You are on page 1of 48

Journal Pre-proofs

Review

Recent advances in energy storage mechanism of aqueous zinc-ion batteries

Duo Chen, Mengjie Lu, Dong Cai, Hang Yang, Wei Han

PII: S2095-4956(20)30428-9
DOI: https://doi.org/10.1016/j.jechem.2020.06.016
Reference: JECHEM 1392

To appear in: Journal of Energy Chemistry

Received Date: 19 May 2020


Revised Date: 15 June 2020
Accepted Date: 15 June 2020

Please cite this article as: D. Chen, M. Lu, D. Cai, H. Yang, W. Han, Recent advances in energy storage
mechanism of aqueous zinc-ion batteries, Journal of Energy Chemistry (2020), doi: https://doi.org/10.1016/
j.jechem.2020.06.016

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by ELSEVIER B.V. and Science Press on behalf of Science Press and Dalian Institute of
Chemical Physics, Chinese Academy of Sciences.
Review

Recent advances in energy storage mechanism of aqueous zinc-ion

batteries
Duo Chena, Mengjie Lua, Dong Caia, Hang Yanga, Wei Hana,b,*

a Sino-Russian International Joint Laboratory for Clean Energy and Energy


Conversion Technology, College of Physics, Jilin university, Changchun 130012,Jilin,
China
b International Center of Future Science, Jilin University, Changchun 130012, Jilin,
China
*Corresponding author.
Email addresses: whan@jlu.edu.cn (W. Han)
Abstract
Aqueous rechargeable zinc-ion batteries (ZIBs) have recently attracted increasing
research interest due to their unparalleled safety, fantastic cost competitiveness and
promising capacity advantages compared with the commercial lithium ion batteries.
However, the disputed energy storage mechanism has been a confusing issue
restraining the development of ZIBs. Although a lot of efforts have been dedicated to
the exploration in battery chemistry, a comprehensive review that focuses on
summarizing the energy storage mechanisms of ZIBs is needed. Herein, the energy
storage mechanisms of aqueous rechargeable ZIBs are systematically reviewed in
detail and summarized as four types, which are traditional Zn2+ insertion chemistry,
dual ions co-insertion, chemical conversion reaction and coordination reaction of Zn2+
with organic cathodes. Furthermore, the promising exploration directions and rational
prospects are also proposed in this review.

Keywords: Zinc-ion batteries; Energy storage mechanism; Rechargeable aqueous


battery; Zn-MnO2 battery; Electrolytic battery
Photos and biographies

Duo Chen received his B.S. in Physics from Jilin University and Tomsk Polytechnic
University in 2016. He is currently pursuing the Ph.D at the Sino-Russian
International Joint Laboratory for Clean Energy and Energy Conversion Technology,
Jilin University, supervised by Prof. Wei Han. His current research focuses on
exploring energy storage mechanisms of Zinc-ion batteries.
Mengjie Lu received her B.S. degree in Physics from Jilin University in 2016. She is
currently pursuing the Ph.D. under the supervision of Prof. Wei Han at the
Sino-Russian International Joint Laboratory for Clean Energy and Energy Conversion
Technology, Jilin University. Her research focuses on the design and synthesis of
electrode materials for energy storage and conversion, including supercapacitors,
metal-ion batteries and water splitting.

Dong Cai is a Ph.D. candidate from the Key Laboratory of Physics and Technology
for Advanced Batteries (Ministry of Education), Jilin University, advised by Dr. Wei
Han. His research interest focuses on electrochemical energy storage systems, mainly
including lithium sulfur/ion batteries, supercapacitors, and full cells based on
transition-metal-based nano-materials together with hieratical porous carbons.

Hang Yang received bachelor's degree (2014–2018) in Physics from Jilin University
(PR China). In 2018, he joined the Sino-Russian International Joint Laboratory for
Clean Energy and Energy Conversion Technology in Jilin University as a master
student advised by Dr. Wei Han. His current research interests include the
development of high-performance zinc-ion batteries and design of next-generation
energy storage devices.

Wei Han received his Ph.D. degree from Tomsk Polytechnic University, Russia, in
1997. He joined the College of Physics, Jilin University as a Professor in 2001. His
current research focused on synthesis of nanomaterials and their potential applications
in energy storage devices and nuclear waste water purification.

1. Introduction
Increasing research interest has been attracted to develop the next-generation
energy storage device as the substitution of lithium-ion batteries (LIBs), considering
the potential safety issue and the resource deficiency [1–3]. In particular, aqueous
rechargeable zinc-ion batteries (ZIBs) are becoming one of the most promising
alternatives owing to their reliable security without any risk of flame or explosion,
huge cost competitiveness, eco-friendliness, high theoretic capacity, impressive
long-term cycling stability and superior rate capability [4–7].
The earliest Zn battery can be traced back to 19th century, when the alkaline
Zn-MnO2 batteries had ever dominated the disposable battery market at that time [8].
However, non-rechargeability of the early alkaline Zn-Mn batteries brought out
severe waste of resources and environmental issues. Although the problems got
alleviated as the development of rechargeable alkaline Zn batteries, the batteries
suffer from notorious zinc dendrites and low efficiency caused by inescapable
formation of byproduct and inferior cycling capability, which impeded its step toward
commercial application [9,10]. In addition, the rechargeable alkaline Zn batteries are
enabled via chemical conversion of cathode (such as Mn-, Co- or Ni-based oxides)
and dissolution/precipitation reactions of Zn anode (Zn + 4OH ― ↔Zn(OH)24 ― +2e ―
↔ ZnO + 2OH ― + H2O + 2e ― ), where there is no zinc ions (de)intercalation in
cathode reaction, therefore, they cannot be named ZIBs [11,12]. Up to recent years,
ZIBs in the true sense using mild Zn2+-containing electrolyte and insertion cathode
behave increasingly impressively because both the cycling reversibility and energy
storage capability are significantly improved compared to traditional alkaline Zn
batteries [13–15].
The main components of rechargeable aqueous ZIBs are Zn metal anode, Zn2+
host-type cathode, electrolyte and separator, where the devise for both electrodes
attracted pours of attentions [16,17]. A series of efforts have been made to develop
sufficient dendrite-free zinc anode to ensure the long-term stable zinc
plating/stripping [18,19]. Despite that, all of the anodes play the same role of
providing reversible Zn2+/Zn conversion for the battery system, following a defined
process of Zn↔Zn2 + +2e ― [20]. Hence diverse electrochemical performances of
ZIBs mainly depend on the varieties of cathodes, which serve as the charge
carrier-host via different electrochemical mechanisms in the energy storage process
[21]. Current cathode materials are mostly obtained from manganese-based oxides
(e.g. α-, β-, δ- or γ-MnO2), vanadium-based oxides or vanadates (e.g. V2O5, VO2,
V2O3, V6O13, NH4V3O8, NaV3O8, CuV2O7), Prussian blue analogues (e.g. ZnHCF,
CuHCF, FeHCF, NiHCF, CoFe(CN)6), other relative materials that possess analogous
crystalline structures (e.g. VS2, VS4, VSe2, MoS2, NASICONs), and some organic
polymers (e.g. Quinone compounds, some MOFs and COFs) [2]. It differs from the
anodes that almost every cathode material seems to possess distinctive reaction
inherence, which could be embodied in various operate voltage ranges, kinetics
characterizations, rate capabilities, cycling stabilities, as well as discharge capacities
[22]. Cathode chemistries are sophisticated that it is generally combined with the
multitudinous influence factors.
Fig. 1. Illustration of key factors in energy storage mechanisms of aqueous
rechargeable ZIBs.
Although numerous researchers for ZIBs about various cathode materials or
battery systems have been reported, the energy storage mechanism is still debatable
and ambiguous [9,17]. Besides the typical Zn2+ intercalation chemistry, other reaction
mechanisms benefitting to zinc-ion storage have been also demonstrated (as seen in
Fig. 1), such as co-insertion of cations (H+, Li+, Na+), structural water in cathode
crystals, reversible chemical conversion reaction with charge transfer, coordination
reaction of Zn2+ with organic cathodes, which would play crucial roles in guiding to
design novel high-performance ZIBs. This review will comprehensively summarize
the state-of-the-art exploration progress in energy storage mechanisms for ZIBs and
further point out the promising strategies toward designing high-performance
cathodes.
2. Energy storage mechanism
2.1. Traditional Zn2+ insertion chemistry
The reversible Zn2+ insertion/extraction in the host materials is the most common
energy storage mechanism, which is similar to traditional Li-ion batteries. In the
discharge process, zinc ions as the charge carriers are intercalated into the cathode,
which receives electron with the decrease of the oxidation state; in the reverse charge
process, the cathode release Zn2+ and gets oxidized at the effect of the electric field.
The process can be summarized as universal equations as follow:
Anode:
Zn↔Zn2 + + 2e ― (1)
Cathode:
M + 𝑥Zn2 + + 2𝑥e ― ↔Zn𝑥M (2)
Overall:
𝑥Zn + M↔Zn𝑥M (3)

where M refer to the cathode materials. Materials with tunnel, layered or open
framework structure which possess sufficient host room for Zn2+ are considered as
ideal cathode materials of ZIBs as shown in Fig. 2(a).
Tunnel structural materials include some vanadium oxides and manganese
oxides, for example, α-MnO2 (2 × 2 tunnels), β-MnO2 (1 × 1 tunnels), γ-MnO2 (1 × 1
and 1 × 2 tunnels), todorokite-type MnO2 (3 × 3 tunnels), VO2 (B) (2 × 2 tunnels)
[2,10]. In the early stage, manganese oxides got the most commonly studied and
Zn/α-MnO2 battery was successfully fabricated firstly in 2012 [23]. Although the
mechanism of Zn2+ reversible (de)intercalation in tunnels is recognized to be defined,
the dissolution of cathode and the accompanied phase change during electrochemical
process still make the mechanism controversial for Zn/MnO2 batteries, for instance,
the phase transition between Zn-birnessite and α-MnO2 (Fig. 2b) [24,25]. Unlike the
undesired phase changes of MnO2, the tunnel structural VO2 (B) possesses long-term
robust stability [26]. VO2 (B) was firstly reported as an efficient cathode by Yang et
al. at 2018 [27]. Considered that the big-size tunnels (0.82 and 0.5 nm2 along the b-
and c-axes) provided unimpeded Zn2+ diffusion pathway, VO2 (B) electrode exhibited
high reversible capacity of 357 mAh g−1 with impressive rate capability (171 mAh g−1
at an ultrahigh current density of 300 C). Besides, as seen in Fig. 2(c), Mai et al.
demonstrated the highly reversible single-phase reaction of the Zn2+ insertion in VO2
utilizing operando techniques and corresponding qualitative analyses [28].
Fig. 2. The traditional Zn2+ insertion chemistry and basic structures of cathode
materials. (a) Crystalline structures of tunnel α-MnO2, layered δ-MnO2, tunnel VO2
(B), layered V2O5, open framework of MFe(CN)6 (M=Fe, Co, Ni, Cu, Mn),
respectively from left to right. (b) Schematic of the phase transition of α-MnO2 as
zinc-ion intercalation. Reproduced with permission [24]. Copyright 2014, Nature
Publishing Group. (c) Diagram of the single-phase reaction of VO2 in different
discharge stages. Reproduced with permission [28]. Copyright 2019, American
Chemical Society. (d) Cycling performance of δ-MnO2 cathode. Reproduced with
permission [29]. Copyright 2015, Elsevier. (e) Schematic of the Zn2+ (de)intercalation
in the CoFe(CN)6 electrode with two-species redox reactions. Reproduced with
permission [40]. Copyright 2019, Wiley-VCH.
Layered structure generally exists in a small part of manganese oxides, a number
of vanadium oxides, vanadates, and so on. Typically, layered δ-MnO2 can deliver a
high capacity of 252 mAh g−1 at the current density of 83 mA g−1 when introduced in
aqueous ZIBs application [29]. Unfortunately, the phase transition to spinel ZnMn2O4
(i.e., Zn2 + +2MnO2 +2e ― ↔ZnMn2O4) was observed in the dis/charge process,

