• Keine Ergebnisse gefunden

A review on recent developments of vanadium‑based cathode for rechargeable zinc‑ion batteries Tungsten

N/A
N/A
Protected

Academic year: 2022

Aktie "A review on recent developments of vanadium‑based cathode for rechargeable zinc‑ion batteries Tungsten"

Copied!
16
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Tungsten (2021) 3:289–304

https://doi.org/10.1007/s42864-021-00091-9 REVIEW PAPER

A review on recent developments of vanadium‑based cathode for rechargeable zinc‑ion batteries

Yan Wu1,2  · Tian‑Yi Song2  · Li‑Na Chen1

Received: 29 December 2020 / Revised: 2 February 2021 / Accepted: 2 February 2021 / Published online: 1 July 2021

© The Nonferrous Metals Society of China 2021

Abstract

Benefiting from their high safety, low cost, and excellent performance, aqueous zinc-ion batteries are regarded as a promis- ing candidate for next-generation commercial energy storage devices. High-performance cathodes are urgently needed to accelerate practical application of zinc-ion batteries (ZIBs). Among various cathodes reported previously, vanadium-based materials attract a great deal of attention since they hold high capacity and good cycling stability. Though fruitful achieve- ments have been made, there are amounts of crystal structures and energy storage mechanisms are still unclear, which will significantly affect performance of full batteries. This review presents a comprehensive overview of structure characteristics, relevant electrochemical behavior, and  proposed energy storage mechanisms of reported vanadium-based materials which provide effective Zn-storage performance. Meanwhile,recent developments of vanadium-based materials for ZIBs are sys- tematically summarized. Last but not least,  the future perspectives are also discussed. We hope that this review can provide suggestions to design high-performance electrode materials and promote the development of ZIBs.

Keywords Aqueous zinc-ion batteries · Vanadium-based materials · Cathode · Structure · Energy storage mechanism

1 Introduction

Considering the limit of traditional energy resources stocks, serious environmental pollution problems, and continuous developments of electric vehicles and portable electronics, it is urgent to explore a reversible energy storage device for renew- able energy with low cost, good safety, as well as remarkable energy density. Currently, the most commonly commercial choice is lithium-ion batteries (LIBs) [1] with superior energy densities and cycle stability, but their large-scale application is strongly impeded by the high cost, low safety, and the toxic substance produced in cycling. This motivates people to explore alternative rechargeable batteries. Zn-ion batteries (ZIBs) [2]

have attracted a lot of attention, because Zn2+ ions have simi- lar ionic radius with Li+ ions (0.075 nm vs. 0.076 nm) [3], indicating the similar energy barriers. Divalent zinc ions have

higher electrostatic interaction when inserting into the electrode materials than that of lithium ions [4]. In addition, zinc metal shows high stability under atmospheric environment and can be used directly as anode materials for ZIBs, enabling low cost [5], low redox potential (− 0.76 V vs standard hydrogen electrode) [6], and high theoretical capacity (≈ 820 mAh·g−1) [7]. Fur- thermore, the mild aqueous electrolytes with high safety, low price, and easy-to-use [8] have an advantage over their organic counterparts [9]. Therefore, there are a lot of works focused on investigating appropriate positive materials for aqueous ZIBs (AZIBs), such as the Mn-based materials [10], V-based mate- rial [11], and Prussian blue analogues (PBAs) [12]. Among them, Mn-based materials and PBAs show high operating volt- age, but their capacity and reversibility during cycling are still not satisfactory [13]. These disadvantages hugely limit their further application of next-generation AZIBs. In this case, V-based materials are regarded as the most promising candi- date, which provide high capacity and excellent cycle stability [14–16]. Crystal structure of the active materials and Zn-ion storage mechanisms have significant effect on the electrochemi- cal behaviors. Various crystal structures of V-based materials have been explored and their energy storage mechanisms are still controversial. Thus, in terms of crystal structure and stor- age mechanism, this review aims to provide a comprehensive

Tungsten

www.springer.com/42864

* Li-Na Chen linachen@hit.edu.cn

1 School of Materials Science and Engineering, Harbin Institute of Technology, Shenzhen 518055, China

2 Department of Chemistry, Center of Super‐Diamond and Advanced Films (COSDAF), City University of Hong Kong, Hong Kong 999077, China

(2)

290 Y. Wu et al.

summary of the developments of V-based materials for AZIBs.

We believe that this review could provide a suggestion in designing high-performance cathode materials and promoting the development of AZIBs.

2 Structure of vanadium‑based materials

Vanadium-based materials mainly contain vanadium oxides and metal vanadate, varying from layered structure, tunnel based structure, other structures, etc. (Fig. 1) [17–19].

2.1 Layered structure 2.1.1 Normal layer

Thanks to their abundant zinc-ion storage sites and favorable ions transfer channels, layered vanadium‐based oxides have been identified as promising candidates for AZIBs. V2O5 consisting of VO6 square pyramids [20] is the most common layered V-based electrode material for AZIBs, displaying

a remarkable theoretical capacity of 589 mAh·g−1 [21].

Layered V2O5 cathode prepared by atomic layer deposition showed a high capacity of 513 mAh·g−1 at 0.5 A·g−1 and reversible capacity of 439 mAh·g−1 at 5 A·g−1 even after 1000 cycles. Interestingly, the electrolyte’s pH value could affect phase transformation during cycling. For instance, no new phase produced but only partial dissolution took place if pH value is below 3.2, but the phase transition occurred when pH values are above 3.8 [10]. Zhang et al. [22] discov- ered that the morphology of V2O5 cathode changed during the electrochemical process. In detail, the commercial V2O5 powder changed into porous nanosheets with more active sites for Zn2+ storage (Fig. 2a).

