Skip to content
BY 4.0 license Open Access Published by De Gruyter April 18, 2022

Robust synthesis of a composite phase of copper vanadium oxide with enhanced performance for durable aqueous Zn-ion batteries

  • Haoyan Cheng , Xuerong Li , Hao Hu EMAIL logo , Tongtong Yuan , Shiqian Zhou , Shuge Dai , Di Zhang and Kunming Pan EMAIL logo
From the journal Nanotechnology Reviews

Abstract

Rechargeable aqueous Zn-ion batteries (AZIBs) have attracted much interest as next-generation power sources due to their economical, safe, and capacity superiorities. However, the cathodes used in AZIBs always suffer from sluggish kinetics, inducing inadequate rate performance and poor cycle ability. Pre-intercalating transition metal element in the cathode materials offers an effective strategy for improving diffusion kinetics of Zn2+ and thus the electrochemical activity. In this work, different proportions of Cu pre-intercalated V2O5 were synthesized to form a composite phase of Cu0.4V2O5 and VO2·nH2O nanosheets through the hydrothermal method. The reversible redox reaction of Cu2+ and Cu0, accompanied by the phase changes of copper vanadate and zinc vanadate, contributes to an excellent battery performance. When the molar ratio between Cu precursor and commercial V2O5 in the reaction solution is 1:2, the obtained material presents an outstanding electrochemical performance with the initial discharge capacity of 332 mAh g−1 at 0.2 A g−1. The enlarged lattice distance together with the high conductivity leads to a high Zn ions diffusion rate of 10−5 cm2 s−1. Even after 1,000 cycles at a current density of 2 A g−1, the capacity attenuation is only 0.035% per cycle, exhibiting distinctive activities toward AZIBs.

1 Introduction

The energy storage device used for the collection of other green energies, such as solar and wind, is becoming increasingly urgent [1,2,3,4,5,6]. Rechargeable aqueous Zn-ion batteries (AZIBs) have captured a lot of attention because of their natural abundance, low price, high safety, environment-friendliness, and good conductivity [7,8,9,10], which also possess super high theoretical energy capacity (5,851 mA h cm−3), low redox potential, and impressive electrochemical stability in water [11,12,13,14].

Exploring appropriate cathode materials with favorable capacity and stable structure during the insertion/extraction of Zn2+ is becoming urgent for aqueous ZIBs. Layered VnOm is generally recognized as preferable cathodes for ZIBs due to comparable electrochemical activity and ultralow cost [15,16,17,18,19]. However, the interlayer distance in the original VnOm framework is confined for Zn2+, which sterically hinders the ion diffusion kinetics during cycling, and eventually leads to deteriorative performance and poor commercial applications [20]. Thus, the structure and compositions of V2O5-based materials should be carefully designed for enlarging the interlayer spacing and accelerating the insertion and extraction of Zn2+. Among the reported construction strategies, surface modification of V2O5 is considered one of the valid methods to boost ion diffusion by providing a larger contact area and shortening the ion diffusion path [16,21]. In addition, the introduction of interlayer structural water or/and metal ions (such as Li+ [22], Na+ [23], K+ [24], Ca2+ [25]) as pillars is also beneficial for the enlargement of interlayer spacing of V2O5 and accelerating the diffusion of Zn2+ [17]. For example, co-intercalation of water molecules and zinc in vanadium oxide bronze (Zn0.25V2O5·nH2O) could decrease the activation energy for charge transfer and reversibly enlarge/shrink the layered structure of Zn0.25V2O5 during Zn2+ (de)intercalation, leading to good kinetics, outstanding rate performance, and long-term cycle life [26]. In addition, hydroxyl copper was introduced to V2O5 to achieve copper pyrovanadate (Cu3V2O7(OH)2·2H2O) cathode for aqueous zinc ion batteries to improve the electrochemical performance [27]. Although the intercalation of Zn2+ would give rise to the reduction of Cu ions to metallic Cu0, allowing for an increased conductivity of Cu3V2O7(OH)2·2H2O, the Cu3(OH)2V2O7·2H2O presented fast fading in the initial cycles due to the phase change [28]. It is still a challenge to directly introduce copper in vanadium oxide bronze to obtain layered cathode materials with satisfactory electrochemical properties.