which led to the fast capacity degradation (Fig. 2d). In particular, another report from
Liu et al. demonstrated that no phase transition occurred in layered δ-MnO2, which
endowed it with superior cycling stability than tunnel manganese oxides [30].
Vanadium-based compound is a huge family due to the diversity of V−O coordination
and the crystalline structures, and up to now, many of these have been developed as
the cathode for zinc-ion batteries, including V2O5, V3O7·H2O, V5O12·6H2O, V6O13,
VOPO4, Na3V2(PO4)3, etc. [31–38]. Rechargeable aqueous Zn-V2O5 battery is a
common system representing the work mechanism of layered structural cathodes [10].
According to Zhang’s report, the Zn-V2O5 battery exhibited a high capacity of 470
mAh g−1 at 0.2 A g−1 with excellent rate capability and long-term cyclability (91.1%
capacity retention over 4000 cycles at 5 A g−1) benefiting from the large accessible
interlayer space for Zn2+ and the highly reversible (de)intercalation through the
layered structure. Layer structured V6O13 is also one of the suitable cathode materials.
Peng et al. synthesized oxygen-deficient V6O13 that weakened the strong columbic
ion-lattice interactions with divalent Zn2+, facilitating the reaction kinetic of the ion
intercalation [36]. Therefore, the O-deficient V6O13 showed a high specific capacity
of 400 mAh g−1 at 0.2 A g−1, reaching 95% of its theoretical capacity, and an
outstanding cycling stability with negligible decline of capacity at this small current
density.
Unlike the tunnel or layered structure of Mn-/V-based oxides, Prussian blue
analogues (PBAs) refer to the transition-metal hexacyanoferrates (MFe(CN)6, M=Fe,
Co, Ni, Cu, Mn and etc.) possessing 3D open framework structure with abundant
big-size interstitial sites, which are considered as a promising cathode material for
ZIBs [39,40]. In the discharge process, Zn2+ is inserted into the interstitial tunnels
accompanied with the reduction of the valence state of the transition-metal atom M or
Fe from +3 to +2; and the case would be reverse in charge process. The
electrochemical properties are usually determined by the atom that actively occurred
valence-state change. In 2014, Liu et al. firstly applied the zinc hexacyanoferrate
(ZnHCF) in aqueous zinc-ion battery, which exhibited a relative high operation
voltage of ~1.7 V [39]. However, the battery can only deliver a low capacity of 65.4
mAh g−1 at 1 C, and it is a common phenomenon for majority of PABs-based ZIBs
that exhibited lower capacity. The inferior capacity is believed to be ascribed to single
atom redox activity of the MFe(CN)6 at electrochemical process in most cases. To
address this issue, Zhi et al. successfully fabricate a cobalt hexacyanoferrate
(CoFe(CN)6) electrode with two-species redox reaction both of Co(II)/Co(III) and
Fe(II)/Fe(III) as a high-performance cathode of aqueous ZIB (Fig. 2e) [40]. Due to the
efficient mechanism of two-species electrochemical redox activity, the Zn/
CoFe(CN)6 battery can deliver a significantly promoted capacity of 173.4 mAh g−1 at
0.3 A g−1, simultaneously possess a high operation voltage plateau of 1.75 V.
Beyond that, as one of the significant electrode materials in Li+/Na+ batteries, the
rhombohedral NASICON-type compounds have also exhibited remarkable
performances in zinc-ion storage. However, it was generally acknowledged that the
M1(6b) sites in NASICON were inactive with only M2(18e) sites enabled ion
(de)intercalation, which led to the limited capacity. Cui et al. found a concerted
Na+/Zn2+ transfer mechanism in Zn/Na3V2(PO4)3 battery, that both M1 and M2 sites
could be occupied by Na+/Zn2+ [41]. The concerted intercalation in Na3V2(PO4)3 not
only enhanced the energy storage capacity by activating the M1 sites of
NASICON-type compounds, but also improved the structural stability by the
occupation of mixed ions.
Fig. 3. Facilitation effect of structural water. (a) Crystal structure of hydrated
V2O5·nH2O in fully discharged state. (b) Comparation of rate capabilities of
V2O5·nH2O (VOG) and V2O5 (VOG-350). Reproduced with permission [42].
Copyright 2017, Wiley-VCH. (c) Diffusion paths of zinc-ion in V6O13 with hydrated
intercalation and (d) without water. (e) Diffusion energy barriers for Zn2+ in paths (c)
and (d). Reproduced with permission [43]. Copyright 2019, Wiley-VCH. (f) Zn2+ host
sites in hydrate NaCaVO (topside) and dry-NaCaVO (downside). Reproduced with
permission [44]. Copyright 2019, Wiley-VCH. (g) Solvent water and Zn2+ were
migrated with opposite direction in Ca0.67V8O20·3.5H2O cathode. Reproduced with
permission [45]. Copyright 2019, American Chemical Society.
2.1.1. Facilitation effect of structural water
Structural water between the oxide layers is demonstrated to play a crucial role
in the Zn2+ insertion process. The ZIBs based on the bared oxide cathode usually
suffer from the sluggish diffusion kinetics due to the strong electrostatic repulsion
originated from the big ratio between charge and mass of the zinc ion [3,27]. The
introduced structural water in the interlayers can reduce the effective charge of zinc
ions by the solvation, which effectively weaken the electrostatic interactions with the
oxide framework (Fig. 3a) [42]. The phenomenon of the facilitation by structural
water was firstly reported by Yang and co-workers in 2018 via a study on
Zn/V2O5·nH2O battery and the case was called “water-lubricating” effect. To
experimentally demonstrate the water-lubricating effect, the dry sample without
structural water was synthesized via a calcined treatment at 350 °C, which exhibit an
obvious decrease whenever in capacity or rate capability compared to the structural
water-containing sample (Fig. 3b). In addition, the density functional theory (DFT)
calculations were employed to verify the vital role of structural water. Choi et al.
simulated the Zn2+ diffusion in the solid-state matrix of V6O13 with/without hydrated
intercalation (Fig. 3c, d), which exhibited the aqueous case endowed with a
significantly lower energy barrier of 0.87 eV than that of the anhydrous case of 2.85
eV (Fig. 3e) [43]. The similar evidence was also obtained from a
Zn/NaCa0.6V6O16·3H2O (Zn/NaCaVO) system [44]. Compared to the dry-NaCaVO
without structural water, the hydrated NaCaVO possessed larger diffusion-coefficient
and smaller energy barrier for Zn2+ migration throughout the host lattice. The
structural water intercalation not only accelerated the reaction kinetics, but also
enlarged the gallery spacing to provide more accessible sites for Zn2+ storage, as
shown in Fig. 3(f), in which two active sites in hydrated NaCaVO would be
degenerated to only one in dry-NaCaVO due to the decreased gallery spacing.
Thereby, the Zn/NaCaVO aqueous battery could deliver an enhanced capacity of 347
mAh g−1 at the current density of 0.1 A g−1.
Except the structural water, the solvent water is also found to (de)intercalate in
the interlayer during the electrochemical process of aqueous ZIBs. For example, the
insertion of solvent water was demonstrated in the investigation on
Ca0.67V8O20·3.5H2O (CaVO) nanobelts cathode [45]. In the first cycle, the solvent
water was intercalated into CaVO; in the subsequent process, the
Zn2+-(de)intercalation is accompanied by a counter migration of solvent water (Fig.
3g). The insertion of the solvent water can expand the interlayer spacing, facilitating
the diffusion kinetics of Zn2+.
2.1.2. The pre-insertion strategy
Alkali metal (e.g. Li+, Na+, K+, etc.), transition metal (e.g. Fe2+, Co2+, Ni2+, Cu2+,
Zn2+, Mn2+, etc.) or other metal (e.g. Ag+, Mg2+, Ca2+, La3+, etc.) ions pre-inserted in
the layered vanadium or manganese oxides can act as pillars to enlarge the interlayer
spacing, enhancing the ion diffusion during the charge-discharge process [44,46–55].
Besides, the pre-insertion of metal ions is able to prevent the collapse of the layer
structure upon repeated Zn2+ insertion/extraction, which provides an improvement of
cycling stability. As the aforementioned discussion that the water molecule
intercalation could enlarge the interlayer spacing in V2O5·nH2O, the interlayer
distance was further extended via a chemical intercalation of Li+ in the interlayer by
Liang et al. [46]. The expansion of interlayer can be observed by the XRD peak shift