As shown in Fig. 2b, VS2 is a typical layered structure material and active for insertion/extraction of Zn2+. There is a vanadium layer located in two sulfur layers, providing a large interlayer spacing (0.576 nm) and delivering a high capacity of 190.3 mAh·g−1 at 0.05 A·g−1 [23]. As shown in Fig. 2c, VOPO4 is another typical layer structural cathode, in which VO6 octahedra corner-shared with each other and simultaneously connected with PO4 tetrahedra to form layer

Fig. 1 Crystal structures of typi- cal vanadiumbased electrodes for AZIBs. Crystal structure on the left is reproduced with per- mission from Ref. [17], Copy- right 2019 American Chemical Society (ACS). Crystal structure on the upper right is reproduced with permission from Ref. [18], Copyright 2018 Wiley. Crystal structure on the lower right is reproduced with permission from Ref. [19], Copyright 2020 Royal Society of Chemis- try (RSC)

(3)

structure. The VOPO4/Zn battery extended the voltage win- dow to 2.1 V with the activated oxygen redox because of the water‐in‐salt electrolyte, resulting in a higher average operating potential of 1.56 V [24].

Above all, even normal layer V-based cathodes show many advantages. For example, V2O5 delivers high specific capacity and VOPO4 shows high operating voltage. They still face some problems, such as low ion diffusion, low conductivity, dissolution of materials, self-aggregation for most V2O5 cathodes, relatively low specific capacity and poor cycle stability for VS2 as well as VOPO4.

2.1.2 Intercalated layer

Inspired by the excellent property of intercalated layered materials in other types of rechargeable ion batteries, pre- intercalating different ions and molecules into layered vana- dium-based materials were used to obtain lager space for ions transfer.

For instance, as shown in Fig. 3a, by pre-inserting Na+ ions into the V2O5 layer, the interlayer space can be expanded to 0.501 nm, resulting in an excellent capacity of 367.1 mAh·g−1 at 0.1 A·g−1 and 93% retention for 1000 cycles [25]. Ming et al. [17] synthesized Mg2+-intercalated V2O5

Fig. 2 a Schematic illustration of a rechargeable aqueous Zn-V2O5 battery (left) and galvanostatic cycling performance at 0.2, 0.5, and 1.0 A·g−1 and the corresponding coulombic efficiency at 0.2 A·g−1 (right). Reproduced with permission from Ref. [22]. Copyright 2018 ACS. b Schematic illustration of operation mechanism of Zn-VS2 bat- teries (left) and charge/discharge curves of Zn-VS2 batteries at the

current densityfrom 0.05 to 2.0 A·g−1 (right). Reproduced with per- mission from Ref. [23]. Copyright 2017 Wiley. c Crystal structure of VOPO4 (left) and charge/discharge curves of Zn/VOPO4 batteries at 0.05 A·g−1 (right). Reproduced with permission from Ref. [24]. Cop- yright 2019 Wiley

(4)

292 Y. Wu et al.

with large layer space (1.34 nm) as the cathode material for AZIBs, providing a capacity of 353 mAh·g–1 at 0.1 A·g–1 (as shown in Fig. 3b). Cathode materials composed of anionic [V3O8] with the large interlayer space usually exhibit high specific capacity. Therefore, Cai et al. [26] designed the Na+ pre-inserted V3O8 layers with enhanced interlayer distance, which can store more Zn2+ as host materials in AZIBs. When further mixed with nanoribbons/graphene to form nanocom- posites, they achieved excellent reversibility of 223 mAh·g−1 at 0.3 A·g−1. Cations such as NH4+ also show good results for modifying V2O5 cathode. The obtained (NH4)2V10O25·8H2O exhibits a robust bi-layered structure, which makes the assem- bled cell display high energy density of 341 Wh·kg−1 [27].

Pre-intercalation of a great deal of metal ions may decrease the reversible capacity of V2O5-based materials.

Yang et al. [28] pre-intercalated a bit of transition metal ions, such as Fe2+, Co2+, Ni2+, Mn2+, Zn2+, and Cu2+, into the V2O5 interlayer, and confirmed that this universal strat- egy can significantly improve ions transfer kinetic, structure stability, and temperature adaptability of V2O5. Hydrated layer materials such as V2O5·nH2O demonstrated an energy density of ≈144 Wh·kg−1 at 0.3 A·g−1, thanks to the “lubri- cating” effect from H2O [29].

Ions and water co-insertion is also demonstrated to be a positive way to enlarge the layer space and enabled higher electrochemical performance of layered V2O5. For exam- ple, Yang et al. [30] reported a series of LixV2O5·nH2O with a spacing of 1.377 nm in the (001) face as cathodes for AZIBs, delivering high capacities, excellent cycling performance, and excellent temperature adaptability. Kundu et al. [31] prepared the Zn0.25V2O5⋅nH2O, in which V2O5 interlayer was supported by the Zn ions and the interlayer distance of the (200) planes is 0.537 nm. The presence of pillars leads to reversible Zn2+ (de)insertion at high rates, excellent capacities (~ 300 mAh·g−1), and long-term perfor- mance (> 1000 cycles). Compared with Zn0.25V2O5⋅nH2O, Ca0.25V2O5⋅nH2O achieved an expanded layer space of 1.06 nm and delivered higher gravimetric capacity of 340 mAh·g−1 at 0.2 C [32].

