Highly reversible Zn electrodeposition enabled by glutathione-protected copper nanoclusters for aqueous Zn-ion batteries

Lu-Fan Wang , Xiao-Yuan Wang , Shu Xu , Zhou Wu , Guoqiang Sun *, Xiao-Fei Liu *, Ren-Wu Huang and Shuang-Quan Zang
Henan Key Laboratory of Crystalline Molecular Functional Materials, College of Chemistry, Zhengzhou University, Zhengzhou 450001, Henan, China. E-mail: sungq@zzu.edu.cn; liuxiaofei@zzu.edu.cn

Received 16th October 2024 , Accepted 26th November 2024

First published on 27th November 2024


Abstract

Dendrite growth and side reactions occurring in Zn anodes hinder the development and practical application of aqueous Zn-ion batteries. Herein, for the first time, we demonstrated the use of ligand-protected Cu nanoclusters as Zn plating/stripping mediators to construct dendrite-free Zn anodes. The abundant polar functional groups (–COOH, –NH2, and –CO–NH–) on the peripheral ligands modulated the solvation structure and regulated the transport of Zn2+, whereas the zincophilic Cu cores functioned as nucleating agents to guide the uniform nucleation and plating of Zn. Using such bifunctional CuNCs, the asymmetric cell achieves a coulombic efficiency of 99.40% for over 1000 cycles, and the symmetric cell exhibits low voltage hysteresis and superior cycling over 400 h with a high depth of discharge of 40%. These findings may contribute to the design of multifunctional mediators at the atomic level.


Introduction

Aqueous Zn-ion batteries (AZIBs) have garnered extensive attention in the energy-storage field owing to their distinct benefits, including low cost, high theoretical capacity (820 mA h g−1, 5855 mA h cm−3), low redox potential (−0.76 V vs. the standard hydrogen electrode (SHE)), and inherent safety.1–5 Nevertheless, the uncontrollable growth of Zn dendrites, which widens the electrode–electrolyte contact area and exacerbates detrimental side reactions on the Zn anode surface, remains a thorny issue.6,7 Various approaches, including surface modification,8–13 use of electrolyte additives,14–17 separator optimization,18,19 and host design,20–25 have been reported as solutions to these issues. Among these approaches, building three-dimensional current collectors is a feasible strategy to efficiently restrain the growth of Zn dendrites compared to conventional planar Zn anodes. This method reduces the local current density and homogenizes the electrical field on the anode surface, while mitigating volume variations during Zn plating/stripping, thereby enhancing electrode stability.26

Owing to their high specific surface area, ultralight design, and unique vertical–horizontal network structure, carbon cloth (CC) current collector offers inherent advantages for future advancements in Zn metal anodes and flexible energy-storage applications. However, the absence of effective zincophilic sites results in a high nucleation barrier for Zn2+ ions on naked CC.27 Our group doped CC with heteroatoms to increase the number of zincophilic sites, which can induce exposure of the Zn (002) crystal plane and significantly enhance the cell performance.28 Liang et al.29 prepared an amine-functionalized CC host. The introduced amine groups enhanced the binding energy of Zn on the CC surface and directed the nucleation and growth of Zn. Dong et al.30 revealed that a layer of Cu nanosheets on CC can considerably reduce the Zn nucleation overpotential and provide a large number of evenly spaced Zn deposition sites. Most CC modification strategies focus on the modulation of their zincophilic ability while neglecting the optimization of the solvated structure of hydrated Zn2+ ions transported on their surface. Notably, the desolvation of Zn(H2O)62+ has a high energy barrier, resulting in the inhomogeneous deposition of metallic Zn. Adjusting the coordination environment of Zn2+ ions can effectively regulate the solvation structure and chemical properties of hydrated Zn2+ ions, thereby inhibiting dendrites growth and side reactions.31 Thus, the construction of bifunctional modulators with high zincophilic ability and regulation of the solvated structure of Zn2+ ions on the CC surface is an effective method to further enhance the performance of current collectors.