toward lower-degree, where a largest d-spacing of (001) plane was calculated as 13.77
Å, as shown in Fig. 4(a). Compared with the alkali metal ions, the multivalent ions
were deemed to possess stronger ionic bonding with oxygen atoms in the interlayer,
which is conducive to maintain the structural stability. Based on that, Alshareef et al.
designed a bilayer calcium vanadium oxide (Ca0.25V2O5·nH2O) as a high-efficiency
ZIB intercalation cathode, demonstrating an impressive capacity of 340 mAh g−1 at
0.2 C and superior cycling durability at high current density of 80 C [56]. Moreover,
Zhi et al. demonstrated that the Co2+-inserted V2O5·nH2O (Co0.247V2O5·0.944H2O)
was endowed with a large interlayer distance of 10.7 Å (Fig. 4b), which supplied
significantly lower Zn2+ diffusion energy barrier in the interlayer than V2O5·nH2O by
DFT calculations (Fig. 4c), confirming the faster electrochemical kinetics in the
Zn/Co0.247V2O5·0.944H2O batteries [51]. Therefore, the chemical pre-intercalation has
become a universal strategy for boosting the zinc-ion storage, which was proved via a
comprehensive study on the performances of the layered vanadium oxide pre-inserted
with a series of transition metal ions (i.e. Fe2+, Co2+, Ni2+, Mn2+, Zn2+ and Cu2+) by
Zhou and co-workers [48].
Fig. 4. The strategy of cation pre-insertion. (a) XRD patterns of Li+ intercalated
V2O5·nH2O. Reproduced with permission [46]. Copyright 2018, Royal Society of
Chemistry. (b) Crystal structure of Co0.247V2O5·0.944H2O viewed along [110]
direction. (c) The energy barrier (Ea) for Zn2+ diffusion in Co0.247V2O5·0.944H2O and
V2O5·0.944H2O. Reproduced with permission [51]. Copyright 2019, Wiley-VCH. (d)
Element mapping images at discharge state and (e) ex-situ XPS spectra of Ag 3d of
Ag0.4V2O5 electrode. (f) Schematic of CDI mechanism of Zn/Ag0.4V2O5 battery.
Reproduced with permission [52]. Copyright 2018, Elsevier. (g) Schematic
illustration of the CDI reaction process in the Zn/CuV2O6 battery. Reproduced with
permission [59]. Copyright 2019, American Chemical Society.
It must be pointed out that the pre-intercalated ions would occupy the active sites
of Zn2+ even with the possibility of obstructing the diffusion pathway for insertion of
Zn2+, restraining the capacity of zinc-ion storage. Thus, it is important to take some
measures to avoid relative adverse effect. Interestingly, researchers found that some
ions at interlayers could be displaced by Zn2+ accompanied with a reduced
transformation into metal elemental substance in the discharge process [52,57–59].
For instance, except the basic intercalation of Zn2+, a part of the Ag0.4V2O5 cathode
was detected to transform into zinc vanadate (Zn2+x(V3O8)2) during discharging [52].
Some Ag+ in the interlayer of vanadium oxide was displaced by Zn2+ and reduced to
Ag0 (Fig. 4d), which not only provided exceptional capacity via electron transfer in
the redox reaction Ag + ↔Ag0, but also enhanced the electric conductivity of the
whole electrode. And the host structure could be recovered in the charging process as
demonstrated by XPS analysis (Fig. 4e). Such process was defined as the combination
displacement/intercalation (CDI) reaction mechanism as illustrated in Fig. 4(f). The
analogous process was also observed in Zn/CuV2O6 system (Fig. 4g), which was
demonstrated a high capacity of 427 mAh g−1 benefiting from the contribution of the
Cu2 + ↔Cu0 reduction/oxidation reaction [59].
The Zn2+ trapped in the host structure in the initial discharge process could
likewise act as pillars to guarantee the structural stability and enough interlayer
distance [60]. The Zn2+ located at the “dead Zn2+ sites” cannot be extracted out from
the host material during charging, which could be detected by the left Zn trace from
element mapping [61]. Although the phenomenon led to capacity decrease in initial
stage, a superior cycling stability would be contained in subsequent charge-discharge
processes.
The strategy of cation pre-insertion is not only suitable for vanadium oxides, but
also applied in tunnel or birnessite manganese oxides cathodes [25,54,62]. For
example, Lu et al. reported a La3+ pre-inserted δ-MnO2 (LMO) nanoflorets as
high-performance cathodes for aqueous rechargeable ZIBs [54]. Similarly, on the one
hand, the intercalated La3+ in layered δ-MnO2 enabled to enlarge the interlayer
distance, reducing the resistance of Zn2+ (de)intercalation; on the other hand, the ion
was also favorable for supporting the structure to avoid collapse or phase
transformation. Supported by the intercalated La3+, LMO exhibited admirable
capacity of 287.5 mAh g−1 and remarkable cycling stability. Yu and co-workers also
demonstrated the pre-insertion in birnessite as a promising strategy to fabricate
high-performance cathode materials for ZIBs [63]. Due to the pre-insertion of Na+
and crystal water, the layered Na0.55Mn2O4·0.57H2O showed enlarged interlayer
distance (7.25 Å) to facilitate the insertion and diffusion of Zn2+, thereby the cathode
delivered high reversible capacity of 389.8 mAh g−1 at the current of 0.2 A g−1.
Notably, the sodium ions were partially displaced by zinc ions in the initial cycling
stage, which enhanced the structural stability for subsequent cycles.
In addition to the various cations, some organic polymers have also been
reported as the interlayer modifier to pre-intercalate into the layered cathodes. For
example, Xia and co-workers developed polyaniline-intercalated layered MnO2 and
poly(3,4-ethylenedioxythiophene) (PEDOT) inserted NH4V3O8 cathodes for advanced
aqueous ZIBs [64,65]. Polyaniline intercalated V2O5 was also reported as
high-performanced electrode material [66]. The induced polymers could enlarge the
interplanar spacing of the original crystal lattice, providing rapid diffusion channels
for Zn2+. Importantly, the unique π-conjugated structure of the conductive polymers
can effectively constrain the electrostatic interactions between the host framework and
inserted Zn2+, which facilitated the ion diffusion kinetics [66]. Furthermore, the
polymer-intercalated structure as well exhibited improved cycling stability for
manganese dioxide via avoiding the phase changes [64].
2.2. Dual ions co-insertion
Considering the sluggish Zn2+ intercalation originated from the large-size
hydrate Zn2+ and the strong electrostatic repulsion of divalent ion, host materials are
theoretically allowed a simultaneous insertion of other ions with the higher diffusion
kinetics. The intercalative ions (such as H+, Li+, Na+, and so on) could promote the
effective utilization of the active sites at host structure, enhancing the performance of
the ZIBs [67–74].
A typical case is the co-insertion of H+ and Zn2+ in manganese oxides. Wang et
al. firstly proposed the dual-ion co-insertion mechanism in a rechargeable aqueous
Zn/ε-MnO2 system [67]. In this study, the different intercalation processes of H+ and
Zn2+ were vividly reflected by the discharge galvanostatic intermittent titration
technique (GITT) profiles, which can be divided into two regions with different
kinetic characterizations as shown in Fig. 5(a). In region I, the low voltage jump
revealed small ohm and charge transfer resistances, as well as fast ion diffusion,
which corresponded to the H+ insertion; the large voltage jumps in region II suggested
a slow diffusion of Zn2+. When the electrode was discharged to 1.3 V (region I), a
series of typical MnOOH peaks were observed from the XRD pattern (see Fig. 5b),
while the peaks of ZnMnO4 were obtained in the full discharged state at the end of
region II, which gave a convinced proof to H+ and Zn2+ insertion resulting in the
differences of the two regions. However, the various tunnel or layered manganese
oxides cathode (e.g. α-, β-, γ- or δ-phases) are usually hindered by notorious structure
collapse upon cycling on account of phase transition. To address the issue, Pan and
co-workers applied a novel phase of manganese oxide MnO2H0.16(H2O)0.27 to a
structurally robust zinc-ion battery, in which a particular H+/Zn2+ synergistic
intercalation mechanism was demonstrated [71]. In both two discharge stages the dual
ions of H+/Zn2+ are found to insert synergistically, following by the transition of