Herein, we demonstrated the robust synthesis of Cu pre-intercalated V2O5 to obtain cathode materials with a composite phase of Cu0.4V2O5 and VO2·nH2O for AZIBs through a simple one-step hydrothermal reaction (Figure 1), which was marked as (Cu0.4V2O5) x ·(VO2·nH2O) y . The influence of the different ratios of Cu on the structure and electrochemical performance of the final product was discussed in detail. The activity of the (Cu0.4V2O5) x ·(VO2·nH2O) y was found to be highly dependent on its composition. When the molar ratio between Cu precursor and c-V2O5 in the reaction solution was 1:2, the obtained material exhibits a superior reversible capacity of 332 mAh g−1 at 0.2 A g−1 within the voltage window of 0.3–1.6 V. And, the material displayed good long-term performance during 1,000 cycles with the capacity attenuation of 0.035% per cycle at 2 A g−1. We also elucidated that the charge storage chemistry is the multi-step Zn2+ insertion/extraction reaction with the replacement reaction. We hope this work can draw more attention to cathodes in AZIBs and drive the applications of AZIBs in massive-scale energy storage devices in the future.

Figure 1 
               Schematic diagram of the synthesis of (Cu0.4V2O5)
                     x
                  ·(VO2·nH2O)
                     y
                   nanosheets in this work.
Figure 1

Schematic diagram of the synthesis of (Cu0.4V2O5) x ·(VO2·nH2O) y nanosheets in this work.

2 Experimental

2.1 Chemicals and materials

C4H6CuO4·H2O (Guangfu Chemical Reagent, AR, 99%), C6H12O6 (Damao Reagent, AR, 99%), V2O5 (Shanpu Reagent, AR, 99%), and H2O2 (Beijing Chemical works, AR, 30%) were used.

2.2 Materials preparation

The composite phase of copper vanadium oxide nanosheets with different proportions of Cu was synthesized by a one-step hydrothermal method. First, 10 mL of commercial V2O5 (c-V2O5) with different concentrations (0.09, 0.135, 0.18, 0.36 mmol mL−1) was, respectively, added into deionized water (10 mL), followed by decanting of 30 wt% H2O2 (3 mL). The above-mixed solution was heated at 50°C for 20 min under magnetic stirring. Second, C4H6CuO4·H2O (0.9 mmol) and C6H12O6 (0.1 mmol) were subsequently dissolved in another 10 mL of deionized water and stirred for 10 min. Then, the mixture solution of C4H6CuO4·H2O and C6H12O6 was pulled into the above solution containing c-V2O5 and H2O2, followed by heating at 50°C for another 10 min. In the final reaction solution, the molar ratios between C4H6CuO4·H2O and c-V2O5 were 1:1, 1:1.5, 1:2, and 1:4. The mixture reaction solution was hydrothermally reacted at 180°C for 12 h and then collected after washing with DI water and ethanol. After drying at 80°C for 12 h, the materials were heat-treated at 200°C for 2 h under argon ambience. The obtained (Cu0.4V2O5) x ·(VO2·nH2O) y materials were named as CVO-1, CVO-1.5, CVO-2, and CVO-4, according to the molar ratios between Cu precursor and c-V2O5.

2.3 Materials characterizations

The crystallographic characterization of the products was tested by the D8 Advanced X-ray diffraction (XRD) with Cu Kα radiation. Scanning electron microscopy (SEM) and EDS mapping images were taken by the JSM-IT100 microscope. Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images were collected on the JEM-2100 electron microscope.

2.4 Electrochemical measurement

The working electrode material was consisted of a mixture of the active materials (70%), super P (20%), and polyvinylidene fluoride (PVDF) binder (10%) with N-Methyl pyrrolidone (NMP). All materials were dispersed to form a homogeneous slurry. The slurry was then uniformly coated on a stainless-steel foil (200 mesh, Canrd) as a current collector, and dried for 12 h at 60°C. Finally, the stainless-steel foil with active materials was cut into a circular disc (12 mm) with an active material of 2 mg cm−2. CR2032-type coin cells were fabricated with Zn foil as the anode (0.1 mm thick), Zn(CFSO3)2 solution (1.8 M) as electrolyte, and glass microfiber as the separators. The electrochemical performance was evaluated with a Neware battery testing system in the potential range of 0.3–1.6 V. The cycle voltammetry (CV) and electrochemical impedance spectroscopy (EIS) measurements were tested from 0.3 to 1.6 V vs Zn/Zn2+ by an electrochemical workstation (CHI760E) at the room temperature.