Hydrated layer V3O8 also exhibits excellent performance in AZIBs. For example, V3O8 layers were investigated to form V3O7·H2O, in which hydrogen-bond vibration could provide the elastic buffer space to make Zn2+ insertion/extraction easier and maintain the stable structure [33]. Zhang et al. [34] reported a high‐performance V5O12·6H2O nanobelt cathode, which deliv- ered a capacity of 354.8 mAh·g−1 at 0.5 A·g−1, a high energy

Fig. 3 a Schematic illustration of zinc‐storage mechanism in the Na0.33V2O5 electrode (left) and galvanostatic charge/discharge curves of it at 0.2 A·g−1 (right). Reproduced with permission from Ref. [25]. Copyright 2018 Wiley. b Structural characterization of MgxV2O5·nH2O cathodes (left), and scanning electron microscope

(SEM), transmission electron microscope (TEM), high resolution TEM (HRTEM) images, and TEM–energy dispersive X-ray spec- trometry  (EDS) elemental mappings of prepared MgxV2O5·nH2O (right). Reproduced with permission from Ref. [17]. Copyright 2018 ACS

(5)

density of 194 Wh·kg−1, and good cycle stability. Layered barium vanadate with different amount of Ba ions and water atoms were also investigated, in which water co-intercalation in Ba1.2V6O16·3H2O contributed to robust structure, a high capacity of 108.8 mAh·g–1 at 10 A·g–1, and capacity retention of 95.6% over 2000 cycles [35]. The co-introduction of hydrated, monovalent- and divalent-cations into the galleries of V3O8 lay- ers can contribute to pseudocapacitive behavior and improved layer structure stability [36, 37]. For example, the prepared NaCa0.6V6O16·3H2O cathode shows excellent performance [38].

Layered VOPO4 [39] also can be modified by pre-interca- lating H2O molecules to weaken the electrostatic repulsion between the intercalating Zn2+ and the host. Simultaneously, using electrolyte with controlled water amounts could fur- ther improve the ion transfer and reversibility. However, in common Zn(CF3SO3)2(Zn(OTF)2) electrolyte, it still under- goes significant voltage decay. After introducing high con- centration ZnCl2 salt in the electrolyte, Zn/VOPO4 battery delivered a high capacity (170 mAh·g−1) and stable retention of both capacity and voltage over 500 cycles [40].

To sum up, chemical pre-intercalation of ions (such as monovalent ions, Li+, Na+, K+, NH4+, divalent ions, Mg2+, Cu2+, Ba2+.) and molecules (H2O, small organic molecules,

etc.) was proved to be an effective strategy to increase the interlayer space, enhance the stability of normal layered structure, and suppress vanadium dissolution upon the elec- trochemical process. However, the intercalation of metal ions may reduce the specific capacity of layer materials.

2.2 Tunnel‑based structure

Tunnel-based vanadium oxides also exhibit excellent perfor- mance in AZIBs because of their unique tunnel pathways for Zn2+ migration. Taking VO2(B) as an example, as shown in Fig. 4a, it displayed ultrafast kinetics of Zn2+ transportation with little volume change during cycling. After 50 cycles, the VO2(B) cathode still delivered a reversible capacity of 357 mAh·g−1 at 0.25 C [41]. Reduced graphene oxide (rGO)/

VO2 contributed an excellent power density, energy density, and cycling stability, demonstrating an energy density of 65 Wh·kg−1 even at a power density of 7.8 kW·kg−1 as well as a capacity retention of 99% after 1000 cycles at 4 A·g−1 [42].

As depicted in Fig. 4b, α-Zn2V2O7 with distorted trigonal bipyramid ZnO5 polyhedral layers connected via VO4 tetra- hedra could provide the tunnel network for Zn-ion transfer.

Then, a high capacity of 138 mAh·g−1 over 1000 cycles was

Fig. 4 a Schematic illustration of a rechargeable aqueous Zn-VO2(B) battery (left) and its cycling performance at 0.25 C (inset: typical gal- vanostatic charge-discharge curve) (right). Reproduced with permis- sion from Ref. [41]. Copyright 2018 Wiley. b Galvanostatic cycling

performance of a rechargeable aqueous Zn-Zn2V2O7 battery at a cur- rent density of 4.0 A·g−1 (insert: schematic of electrochemical regula- tion in α-Zn2V2O7 nanowires for AZIBs). Reproduced with permis- sion from Ref. [43]. Copyright 2018 RSC

(6)

294 Y. Wu et al.

achieved [43]. Analogously, CuV2O6 has also been considered as the active materials for AZIBs. It demonstrated reversible phase transformation between CuV2O6 and ZnV2O6 during cycling, delivering a high capacity of 427 mAh·g–1 and capac- ity retention of 99.3% after 3000 cycles [18].

In addition, Guo et al. [44] compared the electrochemi- cal performances between different structural silver vana- date cathodes, and found that tunnel structural Ag0.33V2O5 show reversible Zn2+ (de)insertion during cycling and good cycle stability of 83% capacity retention after 100 cycles at 1 A·g−1. Similarly, Guo et al. [45] also compared the elec- trochemical performances between the NaV3O8-type lay- ered structure (Na5V12O32) and β-Na0.33V2O5-type tunneled structure (Na0.76V6O15). They discovered that although the tunneled structure has lower ion diffusion coefficients than layered structure, it displays superior cyclic stability. The channel provided by irreversible exchanging between Ca2+

in Ca0.67V8O20·3.5H2O and Zn2+ enabled a capacity of 466 mAh·g–1 and a balanced energy power-density behavior [46].