MnO2→MnO2H0.25Zn0.1→MnO2H0.45Zn0.2. The mechanism was also proved by


theoretic simulation using DFT, where the calculated discharge plateaus were in good
agreement with the experimental profiles (Fig. 5c), demonstrating the H+/Zn2+
synergistic intercalation reaction.
Fig. 5. Dual ions co-insertion mechanism. (a) GITT profiles of ε-MnO2 electrode and
(b) the corresponding ex-situ XRD evolution in 1.3 V and 1.0 V. Reproduced with
permission [67]. Copyright 2017, American Chemical Society. (c) Measured voltage
plateaus compared with DFT calculations in two discharge stages of
Zn/MnO2H0.16(H2O)0.27 battery. Reproduced with permission [71]. Copyright 2019,
Wiley-VCH. (d) Ex-situ XRD patterns and (e) solid state 1H NMR spectra of
NaV3O8·1.5H2O electrode. Reproduced with permission [68]. Copyright 2018, Nature
Publishing Group. (f) Schematic illustration of the crystalline structure evolution for
FeHCF when Li+ and Zn2+ (de)intercalated. Reproduced with permission [72].
Copyright 2019, Wiley-VCH.
The dual-ion co-insertion reaction was also observed in vanadium oxides based
ZIBs. The H+ intercalation was found to exist in Zn/NaV3O8·1.5H2O system by Niu
and co-workers [68]. In discharged stage, the characteristic peaks of
Zn4SO4(OH)6·4H2O were obtained in the ex-situ XRD patterns (Fig. 5d), which
indicated the consume of OH− originated from the water ionization, revealing the
remanent H+ was supposed to combine into cathode material to keep the neutral
charge system of the electrolyte condition. Simultaneously, the analysis of ex situ
solid-state 1H NMR gave out a direct proof that confirmed the continuous and
reversible insertion/extraction of the proton during discharge/charge process in the
sodium vanadate electrode (Fig. 5e). Xu et al. determined the similar proton
co-intercalation reaction in Zn/V10O24·12H2O battery utilizing the structure change of
proton insertion by-product and Zn2+ insertion product with various zinc salt anions
[69]. Notably, the insertion of Zn2+ would result in the lattice contraction of vanadium
oxides, while the hydronium intercalation tends to expand the lattice spacing. Based
on the neutralization effect between intercalated Zn2+ and hydronium, Zheng et al.
successfully achieved stably cycling of the Zn-vanadium oxide battery in which the
structure of cathode material almost remained constant during charge/discharge [75].
Besides, a new kind of ion insertion could be introduced into the aqueous
zinc-ion batteries when another salt was added in the electrolyte. Batyrbekuly and
co-workers reported a hybrid Zn/V2O5 battery with a dual Li+/Zn2+ intercalation
reaction using 3 M Li2SO4 + 4 M ZnSO4 as electrolyte [76]. Except for the Zn2+
insertion/extraction reaction, a reversible lithiation/delithiation process was
demonstrated via XRD and Raman spectra analysis. Specially, the V2O5 electrode in
aqueous hybrid battery performed better structural stability than that in nonaqueous
LIBs owing to the coinsertion of Zn2+ in the discharge process. As for Prussian blue
analogues, Zhi et al. found that the C-coordinated Fe in iron hexacyanoferrate
(FeHCF) could be activated by high-voltage operation, meanwhile the Zn2+ and Li+
co-intercalation was revealed in the aqueous hybrid-ion FeHCF-based battery [72]. As
seen in Fig. 5(f), the crystalline structure of FeHCF exhibited distortion along with the
zinc and lithium ion insertion at discharged state, while the extraction of these ions
could recover the crystalline structure upon charging, which might be observed by the
evolution of the crystalline interlayer spacing during the discharge/charge process
utilizing ex situ XRD patterns and TEM analysis.
2.3. Chemical conversion reaction
Compared with intercalation chemistry, conversion reactions in battery would
tend to deliver higher capacity due to the direct charge transfer in the reactions that
possess superior theoretic capacity. Hence introducing rational conversion reactions
into ZIBs is a promising and effective method to construct high-performance aqueous
batteries.
Liu et al. demonstrated a conversion reaction mechanism of Zn/α-MnO2 battery
that MnO2 was reacted with H+ into MOOH rather than inserted by Zn2+ [77]. To keep
the neutral charge Zn2+ and the sequent OH− were found to converted as
ZnSO4[Zn(OH)2]3·12H2O (zinc hydroxide sulfate, ZHS) precipitated on surface of
MnO2. Notably, MnSO4 as an additive was verified to restrain the dissolution of
MnO2 and promote the cycling stability, which was widely accepted and applied on
manganese-based zinc-ion battery. The conversion reaction of the reversible
hydrogenation/dehydrogenation of MnO2 accompanied by precipitation/dissolution of
ZHS can be formulated as below:
H2O↔H + + OH ― (4)
MnO2 + H + + e ― ↔MnOOH (5)
1 2+ 1 6 1
Zn + OH ― + ZnSO4 + H2O↔ ZnSO4[Zn(OH)2]3 ∙ 𝑥H2O (6)
2 6 𝑥 6
Fig. 6. Conversion reaction mechanism based on cationic conversion or anionic
redox. (a) Zn2+ and Mn2+ content evolution in different states corresponding to the
discharge-charge curve of Zn/α-MnO2 cell. (b) The pH variation in the first cycle.
Reproduced with permission [78]. Copyright 2016, Wiley-VCH. (c) The galvanostatic
discharge profiles of Zn-MnO2 battery and the schematic for different reaction
mechanisms in D1, D2 and D3 stages. Reproduced with permission [80]. Copyright
2019, Wiley-VCH. (d) Schematic illustration of charge storage mechanisms of
Mn-based ZIBs. Reproduced with permission [7]. Copyright 2020, American
Association. (e) Discharge/charge curves of Zn/VOPO4 battery, which can be divided
into two regions of oxygen redox in high voltage and vanadium redox in low voltage.
(f) Evolution diagram of energy versus density of state in VOPO4 as continually
charging. (g) Schematic illustration of reversible oxygen redox reaction in Zn/VOPO4
battery. Reproduced with permission [37]. Copyright 2019, Wiley-VCH.
However, the above conversion reaction mechanism is similar to a type of H+
insertion chemistry that we have discussed. Beyond that, the conversion reaction of
true sense was also proposed in the Zn/α-MnO2 system. In particular, Oh et al.
revealed another understanding of the formation of ZHS and the energy storage
process of α-MnO2 [78]. A monitoring of the concentration of Mn2+ and Zn2+ in
electrolyte was conducted (Fig. 6a), which displayed an increasing tendency of Mn2+
concentration in discharging process and downtrend upon charging, while an inverse
evolution was obtained for Zn2+. The reversible change of Mn2+ concentration in
dis/charge process strongly revealed that the α-MnO2 underwent a reduction and
dissolution process into Mn2+ when discharging, which was oxidized and deposited
onto cathode in charging reaction, as shown in reaction (7).
MnO2 + 2H2O + 2e ― ↔Mn2 + + 4OH ― (7)
Notably, the pH value of the electrolyte was supposed to rise in the discharging
process ascribed to the generation of OH ― , which was confirmed by the in-situ pH
monitoring (as shown in Fig. 6b). As the pH value increasing, the precipitation of
ZHS would start to form on cathode, which contributed to the decrease of Zn2+
concentration in electrolyte.
Shi et al. proposed a multiple reaction mechanism for the birnessite MnO2
cathode, in which the discharge reactions proceeded by two stages [79]. In the first
stage, the H+ and Zn2+ were intercalated successively as the battery discharging from
1.85 to ~1.3 V, where the MnO2 was transformed into MnOOH and Zn2Mn4O8·H2O;
subsequently, in the second stage, the obtained MnOOH and Zn2Mn4O8·H2O were
thoroughly converted into Mn2+ and ZHS as the voltage continue decreased to 1.0 V.
However, although the content of ion and the surface morphology of electrode were
employed to support the discovery, the inconsistent magnification of the SEM images
cast a wisp of shadow to the proposed conversion mechanism, which required more
analysis approaches to clarify the reaction process.
For the abovementioned disputed reaction mechanism of MnO2, Qiao and
co-workers gave a novel systematical explanation via an investigation of high-voltage
electrolytic Zn/MnO2 battery [80]. The multi-redox reactions of the Zn/MnO2 battery
exhibited strongly voltage-dependence as shown in Fig. 6(c), where the discharge
process at operation voltage of 2.2–0.8 V can be divided into three regions, that is D1
(2.0–1.7), D2 (1.7–1.4), and D3 (1.4–0.8). In the D1 region of high voltage, an
electrolytic conversion reaction was dominated, in which the manganese oxide with
+4 valence state was reduced and dissolved as manganese ions in + 2 into electrolyte;
in the subsequent D2 region, the MnO2 cathode underwent a conversion reaction with
H+ to form MnOOH, which was also considered as a process of the insertion of H+;
and in the D3 region of relative low voltage, a classic Zn2+ intercalation chemistry
was occurred. The battery reaction process could be summarized as below:
𝐷1:MnO2 + 4H + + 2e ― ↔Mn2 + + 2H2O (8)
𝐷2:MnO2 + H + + e ― ↔MnOOH (9)
𝐷3:MnO2 + 0.5Zn2 + + e ― ↔Zn0.5MnO2 (10)
It is worth to point that the electrolytic conversion reaction in D1 region was
found to be increased by tuning the pH of the electrolyte. As the additive of H2SO4
increasing at the initial electrolyte consisted of ZnSO4 and MnSO4, the contribution
ratio of the high-voltage capacity in D1 region was grown from 26% to approximate
100%. Benefiting from the unique two-electron conversion reaction, the fabricated
electrolytic Zn-MnO2 battery enabled a high discharge plateau of ~1.95 V and
delivered a record energy density of about 409 Wh kg−1. Furthermore, Qiao et al.
demonstrated that the electrolysis kinetics of the MnO2/Mn2+ electrolytic conversion
can be promoted via introducing the catalytic Ni2+ into the electrolyte [81].
Based on the abovementioned discussion, the charge storage mechanism in
Mn-based electrode can be summarized in four types, i.e., (i) Zn2+
insertion/extraction, (ii) H+ (de)intercalation accompanied with the
deposition/dissolution of ZHS, (iii) coinsertion/extraction of both H+ and Zn2+ in
different (dis)charge stage, and (iv) electrolysis/electrodeposition of MnO2/Mn2+, as
shown in Fig. 6(d) [7]. The electrochemical performances of different MnO2 cathodes
enabled by different mechanisms are displayed in Table 1, which reveals that the
electrolytic conversion reaction with two-electron transfer could deliver higher energy
storage capacity than others.
Table 1. Electrochemical performances of various MnO2 cathodes.
Materials Operating Discharge Cycling performance Battery mechanism
voltage (V) capacity
K1.33Mn8O16 [62] 1.0–1.8 312 mAh g−1 at 79 mAh g−1 at 5 C after Zn2+
0.1 C 650 cycles insertion/extraction
δ-MnO2 [29] 1.0–1.8 252 mAh g−1 at 112 mAh g−1 at 83 mA Zn2+
83 mA g−1 g-1 after 100 cycles insertion/extraction
Mn2O3 [82] 1.0–1.9 148 mAh g−1 at 82.2 mAh g−1 at 2 A g−1 Zn2+
0.1 A g−1 after 1000 cycles insertion/extraction
δ-MnO2 [30] 0–2.0 330.67 mAh g−1 ~160 mAh g−1 at 1.5 mA Co-insertion of H+
at 1.5 mA cm−2 cm−2 after 10000 cycles and Zn2+
β-MnO2 [83] 1.0–1.8 100 mAh g−1 at 100 mAh g−1 at 1 A g−1 Co-insertion of H+
0.1 A g−1 after 100 cycles and Zn2+
δ-MnO2 [83] 1.0–1.8 126 mAh g−1 at 96 mAh g−1 at 1 A g−1 Co-insertion of H+
0.1 A g−1 after 100 cycles and Zn2+
MnO2 [67] 1.0–1.8 290 mAh g−1 at ~50 mAh g−1 at 6.5 C Co-insertion of H+
0.3 C after 10000 cycles and Zn2+
α-K0.19MnO2 [25] 0.8–1.9 270 mAh g−1 at 1 180 mAh g−1 at 5 C after Co-insertion of H+
C 400 cycles and Zn2+
K0.8Mn8O16 [84] 1.0–1.8 Over 300 mAh 154 mAh g−1 at 1 A g−1 Co-insertion of H+
g−1 at 0.1 A g−1 after 1000 cycles and Zn2+
La3+-δ-MnO2 [54] 0.8–1.9 278.5 mAh g−1 at 71% at 0.2 A g−1 after Co-insertion of H+
0.1 A g−1 200 cycles and Zn2+
Na0.55Mn2O4·0.57H2O 0.8–1.9 389.8 mAh g−1 at 201.6 mAh g−1 at 0.5 A Co-insertion of H+
[63] 0.2 A g−1 g−1 after 400 cycles and Zn2+
PANI-MnO2 [64] 1.0–1.8 298 mAh g−1 at 280 mAh g−1 at 0.2 A Co-insertion of H+
0.05 A g−1 g−1 after 200 cycles and Zn2+
MnO2H0.16(H2O)0.27 0.8–1.8 275.6 mAh g−1 at 131.6 mAh g−1 at 6 C Co-insertion of H+
[71] 0.1 C after 2000 cycles and Zn2+
birnessite-type MnO2 1.0–1.9 279.7 mAh g−1 at 279.7 mAh g−1 at 0.3 A Conversion reaction
[79] 0.3 A g−1 g−1 after 300 cycles
α-MnO2 [77] 1.0–1.9 285 mAh g−1 at 92% at 5 C after 5000 Conversion reaction
C/3 cycles
MnO2 [80] 0.8–2.0 570 mAh g−1 at 2 92% at 30 mA cm−2 after Conversion reaction
mA cm−2 1800 cycles