3 Results and discussion

X-ray diffraction (XRD) was carried out to confirm the crystal structure of the obtained materials. For CVO-4, with the least Cu intercalated V2O5, the characteristic peaks located at 25.4°, 30.3°and 49.5° can be assigned to the peaks of VO2·xH2O (JCPDS No.18-1445) (Figure 2a). The formed VO2·xH2O probably results from the reduction of c-V2O5 by glucose during the hydrothermal process. For CVO-2, CVO-1.5, and CVO-1 with the increased Cu content, the characteristic peaks at 26.43°, 28.59°, 31.41°, and 33.2° can be well indexed to the (−111), (104), (−304), and (−213) planes of Cu0.4V2O5 (JCPDS No.18-0464), indicating Cu2+ has been successfully inserted into the interlayer of V2O5 as pillars. All the obtained materials are the mixed phases of copper vanadate and zinc vanadate, with the chemical formula of (Cu0.4V2O5) x ·(VO2·nH2O) y . It can be speculated that the (Cu0.4V2O5) x ·(VO2·nH2O) y materials are composed of V-O polyhedrons and Cu-O polyhedrons as shown in Figure 1. And, the peak intensity of the Cu0.4V2O5 phase improved with the increase of Cu content, which means that more Cu-O polyhedrons arrange in the rows of V-O polyhedrons. Figure 2b exhibits the cyclic voltammogram (CV) of the samples with different Cu contents between 0.3 and 1.6 V at the rate of 1 mV s−1. The CV plot of c-V2O5 shows two pairs of redox peaks located at 0.99/0.95 V and 0.74/0.49 V, respectively. All the curves of CVO materials show similar shapes with three pairs of redox peaks, indicating the multi-step reaction during the ingress/egress of Zn2+. In general, CVO-2 displays the best electrochemical performance among all the groups, with the largest peak area, indicating the most (de)intercalation and the fast diffusion of Zn2+ during the charge and discharge processes. The CVO-2 electrode presents much better rate performance than other Cu2+ pre-intercalated V2O5 cathodes, with an average-discharged capacity of 321, 317, 294, 249, and 225 mAh g−1 at the rates of 0.3, 0.5, 1, 3, 5 A g−1, respectively (Figure 2c), which may be caused by proper Cu-O polyhedrons arranged in the rows of V-O polyhedrons. When the current density turns back to 0.3 A g−1, the capacity of CVO-2 recovers to 292 mAh g−1, with 91% retention of the initial discharge capacity, suggesting the highly reversible of the electrodes. The cycling performances of the CVO electrodes with different Cu contents are conducted at 0.1 A g−1. CVO-2 exhibits a more stable cycling performance than other cathodes over 100 cycles with a high-capacity retention of 77% (Figure 2d), suggesting that suitable Cu content can efficiently improve the cycling performance by expanding the interlayer spacing. Even the active materials in each cell increased to 4 mg, the cathode of CVO-2 still shows the highest initial capacity of 331 mAh g−1 and the most stable cycling performance (Figure S1).

Figure 2 
               (a) The XRD patterns and (b) CV curves of CVO-1, CVO-1.5, CVO-2, CVO-4, and c-V2O5. (c) Rate performance and (d) cycle performance of CVO-1, CVO-1.5, CVO-2, and CVO-4.
Figure 2

(a) The XRD patterns and (b) CV curves of CVO-1, CVO-1.5, CVO-2, CVO-4, and c-V2O5. (c) Rate performance and (d) cycle performance of CVO-1, CVO-1.5, CVO-2, and CVO-4.

We employed the SEM to observe the morphology of CVO-2 in Figure S2. Most of them present nanosheet structures with a width of 70–160 nm, length of 200–500 nm, and average thickness of 22 nm. The transmission electron microscopy (TEM) images of the CVO-2 are shown in Figure 3a–c. The lattice fringe with a d-spacing of 0.368 nm corresponds to the (−111) face of CVO-2 (Figure 3b). Figure 3c demonstrates that the obtained material was indeed a typical layered structure with an interlayer distance of 9.125 Å, much larger than the lattice spacing of previously reported vanadium-based cathodes, such as K0.5V2O5 (3.62 Å) [24], Zn0.25V2O5·nH2O (5.37 Å) [26], Na2V6O16·3H2O (7.1 Å) [29], and CaV6O16·3H2O (8.08 Å) [30]. The larger lattice spacing can provide more accommodation for the (de)intercalation of Zn2+, resulting in excellent electrochemical performance [31]. As shown in Figure 3d, energy-dispersive X-ray spectroscopy (EDS) mapping clearly shows that Cu, V, and O were uniformly distributed through the rectangular plate.

Figure 3 
               (a) TEM image of CVO-2, showing good uniformity in terms of size and morphology. (b) and (c) High-resolution TEM image. (d) SEM image and elemental mappings of V, O, and Cu of CVO-2.
Figure 3

(a) TEM image of CVO-2, showing good uniformity in terms of size and morphology. (b) and (c) High-resolution TEM image. (d) SEM image and elemental mappings of V, O, and Cu of CVO-2.

