M Zn(ClO4)2 | 0.6–1.8 | 55.6@0.2C | 20, 99.9@0.2C | 48 |
NiFe(CN)6 | 0.5 M Na2SO4 + 0.05 M ZnSO4 | 0.9–1.9 | 76.2@100 | 1000, 81@500 | 152 |
K1.6Mn1.2Fe(CN)6 | 2 M Zn(ClO4)2 in tetraethylene glycol dimethyl ether | 0.7–1.9 | 65@50 | 8500, 94@200 | 33 |
K2MnFe(CN)6 | 30 M KFSI + 1 M Zn(CF3SO3)2 | 0.5–1.9 | 100@200 | 400, 72@200 | 50 |
KMHCF-PVP-80 | 1 M ZnSO4 + 0.1 M MnSO4 | 0.4–1.8 | 78@200 | 400, ≈100@100 | 91 |
KMHCF@PPy | 2 M ZnSO4 + 0.1 M MnSO4 | 0.4–1.8 | 110@100 | 100, 97.8@100 | 135 |
K0.05Fe(III)[Fe(III)(CN)6]·2.6H2O | 1.0 M Zn(OAc)2/([Ch]OAc + 30 wt% water) | 0.8–2.0 | 120@10 | 50, 97@60 | 136 |
Fe[Fe(CN)6]·2.5H2O | A bio-degradable ionic liquid-water mixture | 0.6–1.8 | 54@0.1 mA cm−3 | 50, 99.5@0.1 mA cm−3 | 137 |
FeHCF | 21 M lithium LiTFSI + 1 M zinc bis(trifluoromethanesulfonyl)imide | 0–2.3 | 76@1000 | 10000, 73@3000 | 138 |
CoFe(CN)6 | 4 M (Zn(OTf)2 | 0.7–1.8 | 109.5@6000 | 2200, ≈100@2000 | 60 |
CoMn-PBA HSs | 2 M Zn(CF3SO3)2 | 0.5–1.8 | 128.6@50 | 300, 76.4@10000 | 92 |
V3[Fe(CN)6]2 | 4 M Zn(CF3SO3)2 | 0.4–1.8 | 160@500 | 1000, 87.8@2000 | 93 |
VHCF/CNTs | 2 M ZnSO4 | 0.3–2.0 | 97.8@50 | 1000, 98@3200 | 94 |
VO-PBAs | 21 M (LiTFSI) + 1 M Zn(CF3SO3)2 | 0.3–2.1 | 209.6@100 | 2000, 95.5@1000 | 95 |
ZnHCF@MnO2 | 0.5 M ZnSO4 | 1.4–1.9 | 120@100 | 1000, 77@100 | 107 |
ZnHCF/PANI | 7.5 M ZnCl2 + 4 M NaCl | 0.8–2.1 | 50@50 | 350, 75@500 | 110 |
The concentration of metal ions within the PBA framework exhibits a positive feedback effect on the capacity and cycling performance of sodium ion and potassium ion batteries, when used as cathode materials. However, the impact of the basic metal ion content within the framework of PBAs on the mechanism for Zn2+ removal remains insufficiently elucidated. Insufficient evidence suggests that the framework exerts a favorable influence on Zn2+ during synthesis. In addition, further investigation is variations in water content affect the crystal structure of Prussian blue. Notably, when subjected to vacuum drying, Prussian blue materials used in sodium ion batteries experience a reduction in water content which eliminates water molecules from the organic electrolyte system. In an aqueous-based zinc ion battery, it is unknown whether water will be continuously embedded in the lattice as the electrochemical reaction progresses, such that the amount of water molecules in the lattice will influence the transmission of metal ions and electrons. It is unclear, however, if the water molecules in the lattice will impact the transmission of metal ions and electrons. This is the distinction between an aqueous electrolyte and an organic electrolyte, so the discussion on this aspect is also a breakthrough in elucidating the mechanism. In summary, the research on high-performance PBAs should focus on effectively controlling the phase transition of defect water content and investigating the characteristics of lattice strain. For MnHCF, additional efforts are necessary to suppress the Jahn–Teller effect in order to enhance cycle stability.
In essence, the research on high-performance PBAs should primarily focus on effectively controlling the phase transition of defect water content and investigating the characteristics of lattice strain. Regarding MnHCF, additional endeavors are necessary to suppress the Jahn–Teller effect for enhancing cycle stability. From a practical application standpoint, the consideration of volume energy density becomes pivotal. Extensive research has been dedicated to the doping of transition metals and the development of composite materials thus far. Consequently, increasing novelty also involves enhancing ion utilization efficiency through the design of morphology and structure, such as incorporating hollow structures and core–shell structures. However, there is a dearth of research on modifying the cathode/electrolyte interface, excluding any modifications to the PBA material itself. The energy barrier for the embedding of completely dehydrated metal ions in a crystal lattice is distinctly different when they diffuse from the electrolyte to the electrode interface. Furthermore, the involvement of metal hydrated ions in intercalation forms at the electrode interface plays a pivotal role in battery dynamics. Moreover, exploring system and ion chemistry through simulation calculations and studying interfacial impedance represents a promising avenue for future research. It is worth noting that certain bimetallic PBAs exhibit substantial potential for research and development. In such a structure, one metal component acts as a supportive framework while the other imparts active sites. The diverse interactions among transition metals play a crucial role in mutually enhancing their electrochemical performance, thereby surpassing that of PBAs containing a single metal. Additionally, novel configurations or enhanced characteristics may be unveiled during the course of the investigation, and thus, the exploration of bimetallic PBAs holds potential for ground breaking outcomes. Increasing the specific surface area is a promising approach to enhance the electrochemical efficacy of PBAs. Exploring novel technologies for synthesizing particles with diverse morphologies and structures, as well as controlling various other factors, represents an intriguing research direction that will impact the electrochemical performance of manufactured goods. In conclusion, we anticipate that this study will offer researchers a comprehensive understanding of the current state of cathode materials based on Prussian blue analogues for zinc-ion batteries, while also inspiring them to identify and address challenges, as well as develop novel PBAs materials.
This journal is © The Royal Society of Chemistry 2024 |