Unlike the Mn-based materials, the vanadium-based compounds rarely exhibited


cationic conversion reactions, but some anionic redox reactions reported in recent
literatures. For instance, Zhou et al. reported a reversible Zn/VNxOy battery enabled
by the Zn2+ insertion/extraction and the pseudo-capacitive surface redox reaction
(N3 ― ↔N2 ― ) [85]. Due to the rapid anionic redox reaction in the discharge process,
the battery showed unparalleled rate capability of 150 mAh g−1 at the current density
of 20 A g−1 after 2000 cycles. Another anionic reaction in ZIB was reported by Niu et
al. in the Zn/VOPO4 system, which exhibited highly reversible oxygen redox
chemistry in addition to the traditional intercalation reaction [37]. The water-in-salt
electrolyte was selected in the battery enabled a broader voltage window, activating
the oxygen redox reaction in the high voltage of ~1.86 V (Fig. 6e). In charging
process, after the vanadium was thoroughly oxidized, an electron could be lost from
oxygen to form a local charge hole as the voltage continues increasing (Fig. 6f, g).
The introduced anionic redox reaction not only provided extra capacity and increased
the average work potential, but also enhanced the reversible crystal structure
evolution of VOPO4, guaranteeing the superior rate capability and long-term cycle
life.
In summary, apart from the analysis of cathode materials, the operation voltage
and the electrolyte circumstance (including ion concentration, pH, additive, and so on)
are crucial for the exploration of the energy storage mechanism of ZIBs. Although the
conversion reactions possess higher theoretical capacity, how to guarantee the cycling
durability and high coulombic efficiency is a vital issue toward practical application
which is needed to further address.
2.4. Coordination reaction of Zn2+ with organic cathode
Different from the common layered, tunnel-type or 3D open framework
materials that possess conductive ion intercalative crystal structure, some organic
polymers with abundant bond sites for Zn2+ have been as well developed as the
promising cathode material candidates for ZIBs. The organic cathodes store Zn2+
principally via coordination reactions, even though the materials are of great variety
in chemical constitutions.
Owing to the fascinating tunable morphology, abundant active sites and
multifunctionality, metal-organic frameworks (MOFs), especially conductive MOFs,
have been recently attracted increasingly research attention in energy storage fields
[86]. The application of MOFs toward ZIBs has been developed in some advanced
investigations [87–89]. For example, Stoddart et al. utilized a conductive 2D MOF
Cu3(HHTP)2, that possesses open-framework structures similarly to Prussian blue
analogs, as a high-performance cathode for aqueous rechargeable ZIB [87]. Both the
quinoid structure and the copper in the MOF are demonstrated as the redox center via
experimental analysis as well as theoretical computation, delivering 2.3 electrons for
per Cu3(HHTP)2 as shown in Fig. 7(a). As a consequence, Cu3(HHTP)2 cathode
exhibited rapid diffusion rate of Zn2+, low interface resistance and high reversible
capacity of 228 mAh g−1 at 0.05 A g−1. A vanadium-based MOF (MIL-47, V-MOF)
was also developed as a promising cathode material for ZIBs [88]. The high b value
(approximately equal 1.0) and high capacitive contribution from the kinetic analysis
of the V-MOF/Zn battery revealed the battery reaction was dominated by a
surface-controlled capacitive process, which was contributed to the outstanding rate
capability. The reversible peak shift in the XRD patterns in different dis/charged
states suggested an excellent structural reversibility of the V-MOF as the Zn2+
insertion/extraction. More importantly, the XPS analysis of V 2p peaks showed no
discernible changes between the charged and discharged states, implying the organic
moiety, rather than vanadium-ions, provided direct binding sites for the intercalated
Zn2+ in the energy delivery process, as illustrated in Fig. 7(b). Consequently, the Zn2+
storage mechanism of the V-MOF belongs to a solid-solution reaction with excellent
structural stability, where the Zn2+ was reversibly coordinated with the organic
moieties.
Similarly, 2D covalent organic frameworks (COFs), integrated from
pre-designed symmetric organic building units and self-assembled by π–π stacking,
are endowed with tunable structure and well-defined porosity that have been also
explored as potential cathode materials for ZIBs. For instance, the hydroquinone
based HqTp COF, formed from 2,5-diaminohydroquinone dihydrochloride (Hq) and
1,3,5-triformylphloroglucinol (Tp), was demonstrated by Banerjee et al. as an
efficient polymeric cathode material for Zn2+ storage [90]. The C=O and N−H
functional groups in HqTp were considered as the Zn2+ coordinated sites, in which the
C=O···Zn and Zn···N−H interactions were reversibly formed/broken in the processes
of the zinc ion insertion/extraction, accompanied by the energy release/storage (Fig.
7c). Benefitting from the unique porous structure (pore size of 1.5 nm) in HqTp along
with abundant active sites, the COF cathode exhibited a remarkable capacity of 276
mAh g−1 at the current density of 125 mA g−1 under the operating voltage range of
0.2–1.6 V.
Fig. 7. Coordination reactions of Zn2+ with organic cathodes. (a) Structure of 2D
conductive MOF Cu3(HHTP)2 and the reaction process in the coordination unit.
Reproduced with permission [87]. Copyright 2019, Nature Publishing Group. (b)
Schematic of probable Zn2+ coordinated sites with V-MOF. Reproduced with
permission [88]. Copyright 2019, Elsevier. (c) Structure of the COF HqTp and the
Zn2+ inserted model. Reproduced with permission [90]. Copyright 2019, Royal
Chemical Society. (d) Diagram of redox reaction in a quinone-based cathode (PQ-Δ)
for aqueous ZIB. Reproduced with permission [92]. Copyright 2020, American
Chemical Society.
Quinone compounds are attracted impressive attention in ZIBs [91,92]. Chen et
al. investigated various quinone electrodes, in which especially calyx [4] quinone
(C4Q) exhibited superior Zn2+ storage performances [91]. It was demonstrated that
the zinc ions could be coordinated to the electronegative oxygen atoms in the quinone
compounds. Stoddart and co-workers reported another quinone-based polymer,
phenanthrenequinone triangles (PQ-Δ), as an efficient cathode material for aqueous
zinc battery [92]. The Zn/PQ-Δ battery could deliver a capacity of 203 mAh g−1 at the
low current density, where the experimental specific capacity value corresponded to a
six-electrons transfer in per PQ-Δ molecule, as shown in Fig. 7(d). It revealed the
PQ-Δ was fully reduced to the dianionic state (PQ-Δ6−) in the discharge process,
which demonstrated the favorable coordination between divalent zinc ion and the
dione functional groups. It was also found that the water molecules in hydrated Zn2+
can substantially decrease the interfacial resistance by reducing the desolvation
energy and coulombic repulsion. Furthermore, the robust triangular structure of PQ-Δ
endowed the battery with a long-term cycling stability with no capacity deterioration
over 500 cycles.
In addition, a number of other organic polymers have got research attention in
ZIBs. The pyrene-4,5,9,10-tetraone (PTO) is a representative cathode material for
ZIBs, in which Zn2+ is stored via coordinating with adjacent C=O groups [93]. Cao
and co-workers proposed a poly(3,4,9,10-perylentetracarboxylic
dianhydride)/graphene aerogel (PPTCDA/GA) composites as the high-efficiency and
robust cathode, which can deliver a high capacity of 281 mAh g−1 at 0.1 A g−1 without
any decline after 300 cycles operated at such a low current density [94]. The charge
storage mechanism of PPTCDA was demonstrated to originate from the coordination
reaction between Zn2+ and neighboring enolate groups. Liu et al. introduced
poly(4,4’-thiodiphenol, TDP) integrated with high-conductive activated carbon as the
cathode material for Zn2+/H+ co-insertion energy storage device [95]. As the effective
work material, poly(4,4’-TDP) provided plenty of active sites for capturing Zn2+ and
H+. Concretely, zinc ions would coordinate with two neighboring ― O ― , while the
proton would be bonded to the isolated =O forming −OH in the discharge process.
Similar to other Zn2+/H+ synergistic intercalative batteries, zinc hydroxide sulfate
(ZHS) was also found to be formed in the Zn/poly(4,4’-TDP) system caused by the
precipitation of residual OH- with Zn2+. Recently, Niu et al. reported a novel aqueous
Zn/diquinoxalino [2,3-a:2’,3’-c] phenazine (Zn/HATN) battery, which was enabled
by the reversible H+ uptake/removal behavior via the formation/break of the
coordination between the electroactive pyrazine N atoms and H+ [96]. Even though
the Zn/HATN battery cannot belong to zinc-ion battery, the unique proton insertion
chemistry with the organic polymer would provide new sight to construct
high-performance ZIBs.
To date, the organic polymeric cathode materials have occupied increasingly
important place in ZIBs field due to the highly tunable structures and functionalities.
The mechanisms of these materials are principally enabled by the coordination
reaction with Zn2+ in specific electronegative sites, such as quinoid structures,
adjacent C=O sites, N−H functional groups. By adjusting the distance of the
coordinatable groups to fit the size of zinc ions, the polymers would be developed or
modified as the feasible cathode materials for ZIBs [97]. Moreover, the H+ with lower
coulombic repulsion and higher binding force with organic moieties would be likely
to insert into the cathode to deliver extra capacity during discharging, which should be
raised special concern in designing polymeric cathode materials. However, one is
urgent to address that the organic matter is prone to undergo dissolution or structural
degradation, resulting in the deterioration of the energy storage performances.