Moreover, tunneled sodium super ionic conductor (NASICON)-typed materials with very stable framework and rapid ionic diffusion capability, have been recognized as a promising Zn2+ storage as well as intercalation hosts [47]. After the extraction of Na+ ions, it shows a fast dif- fusion-controlled Zn-ion storage kinetic in Zn(CH3COO)2 electrolyte, delivering a capacity of 97 mAh·g−1 as well as only 26% decay after 100 cycles (as shown in Fig. 5a) [48]. Hu et al. [49] revealed that its high structural stability mainly depends on mixed occupation of Na+/Zn2+ at both two inequivalent Wyckoff sites including 6b sites and 18e sites. The excellent electronic conductivity and mechanical properties were then achieved. In a more recent study, Hu et al. [50] realized a simultaneous Zn2+/Na+ de/insertion of Na3V2(PO)3/rGO microspheres in 2 mol·L−1 Zn(OTF)2

Fig. 5 a Schematic illustration of Zn ions’ insertion and extraction in Na3V2(PO4)3 (up), galvanostatic charge-discharge curves (lower left), and cycling performance at a current density of 0.5 C for the Zn//0.5 mol·L−1 Zn(CH3COO)2//Na3V2(PO4)3 battery. Reproduced with permission from Ref. [48]. Copyright 2016 Elsevier. b Sche-

matic illustration of a rechargeable aqueous zinc‐ion battery with Na3V2(PO4)2F3 as cathode and Zn foil as anode (left) and cycling per- formance of a carbon film functionalizing (CFF)-Zn//N3VPF@C bat- tery at a current density of 1 A·g−1 (right). Reproduced with permis- sion from Ref. [52]. Copyright 2018 Elsevier

(7)

electrolyte and delivered a specific capacity of 114 mAh·g−1 with an operating voltage of 1.23 V (vs. Zn2+/Zn).

Moreover, inspired by that F atoms having a strong affinity with surroundings [51], Na3V2(PO4)2F3 was inves- tigated as positive materials for AZIBs, achieving 0.5 V higher work voltage than that of Na3V2(PO4)3. The oper- ating voltage was further enhanced to 1.62 V after it was combined with carbon nanotube (CNT) and good long cycle performance of only 5% capacity decay over 4000 cycles was obtained (Fig. 5b) [52].

To sum up, V-based materials with tunnel structure show excellent performance. For example, VO2 with a shear-type structure which results in larger resistance to lattice shear- ing against the insertion/extraction of ions, and some NASI- CON-typed materials achieved high working voltage (~ 1.6

V), which is higher than most of V-base materials with other structures. However, they also have some disadvantages, such as low ion and electrical conductivity, which need to be fur- ther improved to meet the demand of next-generation AZIBs.

2.3 Other structures

There are also some vanadium-based cathode materials with other different structures, such as the VS4 (as shown in Fig. 6a) with chain‐like structure. Its loosely stacked struc- ture had open channels for transfer and storage of Zn ions, achieving excellent capacity of 310 mAh·g−1 [19].

Zhang et  al. [53] proved that Mn cation defective ZnMn2O4 with spinel structure (Fig. 6b) has abundant Mn vacancies to enable fast Zn2+ ion diffusion and good

Fig. 6 a Lateral and vertical views of the crystalline structure of VS4 (left), galvanostatic charge/discharge curves of the Zn-VS4 battery at different current densities (middle), and cycling performance of the Zn-VS4 battery at a current density of 2.5 A·g−1 (right). Repro- duced with permission from Ref. [19]. Copyright 2020 RSC. b Sche- matic illustration of Zn-ion storage of ZnMn2O4 (ZMO) cathode in Zn(CF3SO3)2 Electrolyte (left), schematic illustration of Zn2+ inser- tion/extraction in an extended three-dimensional ZMO spinel frame-

work (middle) and proposed Zn2+ diffusion pathway in ZMO spinel without and with Mn vacancies (right). Reproduced with permission from Ref. [53]. Copyright 2016 ACS. c Probable Zn-ion-binding sites for coordination with the V-MOF-48@carbon nanotube fibers (CNTF) (left), specific capacities of V-MOF-12//Zn, V-MOF-24//Zn, V-MOF- 48//Zn, and V-MOF-60//Zn at various current densities (middle) and Ragone plot of fiber-shaped V-MOF-48//Zn batteries (right). Repro- duced with permission from Ref. [54]. Copyright 2019 Elsevier

(8)

296 Y. Wu et al.

structural stability. 3-dimensional (3D) conductive V-based metal–organic frameworks (MOFs) (Fig. 6c) with abundant active sites, good conductivity, and hierarchical porosity were also considered as electrode materials for AZIBs. After being arrayed on carbon nanotube fibers, the combination delivered a volumetric capacity of 101.8 mAh·cm−3 at 0.1 mA·cm−3 and an excellent rate capability. More importantly, its high energy density of 17.4 mWh·cm−3 at a power density

of 1.46 W·cm−3 offered a prospect of conductive MOFs nanowires-based energy storage devices [54].