The electrochemical performance of CVO-2 nanosheets for the zinc ion storage was tested using a CR2032-type coin cell, in which zinc metal was utilized as the anode and CVO-2 as the cathode. All the cells were tested in the voltage window of 0.3–1.6 V vs Zn2+/Zn. Figure 4a displays the galvanostatic charge/discharge (GCD) curves of CVO-2, which matches well with the CV result shown in Figure 2b. The GCD curves show similar charge–discharge capacities in the first three cycles, indicating the high reversibility of CVO-2. Compared with the c-V2O5 cathode shown in Figure S3, CVO-2 displays higher coulombic efficiency (∼100%) and an initial capacity of 332 mAh g−1 at 0.2 A g−1, with an attenuation of 14% after 100 cycles, indicating the high cycling stability (Figure 4b). The compared SEM images of the CVO-2 electrode before and after cycling are shown in Figure S4a and b. Only a slight difference can be observed after cycling, evidencing excellent structural stability. When the electrolyte of Zn(CFSO3)2 solution (1.8 M) was replaced by ZnSO4 solution (1 M), the CVO-2 electrode still presented a high discharge capacity of 324 mAh g−1 at 0.2 A g−1 (Figure S5). Furthermore, even at the rate of 2 A g−1, the CVO-2 electrode exhibits excellent long-term performance with the capacity attenuation of 0.035% per cycle (Figure 4c). In the meantime, the CVO-2 electrode shows the highest capacity at different current densities, which is shown in Figure 4d by comparing with the reported K2V6O16·2.7H2O [32], Cu3V2O7(OH)2·2H2O [28], K2V3O8 [33], V2O5@CNTs [15], (NH4)2V6O16 [34], and K2V8O21 [35] cathodes for AZIBs.

Figure 4 
               (a) Galvanostatic charge/discharge (GCD) curves of CVO-2 in the initial three cycles at 0.2 A g−1. (b) The discharge/charge profiles of CVO-2 at 0.2 A g−1. (c) The cycling performance of CVO-2 at 2 A g−1. (d) Comparison of the CVO-2 electrode with previously reported vanadium oxide cathode materials for ZIBs.
Figure 4

(a) Galvanostatic charge/discharge (GCD) curves of CVO-2 in the initial three cycles at 0.2 A g−1. (b) The discharge/charge profiles of CVO-2 at 0.2 A g−1. (c) The cycling performance of CVO-2 at 2 A g−1. (d) Comparison of the CVO-2 electrode with previously reported vanadium oxide cathode materials for ZIBs.

To clearly explain the electrochemical kinetics of the CVO-2 electrode, the CV measurement at the rates of 0.1–1 mV s−1 was carried out (Figure 5a). It can be observed that the CV curves maintain similar shapes and the reduction peaks move to the left and oxidation peaks move to the right with the increased scan rates, indicating good rate performance. As shown in Figure 5b, a relevant analysis is executed to reveal the relationship between scan rate and peak current, based on equation (1) [36,37,38].

(1) i = a v b .

Figure 5 
               (a) The CV curves at different speeds of CVO-2. (b) log(i) vs log(v) plots in CV curves. (c) Contribution ratio at different scanning speeds. (d) Capacitive contribution at 0.5 mV s−1.
Figure 5

(a) The CV curves at different speeds of CVO-2. (b) log(i) vs log(v) plots in CV curves. (c) Contribution ratio at different scanning speeds. (d) Capacitive contribution at 0.5 mV s−1.

When the b is close to 0.5, it demonstrates that the electrochemical process is mainly controlled by diffusion; while the b is close to 1, it implies a capacitive controlled charge storage mechanism. If 0.5 < b < 1, it indicates the combination of diffusion-limited and capacitive-limited processes. The b values of peak 1 to peak 4 can be calculated to be 0.89, 0.78, 0.81, and 0.8, indicating that the corresponding redox reactions are dominated by the diffusion-controlled progress and capacitive-controlled progress. In addition, the corresponding current at different scanning rates can be declared by equation (2) [39,40]:

(2) i ( V ) = k 1 v + k 2 v 1 / 2 .

Figure 5c illustrates that the calculation results of capacitive contribution are 67.6, 75.6, 85.3, 94.7, and 96.5% at the sweep rates of 0.1, 0.3, 0.5, 0.8, and 1 mV s−1, respectively. When the scan rate is relatively low, the response currents are mainly attributed to the diffusion-controlled process. The capacitive-controlled process plays a crucial role in the electrochemical reactions with the increased scan rates. The red region in Figure 5d corresponding to the capacitive mechanism is accounted for 85.3% at 0.5 mV s−1.

Furthermore, we employed the galvanostatic intermittent titration technique (GITT) to analyze the electrochemical behaviors in Zn/CVO-2 batteries. As shown in Figure 6a, the zigzag curve is generated by the relaxation rest after the discharge/charge progress. According to the following equation (3) [41], the D Zn values can be calculated by the voltage recovery.

(3) D Zn = 4 Π m B V m M B S 2 Δ E S Δ E τ 2 .