Fig. 8. Specific capacity of various cathode materials with different energy storage
mechanisms.
3. Summary and outlook
In summary, we reviewed the recent progress in the exploration of the energy
storage mechanism for aqueous rechargeable ZIBs. For the first time, the mechanism
of ZIBs are summarized in four types, i.e. traditional Zn2+ insertion chemistry, dual
ions co-insertion, chemical conversion reaction and coordination reaction of Zn2+ with
organic cathodes. A visualized summary of battery capacities with different energy
storage mechanisms based on the state-of-the-art cathode materials is shown in Fig. 8,
which reveals that the specific capacity of ZIBs depends on both the cathode material
and working mechanism. Therefore, designing proper electrode materials integrated
with advanced energy storage mechanism would be a promising approach to
developing high-performance rechargeable aqueous batteries to meet the requirements
of practical application. In addition, according to the author’s best knowledge and
understanding, we propose some perspectives toward mechanism exploration and
development of high-performance ZIBs.
3.1. Comprehensive exploration of mechanism
Although the current studies demonstrated some self-consistent theories to
explain relative electrochemical behaviors, the controversial actuality in mechanisms
of ZIBs requires further efforts to excavate and reach a consensus. For this purpose,
some comprehensive investigations are needed. First, operando techniques are
considered as powerful approach to get insight of the reaction process. Apart from the
frequently-used in-situ XRD or Raman, some other advanced techniques such as
in-situ TEM, in-situ SEM would be helpful to realize the material evolution during
charging and discharging. Second, study on the reaction mechanism should combine
all part of the battery system, including electrolyte environment, analysis both of
cathode and anode. The battery reactions usually belong to a kind of solid-liquid
reaction, thereby the interface would play an important role, which should be drawn
more attention. Lastly, DFT calculation could provide required theoretical supporting
to unlock the interaction mechanism in ZIBs. The diffusion pathways of Zn2+ in
crystalline structures of cathode and the corresponding diffusion energy barriers can
be simulated by DFT calculation, which would be helpful to predicting and
demonstrating the potential cathode materials. Besides, the Gibbs energies of ion
adsorption could evaluate the Zn2+ adsorption kinetics in the interfacial process.
3.2. Strategy for designing high-performance cathodes
Mechanism innovation would be a gold key to open up a new stage of ZIBs. In
principle, limitation is exited for the specific capacity of batteries based on
intercalation chemistry where the host material would reach saturated when all
accessible sites were inserted by ions. Fortunately, the discovery of conversion
reactions broadens the potential of ZIBs, such as cationic conversion reaction in
Zn/MnO2 system, which broke through the traditional limitation of the host site
density for Zn2+. However, the cationic conversion reactions have not yet found in
vanadium-based compounds cathodes. As an important material with multiple valent
state and abundant phases that possesses the potential to achieve the reaction,
vanadium-based cathodes would scale new heights in performances if introduced in
the cationic conversion mechanism. Notably, further study on conversion reactions
should pay more attention to the inferior reversibility and deficient cycling stability.
Optimization and modification for the present cathode materials can meet the
needs of the practical energy storage devices. The efforts should be made to restrain
the dissolution or unexpected phase changes for manganese oxides to promote the
reversibility and cycling stability [9,107,108]. In spite of the remarkable capacity and
cycling durability for vanadium oxides in ZIBs, the low discharge voltage plateaus
obstruct the steps toward practical applications, which is a burning question
demanded to settle [26,109,110]. How to rationally synthesize the composite
materials from one and another (for instance, RGO/VO2,
Zn3(OH)2V2O7·2H2O@CNT, V2O5@PEDOT) would be a complex task to create
synergistic effect achieving high-performance ZIBs [25,100,101]. Moreover,
nanoscale materials can improve the utilization of the active sites, thereby the
nano-structural materials are regarded as the promising cathode for ZIBs, where the
structure-function relationship should be paid more research efforts [101,102].
Seeking for new proper cathode materials is still a significant but uphill task. In
addition to the common cathode of Mn-, V-based compounds, Prussian blue
analogues or organic polymers, other novel materials have been also developed as
potential cathode candidates, such as Bi2S3, MoO3, MoS2, Mo2N [6,111–114].
Although these cathodes exhibited remarkable stability or rate capability, the inferior
capacity is seemed to be the main obstacle to competing with other materials.
According to the current reports, layered or tunnel materials, which possess large
diffusion pathways and enable rapid diffusion kinetics for Zn2+, are considered as
ideal search target for zinc-ion storage. Besides, the environmental friendliness and
the inexpensiveness are as well crucial for commercial application in the future.
3.3. Constructing stable Zinc metal anode
The practical application of ZIBs usually suffers from the unsatisfying zinc metal
anode, in which the notorious Zn dendrites are prone to growth due to the
unhomogeneous plating/stripping in the repeated charge/discharge process [15,115].
The growth of Zn dendrites could lead to the increase of surface area, intensifying the
consumption of electrolyte, facilitating the anode corrosion, competitive hydrogen
evolution and low efficiency, even puncturing separators and breaking batteries down
[116]. Therefore, the implementation of ZIBs requires not only high-performance
cathodes, but also stable dendrite-free Zn anodes. To inhibit the formation of Zn
dendrites, the measures are mainly taken in the following aspects currently: (i)
Modifying the anode substrate for homogeneous zinc nucleation; (ii) constructing a
protective artificial solid electrolyte interface (SEI) on the Zn anode; (iii) optimizing
the electrolyte by adjusting additives to guide zinc plating/stripping [19]. Although
massive efforts have been devoted, developing dendrite-free zinc anode with high
efficiency is still a significant challenge. Notably, apart from the fundamental
properties in laboratory study, the future research should be paid more attention to
industrial-grade index (e.g. the mass ratio between cathode and anode, performances
in large materials-loading, durability against high/low temperature), which is crucial
for its commercial application.
3.4. Prospects toward practical application of ZIBs
Compared with LIBs that dominate the current commercial energy storage
markets, the reliable safety and high-power density endow the aqueous rechargeable
ZIBs with unparalleled application potential in flexible smart electronic devices.
Indeed, a number of flexible solid-state zinc-ion batteries or fiber-shape batteries have
demonstrated out the excellent performances, especially when integrated with other
sensors to assemble as self-powered wearable devices [34,117–120]. The
as-assembled solid-state ZIBs even exhibited superb stability against cutting, soaking
in water, combusting, bending, hammering, drilling or other extreme conditions [121–
123]. The core problem is to develop high-performance solid-state electrolyte with
unimpeded Zn2+ transport that can couple with self-standing electrodes. In addition,
constructing anti-freezing hydrogel electrolyte can further broaden the working