In addition, vanadium nitride with surface-oxide (VNxOy) used as electrode materials for AZIBs is face-centered struc- ture, which possesses high electrical conductivity, ionic con- ductivity, and improved supercapacitor‐like surface redox reac- tions. There are Zn2+ and H+ de‐/intercalation accompanied by multiple redox reactions including V3+↔V2+ and N3−↔ N2−,

Fig. 7 a Zinc ion diffusion in layered V2O5: diffusion pathways viewed along the [100], [010], and [001] directions (upside) and corresponding calculated minimum energy paths of Zn2+ diffusion.

Blue, red, and grey balls represent V, O, and Zn atoms, respectively.

Black balls indicate the most energetically favorable location for Zn2+

intercalation. Reproduced with permission from Ref. [58]. Copyright

2019 Elsevier. b Schematic illustration of the Zn-V2O5 battery with 2 mol·L−1 ZnSO4 aqueous electrolyte (left), capacitive contribution current at 0.1 mV·s−1 (upper right), and the capacitive contribution ratio at different scan rates (lower right). Reproduced with permission from Ref. [59]. Copyright 2019 Elsevier

(9)

and release/uptake of anions (OH) on the VNxOy surface, delivering a high capability of 200 mAh·g−1 at 30 A·g−1 [55].

To sum up, the structure of V-based materials used as the active materials for AZIBs is mainly the layered structure, ion or/and water atoms pre-inserted layered materials, tun- neled materials, and so on. Current works mainly focus on enlarging the interlayer space and exploring materials with different structures which could be an excellent host and transmission medium for Zn2+ ions.

3 Storage mechanisms

For the vanadium-based cathode, its energy storage mechanism involves traditional reversible Zn-ion insertion/de-insertion mechanism, H+-Zn2+ co-insertion mechanism, and H2O-Zn2+

co-insertion mechanism. Detailed works are discussed below.

3.1 Zn‑ion insertion/de‑insertion mechanism A lot of vanadium-based cathodes in AZIBs adopt the tra- ditional reversible Zn-ion insertion/extraction mechanism, such as polyaniline (PANI)-intercalated V2O5 [56], double- layered calcium vanadium oxide bronze (Ca0.25V2O5·nH2O, CVO), and two-dimensional (2D) amorphous V2O5/graphene heterostructures (A-V2O5/G) [57]. As shown in Fig. 7a, the density functional theory (DFT) calculation found that the preferred coordination environment of Zn2+ ions is similar to that of Li+ ions and intercalated into the interlayer region with the most favorable pathway along the [100] channel [58]. Figure 7b demonstrated the capacitive contribution ratio at different scan rates of the V2O5 nanopaper and confirmed that Zn-ion storage behavior is usually manipulated by both capacitive reaction and ion diffusion-controlled process [59].

In addition, He et al. [25] reported the Na0.33V2O5 (NVO) electrode with the improvement conductivity benefiting from the intercalation of Na-ions among [V4O12]n layers, in which a new phase ZnxNa0.33V2O5 (0.42 < x < 0.96) formed/disap- peared in a voltage range between 0.7 V and 0.2 V, and confirmed that Na-ions are unchanged during the charge/

discharge process and only play a “binders/pillars” role to maintain structural stability. Similarly, two new phases of Zn0.25V2O5·nH2O and Zn0.29V2O5 also occurred for the LixV2O5·nH2O cathode upon charging/discharging [30].

3.2 H2O–Zn2+ Co‑insertion mechanism

Besides the traditional Zn2+ carrier mechanism, H2O and Zn2+ co-intercalation also commonly exists in vanadium- based cathode [15, 28, 35, 55, 60–62]. For example, Zn2+

accompanied H2O in the form of complex intercalated in the layer gap was also observed in layered MgxV2O5·nH2O

cathod e via X-Ray Diffraction (XRD) analysis and the schematic illustration is shown in Fig. 8a [17]. In lay- ered V2O5 with the chemical pre-intercalation of Cu2+, Zn3(OH)2V2O7·2H2O and Zn0.25V2O5·H2O appeared during the first cycling (Fig. 8b [28]). H2O molecules de-interca- lated with Zn2+ ions into the Zn0.25V2O5·nH2O could interact with the oxygen layers, buffer the high charge density, and make charge transfer easier at the electrode interface [31].

Generally, the multivalent ions usually distort the mate- rial’s original crystal structure, because its strong charge density hinders the diffusion of charge carriers through the host matrix. However, co-intercalation water can shield the electrostatic interactions and facilitate the insertion of mul- tivalent ions at the interface and the host framework [2].

Additionally, it is worth noting that when the interlayer space is enlarged upon cycling, the Zn2+ accompanied by water molecules inserted in the host may be observed [28].

In summary, water molecules have a great positive effect on the process of Zn2+ transport and structure stability.

3.3 H+– Zn2+ Co‑insertion mechanism

H+– Zn2+ co-insertion mechanisms are widely studied in vanadium-based cathode in AZIBs, such as NaV3O8·1.5H2O (NVO) nanobelts cathode. During the first discharging, the water from the aqueous ZnSO4/Na2SO4 electrolyte decom- posed as H+ and OH. Then, OH ions coupled with Zn2+

from ZnSO4 to form Zn4SO4(OH)6·4H2O. Simultane- ously, the H+ coupled with Zn2+ inserted into NVO cath- ode and H3.9NaZn0.5V3O8·1.5H2O was formed. After being fully charged, the H+ and partial Zn2+ are simultaneously extracted and NaZn0.1V3O8·1.5H2O is formed (as shown in Fig. 9a) [63]. Moreover, Li et al. [64] reported that there is a competition between proton and Zn2+ insertion in VO2 cathode. They also found that structural distortion result- ing from proton insertion is minimal (as shown in Fig. 9b).