Figure 6 
               The charge/discharge GITT curves of (a) CVO-2 and (b) c-V2O5. The calculated chemical diffusion coefficient for Zn2+ of (c) CVO-2 and (d) c-V2O5.
Figure 6

The charge/discharge GITT curves of (a) CVO-2 and (b) c-V2O5. The calculated chemical diffusion coefficient for Zn2+ of (c) CVO-2 and (d) c-V2O5.

In the equation, the molar volume of the materials, molecular mass, electrode/electrolyte contact area, and duration of the current pulse is denoted as V m, m B, M B, S, and τ, respectively. The transient potential change and the steady-state voltage change are denoted as E τ and E S.

The D Zn value of CVO-2 is in the order of 10−5 to 10−7 cm2 s−1, which is much higher than that of c-V2O5, CVO-1, CVO-1.5, and CVO-4 (Figure 6c and d, Figures S6–8), revealing super-fast kinetics could improve the diffusion of Zn2+ and promote the rate ability. To further illustrate the electrochemical behavior of the CVO-2 electrode, the electrochemical impedance spectroscopy (EIS) measurement is applied, and the results are shown in Figure S9. The semicircle (high-frequency region) is referred to as the charge transfer resistance (R ct) and relative capacitance [42,43]. The R ct of the CVO-2 electrode after 1 cycle and after 3 cycles were 71 and 77.5 Ω, respectively, which was much smaller than 113 and 118 Ω for the c-V2O5 electrode, indicating fast electrons charge transfers during cycling.

To deeply elucidate the Zn2+ storage mechanism, ex-situ XRD was employed in recording the structural evolution during different states. After totally discharging to 0.3 V, several new peaks located at around 24.7° and 33.1° can be observed (Figure 7a), which can be referred to as the formation of ZnV3O8. The peaks of copper located at 43.8°, 50.3° indicate that the insertion of zinc ions leads to the reduction of Cu2+ to Cu0. According to the results of Figure 7b, the redox peak at 1.07 V in the CV curves corresponds to the oxidation of Cu0 to Cu2+. The redox peaks at 1.32/1.3 V, 0.98/0.97 V, and 0.65/0.5 V portray the V5+/V4+ and V4+/V3+ redox couple in the CVO-2 electrode. Combining the results of ex-situ XRD and CV curves, it can be reasonably hypothesized that the Zn2+ storage mechanism is based on the displacement reaction (Figure 7c). During discharging, the insertion of Zn2+ drives the reductive reaction of Cu2+ to Cu0 with the phase evolution from (Cu0.4V2O5) x ·(VO2·nH2O) y to ZnV3O8. On the charging process, the mixture of ZnV3O8 and Cu0 reversibly converts to (Cu0.4V2O5) x ·(VO2·nH2O) y . The replacement of divalent cations between Zn2+ and Cu2+ provides multiple electrons charge transfers and less structural destructiveness, indicating high capacity and cyclic stability. Meanwhile, the appearance of Cu0 tremendously enhances the conductivity of the CVO electrode.

Figure 7 
               (a) The ex-situ XRD patterns of CVO-2. (b) The CV curves of CVO-2 at 0.1 mV s−1. (c) Schematic of reaction of (Cu0.4V2O5)
                     x
                  ·(VO2·nH2O)
                     y
                   on discharging/charging progress at 0.1 A g−1.
Figure 7

(a) The ex-situ XRD patterns of CVO-2. (b) The CV curves of CVO-2 at 0.1 mV s−1. (c) Schematic of reaction of (Cu0.4V2O5) x ·(VO2·nH2O) y on discharging/charging progress at 0.1 A g−1.

4 Conclusion

We have successfully synthesized the composite phase of copper vanadium oxide materials as the cathodes of AZIBs. When the molar ratios of Cu precursor and c-V2O5 are 1:2, the Cu pre-intercalated V2O5 nanosheets (CVO-2) express the best electrochemical performance, showing the initial discharge capacity of 332 mAh g−1 at the rate of 0.2 A g−1. The super high diffusion of Zn2+ demonstrates the good conductivity of the material. The replacement mechanism provides excellent electrochemical performance. The achievement results clearly indicate that the appropriate intercalation of Cu is able to improve the properties of vanadium oxide materials, which serves as a promising candidate for high-performance AZIBs.


# These authors contributed equally to this work.


  1. Funding information: This work was financially supported by the National Natural Science Foundation of China (Grant No. 52002119 and 52102346), the State Key Lab of Advanced Metals and Materials (No. 2020-Z14), and the Startup Funds from the Henan University of Science and Technology (13480095, 13480096, 13554031, and 13554032).

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The authors state no conflict of interest.

  4. Date available statement: All data generated or analysed during this study are included in this published article [and its supplementary information files].