temperature range to enable ZIBs in extremely cold environment [124,125].
The new functionalities can be unlocked by introducing some special properties
for the aqueous rechargeable ZIBs [126–130]. For example, the aqueous zinc-ion
storage system incorporated with transparent battery architectures would construct an
electrochromic battery, which enables a lot of new applications, including variable
optical attenuators, energy-efficient smart windows, addressable displays, and optical
switches [128]. Therefore, there will be wider prospects for ZIBs in the near-future
applications.
All in all, the energy storage mechanisms of aqueous rechargeable ZIBs were
comprehensively reviewed, which we believe would provide helpful references for
the next studies. However, disputed mechanism in some specific issues still keeps
confusing, requiring further systematic and scientific study to reach a consensus. In
conclusion, considering the unparalleled security, low-cost and eco-friendliness, the
aqueous rechargeable ZIBs would be soon scaled up in commercial applications with
the continue research and optimization.
Acknowledgments
This work was supported by the National Natural Science Foundation of China
(21571080).
References
[1] B. Tang, L. Shan, S. Liang, J. Zhou, Energy Environ. Sci. 12 (2019) 3288-3304.
[2] J. Ming, J. Guo, C. Xia, W. Wang, H. N. Alshareef, Mater. Sci. Eng. R 135 (2019)
58-84.
[3] D. Kundu, B. D. Adams, V. Duffort, S. H. Vajargah, L. F. Nazar, Nat. Energy 1
(2016) 16119.
[4] P. He, M. Yan, G. Zhang, R. Sun, L. Chen, Q. An, L. Mai, Adv. Energy Mater. 7
(2017) 1601920.
[5] G. Wang, B. Kohn, U. Scheler, F. Wang, S. Oswald, M. Loffler, D. Tan, P. Zhang,
J. Zhang, X. Feng, Adv. Mater. 32 (2020) e1905681.
[6] H. Liang, Z. Cao, F. Ming, W. Zhang, D. H. Anjum, Y. Cui, L. Cavallo, H. N.
Alshareef, Nano Lett. 19 (2019) 3199-3206.
[7] D. Chao, W. Zhou, F. Xie, C. Ye, H. Li, M. Jaroniec, S.-Z. Qiao, Sci. Adv. 6
(2020) eaba4098.
[8] M. Winter, R. J. Brodd, Chem. Rev. 104 (2004) 4245-4270.
[9] F. Wan, Z. Niu, Angew. Chem. Int. Ed. 58 (2019) 16358-16367.
[10] N. Zhang, Y. Dong, M. Jia, X. Bian, Y. Wang, M. Qiu, J. Xu, Y. Liu, L. Jiao, F.
Cheng, ACS Energy Lett. 3 (2018) 1366-1372.
[11] M. Song, H. Tan, D. Chao, H. J. Fan, Adv. Funct. Mater. 28 (2018) 1802564.
[12] J. F. Parker, C. N. Chervin, I. R. Pala, M. Machler, M. F. Burz, J. W. Long, D. R.
Rolison, Science 356 (2017) 415–418.
[13] W. Xu, Y. Wang, Nano-Micro Lett. 11 (2019) 90.
[14] X. Liu, J. Yi, K. Wu, Y. Jiang, Y. Liu, B. Zhao, W. Li, J. Zhang, Nanotechnol.
31 (2020) 122001.
[15] J. Guo, J. Ming, Y. Lei, W. Zhang, C. Xia, Y. Cui, H. N. Alshareef, ACS Energy
Lett. 4 (2019) 2776-2781.
[16] H. Li, L. Ma, C. Han, Z. Wang, Z. Liu, Z. Tang, C. Zhi, Nano Energy 62 (2019)
550-587.
[17] G. Fang, J. Zhou, A. Pan, S. Liang, ACS Energy Lett. 3 (2018) 2480-2501.
[18] J. Zheng, Q. Zhao, T. Tang, J. Yin, C. D. Quilty, G. D. Renderos, X. Liu, Y.
Deng, L. Wang, D. C. Bock, C. Jaye, D. Zhang, E. S. Takeuchi, K. J. Takeuchi, A.
C. Marschilok, L. A. Archer, Science 366 (2019) 645–648.
[19] H. Jia, Z. Wang, B. Tawiah, Y. Wang, C.-Y. Chan, B. Fei, F. Pan, Nano Energy
70 (2020) 104523.
[20] Y. Zhao, Y. Zhu, X. Zhang, InfoMat 2, (2020) 237-260.
[21] D. Selvakumaran, A. Pan, S. Liang, G. Cao, J. Mater. Chem. A 7 (2019)
18209-18236.
[22] C. Li, X. Zhang, W. He, G. Xu, R. Sun, J. Power Sources 449 (2020) 227596.
[23] C. Xu, B. Li, H. Du, F. Kang, Angew. Chem. Int. Ed. 51 (2012) 933-935.
[24] B. Lee, C. S. Yoon, H. R. Lee, K. Y. Chung, B. W. Cho, S. H. Oh, Sci. Rep. 4
(2014) 6066.
[25] G. Liu, H. Huang, R. Bi, X. Xiao, T. Ma, L. Zhang, J. Mater. Chem. A 7 (2019)
20806-20812.
[26] X. Dai, F. Wan, L. Zhang, H. Cao, Z. Niu, Energy Storage Mater. 17 (2019)
143-150.
[27] J. Ding, Z. Du, L. Gu, B. Li, L. Wang, S. Wang, Y. Gong, S. Yang, Adv. Mater.
30 (2018) e1800762.
[28] L. Chen, Y. Ruan, G. Zhang, Q. Wei, Y. Jiang, T. Xiong, P. He, W. Yang, M.
Yan, Q. An, L. Mai, Chem. Mater. 31 (2019) 699-706.
[29] M. H. Alfaruqi, J. Gim, S. Kim, J. Song, D. T. Pham, J. Jo, Z. Xiu, V. Mathew, J.
Kim, Electrochem. Commun. 60 (2015) 121-125.
[30] Y. Jiang, D. Ba, Y. Li, J. Liu, Adv. Sci. 7 (2020) 1902795.
[31] J. Zhou, L. Shan, Z. Wu, X. Guo, G. Fang, S. Liang, Chem. Commun. 54 (2018)
4457-4460.
[32] Z. Cao, H. Chu, H. Zhang, Y. Ge, R. Clemente, P. Dong, L. Wang, J. Shen, M.
Ye, P. M. Ajayan, J. Mater. Chem. A 7 (2019) 25262-25267.
[33] P. He, Y. Quan, X. Xu, M. Yan, W. Yang, Q. An, L. He, L. Mai, Small 13 (2017)
1702551.
[34] D. Chao, C. R. Zhu, M. Song, P. Liang, X. Zhang, N. H. Tiep, H. Zhao, J. Wang,
R. Wang, H. Zhang, H. J. Fan, Adv. Mater. 30 (2018) e1803181.
[35] N. Zhang, M. Jia, Y. Dong, Y. Wang, J. Xu, Y. Liu, L. Jiao, F. Cheng, Adv.
Funct. Mater. 29 (2019) 1807331.
[36] M. Liao, J. Wang, L. Ye, H. Sun, Y. Wen, C. Wang, X. Sun, B. Wang, H. Peng,
Angew. Chem. Int. Ed. 59 (2020) 2273 –2278.
[37] F. Wan, Y. Zhang, L. Zhang, D. Liu, C. Wang, L. Song, Z. Niu, J. Chen, Angew.
Chem. Int. Ed. 58 (2019) 7062-7067.
[38] P. Hu, T. Zhu, X. Wang, X. Zhou, X. Wei, X. Yao, W. Luo, C. Shi, K. A.
Owusu, L. Zhou, L. Mai, Nano Energy 58 (2019) 492-498.
[39] L. Zhang, L. Chen, X. Zhou, Z. Liu, Adv. Energy Mater. 5 (2015) 1400930.
[40] L. Ma, S. Chen, C. Long, X. Li, Y. Zhao, Z. Liu, Z. Huang, B. Dong, J. A.
Zapien, C. Zhi, Adv. Energy Mater. 9 (2019) 1902446.
[41] P. Hu, Z. Zou, X. Sun, D. Wang, J. Ma, Q. Kong, D. Xiao, L. Gu, X. Zhou, J.
Zhao, S. Dong, B. He, M. Avdeev, S. Shi, G. Cui, L. Chen, Adv. Mater. 32 (2020)
e1907526.
[42] M. Yan, P. He, Y. Chen, S. Wang, Q. Wei, K. Zhao, X. Xu, Q. An, Y. Shuang,
Y. Shao, K. T. Mueller, L. Mai, J. Liu, J. Yang, Adv. Mater. 30 (2018) 1703725.
[43] J. Shin, D. S. Choi, H. J. Lee, Y. Jung, J. W. Choi, Adv. Energy Mater. 9 (2019)
1900083.
[44] K. Zhu, T. Wu, K. Huang, Adv. Energy Mater. 9 (2019) 1901968.
[45] K. Zhu, T. Wu, K. Huang, ACS nano 13 (2019) 14447-14458.
[46] Y. Yang, Y. Tang, G. Fang, L. Shan, J. Guo, W. Zhang, C. Wang, L. Wang, J.
Zhou, S. Liang, Energy Environ. Sci. 11 (2018) 3157-3162.
[47] S. Islam, M. H. Alfaruqi, D. Y. Putro, V. Soundharrajan, B. Sambandam, J. Jo, S.
Park, S. Lee, V. Mathew, J. Kim, J. Mater. Chem. A 7 (2019) 20335-20347.
[48] Y. Yang, Y. Tang, S. Liang, Z. Wu, G. Fang, X. Cao, C. Wang, T. Lin, A. Pan, J.
Zhou, Nano Energy 61 (2019) 617-625.
[49] P. He, G. Zhang, X. Liao, M. Yan, X. Xu, Q. An, J. Liu, L. Mai, Adv. Energy
Mater. 8 (2018) 1702463.
[50] H. Geng, M. Cheng, B. Wang, Y. Yang, Y. Zhang, C. C. Li, Adv. Funct. Mater.
30 (2019) 1907684.
[51] L. Ma, N. Li, C. Long, B. Dong, D. Fang, Z. Liu, Y. Zhao, X. Li, J. Fan, S. Chen,
S. Zhang, C. Zhi, Adv. Funct. Mater. 29 (2019) 1906142.
[52] L. Shan, Y. Yang, W. Zhang, H. Chen, G. Fang, J. Zhou, S. Liang, Energy
Storage Mater. 18 (2019) 10-14.
[53] F. Ming, H. Liang, Y. Lei, S. Kandambeth, M. Eddaoudi, H. N. Alshareef, ACS
Energy Lett. 3 (2018) 2602-2609.
[54] H. Zhang, Q. Liu, J. Wang, K. Chen, D. Xue, J. Liu, X. Lu, J. Mater. Chem. A 7
(2019) 22079-22083.
[55] W. Zhang, C. Tang, B. Lan, L. Chen, W. Tang, C. Zuo, S. Dong, Q. An, P. Luo,
J. Alloys Compd. 819 (2020) 152971.
[56] C. Xia, J. Guo, P. Li, X. Zhang, H. N. Alshareef, Angew. Chem. Int. Ed. 57
(2018) 3943-3948.
[57] H. Liu, J. G. Wang, H. Sun, Y. Li, J. Yang, C. Wei, F. Kang, J. Colloid Interface
Sci. 560 (2020) 659-666.
[58] B. Lan, Z. Peng, L. Chen, C. Tang, S. Dong, C. Chen, M. Zhou, C. Chen, Q. An,
P. Luo, J. Alloys Compd. 787 (2019) 9-16.
[59] Y. Liu, Q. Li, K. Ma, G. Yang, C. Wang, ACS Nano 13 (2019) 12081-12089.
[60] S. Chen, Y. Zhang, H. Geng, Y. Yang, X. Rui, C. C. Li, J. Power Sources 441
(2019) 227192.
[61] F. Liu, Z. Chen, G. Fang, Z. Wang, Y. Cai, B. Tang, J. Zhou, S. Liang,
Nano-Micro Lett. 11 (2019) 25.
[62] K. Sada, B. Senthilkumar, P. Barpanda, J. Mater. Chem. A 7 (2019)
23981-23988.
[63] X.-Z. Zhai, J. Qu, S.-M. Hao, Y.-Q. Jing, W. Chang, J. Wang, W. Li, Y.
Abdelkrim, H. Yuan, Z.-Z. Yu, Nano-Micro Lett. 12 (2020) 56.