The similar H+ and Zn2+ co-insertion mechanism was also reported in Zn0.3V2O5·1.5H2O cathode [65], carbon nano- tube delaminated V2C novel 2D nanostructured materials (MXene) cathode [66], as well as KV3O8·0.75H2O (KVO)/

single wall carbon nanotube (SWCNT) cathode [62].

3.4 Cationic and anionic oxidation co‑exist mechanism

To our knowledge, most vanadium-based cathodes per- form an insertion/extraction mechanism with the charge compensation from the V5+/V4+/V3+ redox reaction.

Nonetheless, reversible oxygen redox chemistry also exists, especially for VOPO4 cathode. In these materials, charge loss generated from the process of Zn2+ insertion/

extraction in the low-voltage range usually compensates

(10)

298 Y. Wu et al.

by the V4+/V5+ redox reaction. Lattice oxygen oxidation and reduction reaction proceeded without Zn2+ insertion/

extraction during the voltage range of 1.8– 2.1 V. The oxy- gen oxidation process increases the energy density and enhances the rate capability and cycling performance [24].

Additionally, Fang et al. [55] reported a novel simultane- ous cationic and anionic redox reaction in the VNxOy cath- ode. The Zn2+ insertion/extraction reaction depends on a redox reaction including V3+ ↔ V2+ and N3− ↔ N2− (as shown in Fig. 10a). In conclusion, cationic and anionic

Fig. 8 a Schematic illustrations of storage mechanism of MgxV2O5·nH2O cathode upon charging/discharging. Reproduced with permission from Ref. [17]. Copyright 2018 ACS. b Ex-situ XRD pat- terns and corresponding galvanostatic charge/discharge (GCD) curves

at a current density of 0.1 A·g−1 (up), and schematic illustration of storage mechanism in the first cycle of Cu0.1V2O5·0.08H2O electrode (lower part). Reproduced with permission from Ref. [28]. Copyright 2019 Elsevier

(11)

oxidation co-exist mechanism has a profound significance for high energy density cathode and still needs further investigation.

3.5 Insertion and conversion co‑exist mechanism Apart from the traditional intercalation mechanism, con- version reaction also exists in some vanadium-based cath- odes. For example, VS4@rGO cathode adopted an interca- lation and conversion coexistence mechanism. There was a Zn3(OH)2V2O7·2H2O phase generated in the full discharge state of 0.35 V and diminished after charging to 0.8 V.

Until charged to 1.8 V, the Zn3(OH)2V2O7·2H2O phase was reformed again with orthorhombic sulfur formation. There- fore, during the charging process, the conversion can be expressed as Eq. (1):

However, upon the discharge process, other compli- cated conversion reactions also occurred (as shown in Fig. 10b) [67]. The conversion reaction with the forma- tion of Zn3(OH)2V2O7·2H2O can also observe in the Mg0.19V2O5·0.99H2O cathode [68]. In addition, Ding et al. [69] observed the product of NH4+ in electrolyte for VN0.9O0.15 in initial charge state, which suggested that VN(Opoor) was converted to VN1−x(Orich) with large V and O vacancies/defects, accor ding to the following Eq.  (2):

(1) 2VS4+ 11H2O + 3Zn2++

→ Zn3(OH)2V2O7⋅2H2O + 8S + 16H++ 10e.

(2) VN(

Opoor)

+ 2xH2O→VN1−x( Orich)

+ xNH+4 + xe.

Fig. 9 a Second charge/discharge curve of NaV3O8·1.5H2O nanobelts at 0.1 A·g−1 (left), corresponding ex situ XRD patterns (right). Repro- duced with permission from Ref. [63]. Open Access, Copyright 2018 Fang Wan. b Schematic illustration of the electrochemical process in

a Zn-VO2 battery during discharging in a ZnSO4 aqueous electrolyte.

Process 1, Zn loses electrons and changes to Zn2+ cations. Process 2, hydrolysis of Zn2+ ions. Process 3 and 3′, Zn4(OH)6SO4·5H2O depo- sition and H+ insertion. [64]. Copyright 2019 Wiley

(12)

300 Y. Wu et al.

The extra X-ray photoelectron spectroscopy (XPS) testing process demonstrated a complete conversion reaction upon initial charging. Additionally, the conversion reaction also happened in some vanadium oxide with polyaniline inter- calation [70, 71].

In summary, even the energy storage mechanisms of AZIBs are very complex and affected by various factors.

Structure of vanadium-based materials plays very important role. For example, as depicted in Table 1, layered V-based cathode always demonstrate insertion mechanism. Conver- sion mechanism, and conversion and insertion co-mech- anism are more common in V-based materials with other structures. Therefore, clearly understanding the mechanism has great significance in designing a high-performance cath- ode. It is still a challenge and needs further investigation through various advanced characterization technology.