References

[1] Guerra OJ. Beyond short-duration energy storage. Nat Energy. 2021;6(5):460–1.10.1038/s41560-021-00837-2Search in Google Scholar

[2] Liu MK, Zhang P, Qu ZH, Yan Y, Lai C, Liu TX, et al. Conductive carbon nanofiber interpenetrated graphene architecture for ultra-stable sodium ion battery. Nat Commun. 2019;10:3917.10.1038/s41467-019-11925-zSearch in Google Scholar PubMed PubMed Central

[3] Hu H, Cheng HY, Song KX, Dai SG, Liu Y, Stock H-R, et al. Nd3+ ions induced rational morphology control of transition metal oxides for high energy storage performance. J Power Sources. 2020;472:228599.10.1016/j.jpowsour.2020.228599Search in Google Scholar

[4] Tan J, Li D, Liu Y, Zhang P, Qu Z, Yan Y, et al. Self-supported 3D aerogel network lithium-sulfur battery cathode: sulfur spheres wrapped with phosphorus doped graphene and bridged with carbon nanofibers. J Mater Chem A. 2020;8:7980–90.10.1039/D0TA00284DSearch in Google Scholar

[5] Yan Y, Cheng HY, Qu ZH, Yu R, Liu F, Ma QW, et al. Recent progress on the synthesis and oxygen reduction applications of Fe-based single-atom and double-atom catalysts. J Mater Chem A. 2021;9:19489–507.10.1039/D1TA02769GSearch in Google Scholar

[6] Zhao S, Kang YJ, Liu MJ, Wen BH, Fang Q, Tang YY, et al. Modulating the electronic structure of nanomaterials to enhance polysulfides confinement for advanced lithium–sulfur batteries. J Mater Chem A. 2021;9:18927–46.10.1039/D1TA02741GSearch in Google Scholar

[7] Li XR, Cheng HY, Hu H, Pan KM, Yuan TT, Xia WT. Recent advances of vanadium-based cathode materials for zinc-ion batteries. Chin Chem Lett. 2021;32(12):3753–61.10.1016/j.cclet.2021.04.045Search in Google Scholar

[8] Yi TF, Qiu LY, Qu JP, Liu HY, Zhang JH, Zhu YR. Towards high-performance cathodes: design and energy storage mechanism of vanadium oxides-based materials for aqueous Zn-ion batteries. Coord Chem Rev. 2021;446:214124.10.1016/j.ccr.2021.214124Search in Google Scholar

[9] Xu CX, Jiang JJ. Electrolytes speed up development of zinc batteries. Rare Met. 2021;40(4):749–51.10.1007/s12598-020-01628-6Search in Google Scholar

[10] Yan Y, Liang S, Wang X, Zhang MY, Hao SM, Cui X, et al. Robust wrinkled MoS2/N-C bifunctional electrocatalysts interfaced with single Fe atoms for wearable zinc-air batteries. Proc Natl Acad Sci USA. 2021;118(40):e2110036118.10.1073/pnas.2110036118Search in Google Scholar PubMed PubMed Central

[11] Liu ZX, Yang YQ, Liang SQ, Lu BG, Zhou J. PH-buffer contained electrolyte for self-adjusted cathode-free Zn–MnO2 batteries with coexistence of dual mechanisms. Small Struct. 2021;2(11):2100119.10.1002/sstr.202100119Search in Google Scholar

[12] Li CP, Xie XS, Liu H, Wang PJ, Deng CB, Lu BG, et al. Integrated ‘all-in-one’ strategy to stabilize zinc anodes for high-performance zinc-ion batteries. Natl Sci Rev. 2022;9(3):nwab 177.10.1093/nsr/nwab177Search in Google Scholar PubMed PubMed Central

[13] Fang GZ, Zhou J, Pan AQ, Liang SQ. Recent advances in aqueous zinc-ion batteries. ACS Energy Lett. 2018;3(10):2480–501.10.1021/acsenergylett.8b01426Search in Google Scholar

[14] Luo SH, Cao XX, Su Q, Zhang YP, Liu SY, Xie XF, et al. Layered barium vanadate cathodes for aqueous zinc batteries: enhancing cycling stability through inhibition of vanadium dissolution. ACS Appl Energy Mater. 2021;4(6):6197–204.10.1021/acsaem.1c00979Search in Google Scholar

[15] Chen HZ, Qin HG, Chen LL, Wu J, Yang ZH. V2O5@CNTs as cathode of aqueous zinc ion battery with high rate and high stability. J Alloys Compd. 2020;842:155912.10.1016/j.jallcom.2020.155912Search in Google Scholar