[64] J. Huang, Z. Wang, M. Hou, X. Dong, Y. Liu, Y. Wang, Y. Xia, Nat. Commun. 9
(2018) 2906.
[65] D. Bin, W. Huo, Y. Yuan, J. Huang, Y. Liu, Y. Zhang, F. Dong, Y. Wang, Y.
Xia, Chem 6 (2020) 968-984.
[66] S. Liu, H. Zhu, B. Zhang, G. Li, H. Zhu, Y. Ren, H. Geng, Y. Yang, Q. Liu, C.C.
Li, Adv. Mater. (2020) e2001113. DOI: 10.1002/adma.202001113.
[67] W. Sun, F. Wang, S. Hou, C. Yang, X. Fan, Z. Ma, T. Gao, F. Han, R. Hu, M.
Zhu, C. Wang, J. Am. Chem. Soc. 139 (2017) 9775-9778.
[68] F. Wan, L. Zhang, X. Dai, X. Wang, Z. Niu, J. Chen, Nat. Commun. 9 (2018)
1656.
[69] W. Liu, L. Dong, B. Jiang, Y. Huang, X. Wang, C. Xu, Z. Kang, J. Mou, F.
Kang, Electrochim. Acta 320 (2019) 134565.
[70] X. Gao, H. Wu, W. Li, Y. Tian, Y. Zhang, H. Wu, L. Yang, G. Zou, H. Hou, X.
Ji, Small 16 (2020) e1905842.
[71] Q. Zhao, X. Chen, Z. Wang, L. Yang, R. Qin, J. Yang, Y. Song, S. Ding, M.
Weng, W. Huang, J. Liu, W. Zhao, G. Qian, K. Yang, Y. Cui, H. Chen, F. Pan,
Small 15 (2019) e1904545.
[72] Q. Yang, F. Mo, Z. Liu, L. Ma, X. Li, D. Fang, S. Chen, S. Zhang, C. Zhi, Adv.
Mater. 31 (2019) e1901521.
[73] C. Li, J. Wu, F. Ma, Y. Chen, L. Fu, Y. Zhu, Y. Zhang, P. Wang, Y. Wu, W.
Huang, ACS Appl. Energy Mater. 2 (2019) 6984-6989.
[74] P. Gao, Q. Ru, H. Yan, S. Cheng, Y. Liu, X. Hou, L. Wei, F. Chi‐Chung Ling,
ChemElectroChem 7 (2020) 283-288.
[75] L. Wang, K. Huang, J. Chen, J. Zheng, Sci. Adv. 5 (2019) eaax4279.
[76] D. Batyrbekuly, S. Cajoly, B. Laik, J. P. Pereira-Ramos, N. Emery, Z. Bakenov,
R. Baddour-Hadjean, ChemSusChem 13 (2019) 724-731.
[77] H. Pan, Y. Shao, P. Yan, Y. Cheng, K. S. Han, Z. Nie, C. Wang, J. Yang, X. Li,
P. Bhattacharya, K. T. Mueller, J. Liu, Nat. Energy 1 (2016) 16039.
[78] B. Lee, H. Seo, H. Lee, C. Yoon, J. Kim, K. Chung, B. Cho, S. H. Oh,
ChemSusChem 9 (2016) 2948–2956.
[79] G. Li, Z. Huang, J. Chen, F. Yao, J. Liu, O. L. Li, S. Sun, Z. Shi, J. Mater. Chem.
A 8 (2020) 1975-1985.
[80] D. Chao, W. Zhou, C. Ye, Q. Zhang, Y. Chen, L. Gu, K. Davey, S. Z. Qiao,
Angew. Chem. Int. Ed. 58 (2019) 7823-7828.
[81] D. Chao, C. Ye, F. Xie, W. Zhou, Q. Zhang, Q. Gu, K. Davey, L. Gu, S.Z. Qiao,
Adv. Mater. (2020) 2001894. DOI: 10.1002/adma.202001894.
[82] B. Jiang, C. Xu, C. Wu, L. Dong, J. Li, F. Kang, Electrochim. Acta 229 (2017)
422-428.
[83] C. Guo, Q. Zhou, H. Liu, S. Tian, B. Chen, J. Zhao, J. Li, Electrochim. Acta, 324
(2019) 134867.
[84] G. Fang, C. Zhu, M. Chen, J. Zhou, B. Tang, X. Cao, X. Zheng, A. Pan, S. Liang,
Adv. Funct. Mater. 29 (2019) 1808375.
[85] G. Fang, S. Liang, Z. Chen, P. Cui, X. Zheng, A. Pan, B. Lu, X. Lu, J. Zhou,
Adv. Funct. Mater. 29 (2019) 1905267.
[86] D. Cai, M. Lu, Li, J. Cao, D. Chen, H. Tu, J. Li, W. Han, Small 15 (2019)
e1902605.
[87] K. W. Nam, S. S. Park, R. Dos Reis, V. P. Dravid, H. Kim, C. A. Mirkin, J. F.
Stoddart, Nat. Commun. 10 (2019) 4948.
[88] B. He, Q. Zhang, P. Man, Z. Zhou, C. Li, Q. Li, L. Xie, X. Wang, H. Pang, Y.
Yao, Nano Energy 64 (2019) 103935.
[89] Y. Ding, Y. Peng, W. Chen, Y. Niu, S. Wu, X. Zhang, L. Hu, Appl. Surf. Sci.
493 (2019) 368-374.
[90] M. A. Khayum, M. Ghosh, V. Vijayakumar, A. Halder, M. Nurhuda, S. Kumar,
M. Addicoat, S. Kurungot, R. Banerjee, Chem. Sci. 10 (2019) 8889-8894.
[91] Q. Zhao, W. Huang, Z. Luo, L. Liu, Y. Lu, Y. Li, L. Li, J. Hu, H. Ma, J. Chen,
Sci. Adv. 4 (2018) eaao1761.
[92] K. W. Nam, H. Kim, Y. Beldjoudi, T. W. Kwon, D. J. Kim, J. F. Stoddart, J. Am.
Chem. Soc. 142 (2020) 2541-2548.
[93] Z. Guo, Y. Ma, X. Dong, J. Huang, Y. Wang, Y. Xia, Angew. Chem. Int. Ed. 57
(2018) 11737-11741.
[94] R. Cang, K. Ye, K. Zhu, J. Yan, J. Yin, K. Cheng, G. Wang, D. Cao, J. Energy
Chem. 45 (2020) 52-58.
[95] T. Xin, Y. Wang, N. Wang, Y. Zhao, H. Li, Z. Zhang, J. Liu, J. Mater. Chem. A
7 (2019) 23076-23083.
[96] Z. Tie, L. Liu, S. Deng, D. Zhao, Z. Niu, Angew. Chem. Int. Ed. 59 (2020)
4920-4924.
[97] Y. Wang, Q. Liu, Q. Zhang, B. Peng, H. Deng, Angew. Chem. Int. Ed. 57 (2018)
7120-7125.
[98] P. Hu, M. Yan, T. Zhu, X. Wang, X. Wei, J. Li, L. Zhou, Z. Li, L. Chen, L. Mai,
ACS Appl. Mater. Interfaces 9 (2017) 42717-42722.
[99] X. Wang, L. Ma, J. Sun, ACS Appl. Mater. Interfaces 11 (2019) 41297-41303.
[100] J. Ding, Z. Du, B. Li, L. Wang, S. Wang, Y. Gong, S. Yang, Adv. Mater. 31
(2019) e1904369.
[101] P. Hu, T. Zhu, J. Ma, C. Cai, G. Hu, X. Wang, Z. Liu, L. Zhou, L. Mai, Chem.
Commun. 55 (2019) 8486-8489.
[102] H. Qin, L. Chen, L. Wang, X. Chen, Z. Yang, Electrochim. Acta, 306 (2019)
307-316.
[103] C. Liu, Z. Neale, J. Zheng, X. Jia, J. Huang, M. Yan, M. Tian, M. Wang, J.
Yang, G. Cao, Energy Environ. Sci. 12 (2019) 2273-2285.
[104] J. Lai, H. Tang, X. Zhu, Y. Wang, J. Mater. Chem. A 7 (2019) 23140-23148.
[105] L. Gou, K. L. Mou, X. Y. Fan, M. J. Zhao, Y. Wang, D. Xue, D. L. Li, Dalton
Trans. 49 (2020) 711-718.
[106] L. Qian, T. Wei, K. Ma, G. Yang, C. Wang, ACS Appl. Mater. Interfaces 11
(2019) 20888-20894.
[107] Z. Liu, Y. Huang, Y. Huang, Q. Yang, X. Li, Z. Huang, C. Zhi, Chem. Soc.
Rev. 49 (2020) 180-232.
[108] X. Xu, F. Xiong, J. Meng, X. Wang, C. Niu, Q. An, L. Mai, Adv. Funct. Mater.
30 (2020) 1904398.
[109] Z. Pan, J. Yang, J. Yang, Q. Zhang, H. Zhang, X. Li, Z. Kou, Y. Zhang, H.
Chen, C. Yan, J. Wang, ACS Nano 14 (2020) 842-853.
[110] D. Xu, H. Wang, F. Li, Z. Guan, R. Wang, B. He, Y. Gong, X. Hu, Adv. Mater.
Interfaces 6 (2019) 1801506.
[111] T. Xiong, Y. Wang, B. Yin, W. Shi, W. S. V. Lee, J. Xue, Nano-Micro Lett. 12
(2019) 8.
[112] W. Liu, J. Hao, C. Xu, J. Mou, L. Dong, F. Jiang, Z. Kang, J. Wu, B. Jiang, F.
Kang, Chem. Commun. 53 (2017) 6872-6874.
[113] W. Xu, K. Zhao, Y. Wang, Energy Storage Mater. 15 (2018) 374-379.
[114] Y. Liu, J. Wang, Y. Zeng, J. Liu, X. Liu, X. Lu, Small 16 (2020) 1907458.
[115] C. Li, X. Xie, S. Liang, J. Zhou, Energy Environ. Mater. (2020). DOI:
10.1002/eem2.12067.
[116] Q. Zhang, J. Luan, Y. Tang, X. Ji, H. Wang, Angew. Chem. Int. Ed. (2020).
DOI: 10.1002/anie.202000162.
[117] H. Li, Z. Liu, G. Liang, Y. Huang, Y. Huang, M. Zhu, Z. Pei, Q. Xue, Z. Tang,
Y. Wang, B. Li, C. Zhi, ACS Nano 12 (2018) 3140-3148.
[118] H. Li, C. Han, Y. Huang, Y. Huang, M. Zhu, Z. Pei, Q. Xue, Z. Wang, Z. Liu,
Z. Tang, Y. Wang, F. Kang, B. Li, C. Zhi, Energy Environ. Sci. 11 (2018)
941-951.
[119] X. Li, Y. Tang, H. Lv, W. Wang, F. Mo, G. Liang, C. Zhi, H. Li, Nanoscale 11
(2019) 17992-18008.
[120] J. Liu, N. Nie, J. Wang, M. Hu, J. Zhang, M. Li, Y. Huang, Mater. Today
Energy 16 (2020) 100372.
[121] Q. Han, X. Chi, Y. Liu, L. Wang, Y. Du, Y. Ren, Y. Liu, J. Mater. Chem. A 7
(2019) 22287-22295.
[122] Y. Huang, J. Zhang, J. Liu, Z. Li, S. Jin, Z. Li, S. Zhang, H. Zhou, Mater.
Today Energy 14 (2019) 100349.
[123] X. Li, L. Ma, Y. Zhao, Q. Yang, D. Wang, Z. Huang, G. Liang, F. Mo, Z. Liu,
C. Zhi, Mater. Today Energy 14 (2019) 100361.
[124] F. Mo, G. Liang, Q. Meng, Z. Liu, H. Li, J. Fan, C. Zhi, Energy Environ. Sci.
12 (2019) 706-715.
[125] M. Zhu, X. Wang, H. Tang, J. Wang, Q. Hao, L. Liu, Y. Li, K. Zhang, O. G.
Schmidt, Adv. Funct. Mater. 30 (2019) 1907218.
[126] T. Xue, H. Jin Fan, J. Energy Chem. (2020). DOI:
10.1016/j.jechem.2020.05.056.
[127] H. Li, C. J. Firby, A. Y. Elezzabi, Joule 3 (2019) 2268-2278.
[128] H. Li, L. McRae, C. J. Firby, A. Y. Elezzabi, Adv. Mater. 31 (2019) e1807065.
[129] H. Li, A. Y. Elezzabi, Nanoscale Horiz. 5 (2020) 691-695.
[130] L. Zhang, D. Chao, P. Yang, L. Weber, J. Li, T. Kraus, H.J. Fan, Adv. Energy
Mater. 10 (2020) 2000142.
Graphical abstract

A review focused on energy storage mechanism of aqueous zinc-ion batteries (ZIBs)


is present, in which the battery reaction, cathode optimization strategy and underlying
prospect are comprehensively discussed.
Highlights

 The energy storage mechanisms of aqueous ZIBs are systematically

reviewed.
 Battery reactions for ZIBs are firstly summarized in four basic categories.

 Perspectives toward mechanism exploration and development of

high-performance ZIBs are proposed.

You might also like