4 Conclusions and perspectives

Vanadium-based materials have been regarded as a kind of promising cathode materials for AZIBs. This review sys- tematically summaries main structures of V-based cathode, including layered, tunnel and other structure. Layered struc- ture attracts much attention due to their large number of active sites and intrinsic fast ions diffusion channels. By pre-intercalating guest ions, atoms, or even molecules into the structure, the interlayer space can be enlarged, and the structure can be stabilized. Thus, enhanced performance can be achieved. Furthermore, electrochemical reactions in aque- ous systems are much more complicated than those in their organic counterpart. Various energy storage mechanisms have been revealed, including Zn-ion insertion/extraction mechanism, H2O-Zn2+ co-insertion, H+-Zn2+ co-insertion,

Fig. 10 a High angle annular dark field (HAADF) and elemental mapping images of VNxOyna noflake (left) and ex situ XRD patterns of VNxOy electrodes during initial cycle (right). Reproduced with permission from Ref. [55]. Copyright 2019 Wiley. b Ex-situ XRD

patterns in the first charge/discharge cycling at 0.1 A·g−1. The inset shows the enlarged image of the XRD pattern of the D0.35 state in the 2 theta range from 11.4° to 12.6°. Reproduced with permission from Ref. [67]. Copyright 2018 RSC

(13)

Table 1 Typical vanadium-based cathode materials in AZIBs Cathode materialsRef.ElectrolyteCapacityCyclabilityMechanism Normal layer structure Atomic layer deposition (ALD)-derived V2O5[21]3 mol·L−1 Zn(CF3SO3)2513 mAh·g−1 (0.5 A·g−1)439 mAh·g−1 1000 (5 A·g−1)Zn2+/H+ co-intercalation Ball-milled commercial V2O5[22]3 mol·L−1 Zn(CF3SO3)2470 mAh·g−1 (0.2 A·g−1) 91.1% 4000 (5 A·g

−1)Zn2+/H2O co-intercalation VS2 nanosheets [23]1 mol·L−1 ZnSO4190.3 mAh·g−1 (0.05 A·g−1)

98.0% (0.5 A·g

−1)Zn2+/H2O co-intercalation VOPO4[24]21 mol·L−1 LiTFSI /1 mol·L−1 Zn(Tr)2139 mAh·g−1 (0.05 A·g−1)

93% 1000 (1 A·g

−1)Oxygen redox Zn2+ (de)intercalation Intercalated layer structure Na0.33V2O5 (NVO) nanowire [25]3 mol·L−1 Zn(CF3SO3)2367.1 mAh·g−1 (0.1 A·g−1)

93% 1000 (1 A·g

−1)Zn2+ (de)intercalation (NH4)2V10O25·8H2O [27]1 mol·L−1 ZnSO4376 mAh·g−1 (0.5 A·g−1)

93% 1000 (10 A·g

−1)Zn2+ (de)intercalation Zn0.25V2O5·nH2O [31]1 mol·L−1 ZnSO4220 mAh·g−1 (4.5 A·g−1)

80% 1000 (1.2 A·g

−1)Zn2+/H2O co-intercalation Ca0.24V2O5·0.83H2O nano- belt [32]1 mol·L−1 ZnSO4340 mAh·g−1 (0.2 C)

96% (80 C) 3000

Capacitive behavior Zn2+ (de)intercalation Mg0.34V2O5· 0.84H2O nanobelt [17]3 mol·L−1 Zn(CF3SO3)2264 mAh·g−1 (1 A·g−1)97% (5 A·g−1) 2000Capacitive behavior Zn2+ / Mg2+ intercalation CuxV2O5·nH2O [28]2 mol·L−1 ZnSO4410 mAh·g−1 (0.5 A·g−1)180 mAh·g−1 10,000 (10 A·g−1)Zn2+/H2O co-intercalation NaCa0.6V6O16·3H2O nanobelt [38]3 mol·L−1 Zn(CF3SO3)2347 mAh·g−1 (0.1 A·g−1)83% (5 A·g−1) 10,000Pseudocapacitive Zn2+/H2O co-intercalation H2V3O8 nanowire [36]3 mol·L−1 Zn(CF3SO3)2423.8 mAh·g−1 (0.1 A·g−1)94.3% (5 A·g−1) 1000Capacitive behavior Zn2+ (de)intercalation V3O7·H2O nanogrid [33]3 mol·L−1 Zn(CF3SO3)2481.3 mAh·g−1 (0.1 A·g−1)85.4% (5 A·g−1) 1000Pseudocapacitance Zn2+ (de)intercalation Polypyrrole (PPy)-VOPO4[39]1 mol·L−1 Zn(CF3SO3)2 in acetonitrile 10% water86 mAh·g−1 (25 mA·g−1)86 mAh·g−1 350 (0.1 A·g−1)Zn2+ (de)intercalation VOPO4·H2O [40]13 mol·L−1 ZnCl2/ 0.8 mol·L−1 H3PO4 mixed electrolyte

170 mAh·g−198 mAh·g−1 500 (2A·g−1)Capacitive behavior Zn2+/H+ (de)intercalation Tunnel structure VO2 (B) nanofibers [41]3 mol·L−1 Zn(CF3SO3)2357 mAh·g−1 (75 mA·g−1)171 mAh·g−1 (90 A·g−1)Zn2+ (de)intercalation VO2 nanorods [65]1 mol·L−1  ZnSO4272 mAh·g−1 (3.0 A·g−1)75.5% (1 A·g−1) 945Zn2+/H+ (de)intercalation α-Zn2V2O7 nanowire [43]1 mol·L−1 ZnSO4170 mAh·g−1 (4.4 A·g−1)85% (4 A·g−1) 1000Zn2+ (de)intercalation

(14)

302 Y. Wu et al.

cationic and anionic oxidation co-exist, and insertion and conversion co-exist mechanisms. However, they are still in debate and need to be further investigated. Therefore, future efforts about vanadium-based cathode are suggested to be made in the following directions.