[16] Liu F, Chen ZX, Fang GZ, Wang ZQ, Cai YS, Tang BY, et al. V2O5 nanospheres with mixed vanadium valences as high electrochemically active aqueous zinc-ion battery cathode. Nano-Micro Lett. 2019;11(1):25.10.1007/s40820-019-0256-2Search in Google Scholar PubMed PubMed Central

[17] Yan MY, He P, Chen Y, Wang SY, Wei QL, Zhao KN, et al. Water-lubricated intercalation in V2O5·nH2O for high-capacity and high-rate aqueous rechargeable zinc batteries. Adv Mater. 2018;30(1):1703725.10.1149/MA2018-01/1/56Search in Google Scholar

[18] Wang SY, Zhu KJ, Yang LY, Li HZ, Wang SY, Tang SS, et al. Synthesis and study of V2O5/rGO nanocomposite as a cathode material for aqueous zinc ion battery. Ionics. 2020;26:5607–15.10.1007/s11581-020-03705-3Search in Google Scholar

[19] Chen S, Li K, Hui KS, Zhang JT. Regulation on lamellar structure of vanadium oxide via polyaniline intercalation for high-performance aqueous zinc-ion battery. Adv Funct Mater. 2020;30(43):2003890.10.1002/adfm.202003890Search in Google Scholar

[20] Liu SC, Zhu H, Zhang BH, Li G, Zhu HK, Ren Y, et al. Tuning the kinetics of zinc-ion insertion/extraction in V2O5 by in situ polyaniline intercalation enables improved aqueous zinc-ion storage performance. Adv Mater. 2020;32(26):2001113.10.1002/adma.202001113Search in Google Scholar PubMed

[21] He P, Zhang GB, Liao XB, Yan MY, Xu X, An QY, et al. Sodium ion stabilized vanadium oxide nanowire cathode for high-performance zinc-ion batteries. Adv Energy Mater. 2018;8(10):1702463.10.1002/aenm.201702463Search in Google Scholar

[22] Yang YQ, Tang Y, Fang GZ, Shan LT, Guo JS, Zhang WY, et al. Li+ intercalated V2O5·nH2O with enlarged layer spacing and fast ion diffusion as an aqueous zinc-ion battery cathode. Energy Environ Sci. 2018;11(11):3157–62.10.1039/C8EE01651HSearch in Google Scholar

[23] Shan LT, Wang YR, Liang SQ, Tang BY, Yang YQ, Wang ZQ, et al. Interfacial adsorption-insertion mechanism induced by phase boundary toward better aqueous Zn-ion battery. InfoMat. 2021;3(9):1028–36.10.1002/inf2.12223Search in Google Scholar

[24] Hao Y, Zhang SM, Tao P, Shen T, Huang ZJ, Yan JK, et al. Pillaring effect of K ion anchoring for stable V2O5-based zinc ion battery cathode. ChemNanoMat. 2020;6(5):797–805.10.1002/cnma.202000105Search in Google Scholar

[25] Xia C, Guo J, Li P, Zhang XX, Alshareef HN. Highly stable aqueous zinc-ion storage using a layered calcium vanadium oxide bronze cathode. Angew Chem Int Ed. 2018;57(15):3943–8.10.1002/anie.201713291Search in Google Scholar PubMed

[26] Kundu D, Adams BD, Duffort V, Vajargah SH, Nazar LF. A high-capacity and long-life aqueous rechargeable zinc battery using a metal oxide intercalation cathode. Nat Energy. 2016;1(10):16119.10.1038/nenergy.2016.119Search in Google Scholar

[27] Shan LT, Zhou J, Han MM, Fang GZ, Cao XX, Wu XW, et al. Reversible Zn-driven reduction displacement reaction in aqueous zinc-ion battery. J Mater Chem A. 2019;7(13):7355–9.10.1039/C9TA00125ESearch in Google Scholar

[28] Chen LL, Yang ZH, Wu J, Chen HZ, Meng JL. Energy storage performance and mechanism of the novel copper pyrovanadate Cu3V2O7(OH)2·2H2O cathode for aqueous zinc ion batteries. Electrochim Acta. 2020;330:135347.10.1016/j.electacta.2019.135347Search in Google Scholar

[29] Soundharrajan V, Sambandam B, Kim SJ, Alfaruqi MH, Putro DY, Jo J, et al. Na2V6O16·3H2O barnesite nanorod: an open door to display a stable and high energy for aqueous rechargeable Zn-ion batteries as cathodes. Nano Lett. 2018;18(4):2402–10.10.1021/acs.nanolett.7b05403Search in Google Scholar PubMed

[30] Liu X, Zhang H, Geiger D, Han J, Varzi A, Kaiser U, et al. Calcium vanadate sub-microfibers as highly reversible host cathode material for aqueous zinc-ion batteries. Chem Commun. 2019;55(16):2265–8.10.1039/C8CC07243DSearch in Google Scholar

[31] Zhao HN, Fu Q, Yang D, Sarapulova A, Pang Q, Meng Y, et al. In operando synchrotron studies of NH4+ preintercalated V2O5·nH2O nanobelts as the cathode material for aqueous rechargeable zinc batteries. ACS Nano. 2020;14(9):11809–20.10.1021/acsnano.0c04669Search in Google Scholar PubMed

[32] Sambandam B, Soundharrajan V, Kim SG, Alfaruqi MH, Jo J, Kim SH, et al. K2V6O16·2.7H2O nanorod cathode: an advanced intercalation system for high energy aqueous rechargeable Zn-ion batteries. J Mater Chem A. 2018;6(32):15530–9.10.1039/C8TA02018CSearch in Google Scholar

[33] Li Z, Wu BK, Yan MY, He L, Xu L, Zhang GB, et al. Novel charging-optimized cathode for a fast and high-capacity zinc-ion battery. ACS Appl Mater Interfaces. 2020;12(9):10420–7.10.1021/acsami.9b21579Search in Google Scholar PubMed

[34] Xu L, Zhang YF, Jiang HM, Zheng JQ, Dong XY, Hu T, et al. Facile hydrothermal synthesis and electrochemical properties of (NH4)2V6O16 nanobelts for aqueous rechargeable zinc ion batteries. Colloids Surf A. 2020;593:124621.10.1016/j.colsurfa.2020.124621Search in Google Scholar

[35] Tang BY, Fang GZ, Zhou J, Wang LB, Lei YP, Wang C, et al. Potassium vanadates with stable structure and fast ion diffusion channel as cathode for rechargeable aqueous zinc-ion batteries. Nano Energy. 2018;51:579–87.10.1016/j.nanoen.2018.07.014Search in Google Scholar

[36] Guo C, Wang Q, He J, Wu C, Xie K, Liu Y, et al. Rational design of unique ZnO/ZnS@N-C heterostructures for high performance lithium-ion batteries. J Power Sources. 2021;482:228915. J Phys Chem Lett. 2020;11:905–12.10.1021/acs.jpclett.9b03677Search in Google Scholar PubMed

[37] Deng SZ, Yuan ZS, Tie ZW, Wang CD, Song L, Niu ZQ. Electrochemically induced metal-organic-framework-derived amorphous V2O5 for superior rate aqueous zinc-ion batteries. Angew Chem Int Ed. 2020;59(49):22002–6.10.1002/anie.202010287Search in Google Scholar PubMed

[38] Chen LL, Yang ZH, Huang YG. Monoclinic VO2(D) hollow nanospheres with super-long cycle life for aqueous zinc ion batteries. Nanoscale. 2019;11(27):13032–9.10.1039/C9NR03129DSearch in Google Scholar PubMed

[39] Tang F, Gao JY, Ruan QY, Wu XW, Wu XS, Zhang T, et al. Graphene-wrapped MnO/C composites by MOFs-derived as cathode material for aqueous zinc ion batteries. Electrochim Acta. 2020;353:136570.10.1016/j.electacta.2020.136570Search in Google Scholar

[40] Chen HD, Huang JJ, Tian SH, Liu L, Qin TF, Song L, et al. Interlayer modification of pseudocapacitive vanadium oxide and Zn(H2O)n2+ migration regulation for ultrahigh rate and durable aqueous zinc-ion batteries. Adv Sci. 2021;8(14):2004924.10.1002/advs.202004924Search in Google Scholar PubMed PubMed Central

[41] Pang Q, He W, Yu XY, Yang SY, Zhao HN, Fu Y, et al. Aluminium pre-intercalated orthorhombic V2O5 as high-performance cathode material for aqueous zinc-ion batteries. Appl Surf Sci. 2021;538:148043.10.1016/j.apsusc.2020.148043Search in Google Scholar

[42] Zhao X, Zhao TK, Peng XR, Yang L, Shu Y, Jiang T, et al. In situ synthesis of expanded graphite embedded with amorphous carbon-coated aluminum particles as anode materials for lithium-ion batteries. Nanotechnol Rev. 2020;9(1):436–44.10.1515/ntrev-2020-0033Search in Google Scholar

[43] Hu H, Hu YB, Cheng HY, Dai SG, Song KX, Liu ML. Organic polysulfanes grafted on porous graphene as an electrode for high-performance lithium organosulfur batteries. J Power Sources. 2021;491:229617.10.1016/j.jpowsour.2021.229617Search in Google Scholar

Received: 2022-01-10
Revised: 2022-02-22
Accepted: 2022-03-16
Published Online: 2022-04-18

© 2022 Haoyan Cheng et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 12.6.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2022-0103/html
Scroll to top button