First, to obtain high-performance cathode materials for ZIBs, intrinsic structure should be properly controlled, vana- dium-based materials with large number of active sites, fast and favorable ion diffusion channels, and stable structure must be designed and fabricated. Such a s introducing metal cations act as "pillars" to enhance the structural stability, expanding the interlayer space to buffer the Zn2+ electro- static force through pre-intercal ating cations or water mol- ecules, combining the conductive substrate or matrix to form composites to improve its electronic conductivity, downsiz- ing the materials to nanoscale to shorten the transfer path thereby increasing the ion transfer kinetic, and designing the compatible electrolyte with the cathode or adding additives to electrolyte to activate the anionic redox of cathode thereby further increasing its capacity.

Second, to further explore the electrochemical reactions during the charging/discharging process, high-resolution in- situ characterization technologies and simulation are also needed, which will contribute a lot to design V-based cath- odes with ideal properties for AZIBs.

Acknowledgements This work was financially supported by the School Research Startup Expenses of Harbin Institute of Technology (Shenz- hen) (Grant No. DD29100027).

Compliance with Ethical Statement

Conflict of interest The authors declare no conflict of interest.

References

1. Zhang Y, Tang Q, Zhang Y, Wang J, Stimming U, Lee AA.

Identifying degradation patterns of lithium ion batteries from impedance spectroscopy using machine learning. Nat Commun.

2020;11(1):1.

2. Shin J, Choi DS, Lee HJ, Jung Y, Choi JW. Hydrated intercala- tion for high-performance aqueous zinc ion batteries. Adv Energy Mater. 2019;9(14):1900083.

3. Tang B, Shan L, Liang S, Zhou J. Issues and opportuni- ties facing aqueous zinc-ion batteries. Energy Environ Sci.

2019;12(11):3288.

4. He P, Chen Q, Yan M, Xu X, Zhou L, Mai L, Nan CW. Building better zinc-ion batteries: A materials perspective. Energy Chem.

2019;1(3): 100022.

5. Li H, Yang Q, Mo F, Liang G, Liu Z, Tang Z, Ma L, Liu J, Shi Z, Zhi C. MoS2 nanosheets with expanded interlayer spacing for rechargeable aqueous Zn-ion batteries. Energy Stor Mater.

2019;19:94.

6. Wu Z, Yuan X, Jiang M, Wang L, Huang Q, Fu L, Wu Y. Zinc- carbon paper composites as anodes for Zn ion batteries: keys Table 1 (continued) Cathode materialsRef.ElectrolyteCapacityCyclabilityMechanism Ag0.33V2O5[45]2 mol·L−1 ZnSO4350 mAh·g−1 (50 mA·g−1)83% (1 A·g−1) 100displacement mechanism Zn2+ (de)intercalation Na3V2(PO4)3[49]0.5 mol·L−1 Zn(CH3COO)297 mAh·g−1 (0.5 C)

74% (0.5 C) 100

Zn2+ (de)intercalation Na3V2(PO4)3/rGO micro- spheres [51]2 mol·L−1 Zn(CF3SO3)2107 mAh·g−1 (50 mA·g−1 )75% (0.5 A·g−1 ) 200Zn2+ /Na+ co-intercalation

Carbon coated Na

3V2(PO4)2F3[53]2 mol·L−1 Zn(CF3SO3)261.7 mAh·g−1 (0.2 A·g−1)95% (1 A·g−1) 4000Zn2+ (de)intercalation Other structure VNxOy nanoflake [56]2 mol·L−1 ZnSO4200 mAh·g−1 (30 A·g−1)150 mAh·g−1 2000 (20 A·g−1)Zn2+/H+ intercalation Surface anionic redox VS4[19]1 mol·L−1 ZnSO4310 mAh·g−1 (0.1 A·g−1)85% (2.5 A·g−1) 500Capacitive Zn2+ (de)intercalation VS4@rGO [68]1 mol·L−1 Zn(CF3SO3)2180 mAh·g−1 (1 A·g−1)93.3% (1 A·g−1) 165Conversion Zn2+ (de)intercalation VN0.9O0.15[70]3 mol·L−1 Zn(CF3SO3)2124 mAh·g−1 (102.4 A·g−1)

100% 1500 (4.26 A·g

−1)Conversion reaction

Referenzen

ÄHNLICHE DOKUMENTE

In summary, doping/coating of tungsten and related ele- ments shows great potential to improve the electrochemical performances of layered structure cathode materials (NCM and NCA)

In recent years, vanadium oxides, as cathode materials for LIBs, have attracted wide attention [9–12] Their rich valence states impart vana- dium oxide electrodes with

To examine the utility of the Mg foil quasi- reference electrode a micro polarization test against the counter electrode in 0.25 M APC electrolyte was carried out.

The multi-valence nature of vanadium means that its geochemical behaviour will be ƒO 2 -dependent, so that its concentration or V/Sc (or V/Ga), can serve as proxies for

Major element compositions of clinopyroxene and other minerals necessary for computing oxidation state were measured with JEOL JXA-8900 superprobe and are reported

1-3 Owing to their non-rocking-chair operation mechanism, how- ever, the practical deployment of graphite dual-ion batteries is inherently limited by the need for large quantities

1) The material contains a readily reducible/oxidizable ion; for example a transition met- al ion. 2) The material reacts with lithium in a reversible manner and the lithium

Aqueous zinc-ion batteries (ZIBs) are another cost-effective rechargeable battery for stationary grid energy. In stationary applications, high energy density is not the main