Metal nanoclusters (NCs) are a type of nanomaterial sized 1–3 nm with a metal core and a surface modified by different organic ligands. They have a monodisperse nature and precise atomic structure, facilitating the analysis of the action site and mechanism during application. In addition, ligand-protected NCs have good stability.32–35 By selecting different metals and organic ligands with characteristic functional groups to regulate the synthesis, the target clusters can be obtained more accurately and quickly.36 Consequently, NCs are widely used in electrochemistry owing to their unique advantages. For example, using CuNCs as precursors, we have synthesized zinc-metal-free anode materials for high-performance AZIBs by precisely controlling the pyrolysis process of the NCs.37 Beyond that, Wang et al. first proposed the improved performance of organic-covered metal NCs as active catalysts than those with partially or completely removed surface ligands.38 With reasonable ligand selection and synthesis regulation, even intact metal NCs can be directly applied to electrochemistry to achieve unexpected results through the synergistic effect of ligands and cores.

In this paper, for the first time, we present a new strategy to construct stable dendrite-free Zn anodes using a CC substrate modified with glutathione (GSH)-protected CuNCs (GSH-CuNCs) (Fig. 1a). The polar functional groups (–COOH, –NH2, and –CO–NH–) on the peripheral GSH ligand have high binding energy with Zn2+ to promote the desolvation of Zn(H2O)62+ and accelerate the Zn2+ transport kinetics. In addition, the Cu cores can be regarded as Zn-deposited seeds that guide the homogeneous nucleation and growth of Zn on the CC substrate. The asymmetric cell employing CuNCs achieved a coulombic efficiency (CE) of 99.40% after 1000 stable plating/stripping cycles at 1.0 mA cm−2. Meanwhile, the lifetime of the symmetric cell with CuNCs-modified current collector was improved by more than 500 h, which is 12 times higher than that of pristine CC (PCC) (Fig. 1b; Fig. S1, ESI). This initial application of NCs in a dendrite-free Zn anode design provides a reference for high-performance electrode modification and promotes the development of NCs for energy-storage devices.


image file: d4qi02605e-f1.tif
Fig. 1 (a) Preparation process diagram of NCC and NCC@CuNCs. (b) The comparison histogram of cycle times of symmetric cells based on three different scaffolds at a current density 1 mA cm−2 and an areal capacity of 1 mA h cm−2.

Experimental section

Experimental materials

CC was purchased from Keshenghe (Suzhou, China). L-Glutathione (L-GSH, 99%) was purchased from J&K Scientific. Sodium hydroxide (≥96.0%) was purchased from Tianjin Fengchuan chemical reagent technology Co., Ltd. Zinc sulfate heptahydrate (AR) was purchased from Aladdin. Acetone (≥99.5%), ammonia solution (25%–28%), manganese sulfate monohydrate (99.99%), and potassium permanganate (≥99.5%) were purchased from Luoyang chemical reagent plant. Copper sulfate pentahydrate (AR, 99%) was purchased from Macklin. The above experimental materials are used directly for experiments without further purification.

Preparation of NCC@CuNCs

GSH-CuNCs were synthesized according to a previous report.39 PCC was placed in ammonia solution by adopting solvothermal method at 150 °C for 24 h for N-doped treatment, and then rinsed with ultra-pure water to PH = 7 and dried at 60 °C for 12 h to obtain nitrogen-doped carbon cloth (NCC). The prepared NCC was immersed in CuNCs solution for 12 h at 4 °C, then washed with ultra-pure water and dried to obtain NCC anchored with GSH-CuNCs (NCC@CuNCs).

Preparation of NCC@Zn and NCC@CuNCs@Zn anodes

The NCC@Zn and NCC@CuNCs@Zn were prepared by electrochemical deposition method on a LAND-CT2001A battery testing system. The CR2032 coin-type battery was assembled with 2 M ZnSO4 as electrolyte, zinc foil (diameter 12 mm and thickness 100 μm) and NCC/NCC@CuNCs of the same size were employed as counter and working electrodes, respectively. After that, metallic zinc was electrodeposited on CC at current density of 5 mA cm−2 for 1 h with a deposition capacity of 5 mA h cm−2. Finally, the working electrodes were removed from the coin-type cells as well as cleaned with ultra-pure water and dried at 60 °C to obtain NCC@Zn and NCC@CuNCs@Zn electrodes.

Preparation of MnO2 cathode

0.1 M KMnO4 solution and 0.6 M MnSO4 solution were prepared and mixed in equal volumes, stirred magnetically for 10 minutes, then transferred to the polytetrafluoroethylene (PTFE) lining of stainless steel reactor and reacted in an oven at 140 °C for 12 h. Next, the reaction liquid was centrifuged, and washed with ultrapure water repeatedly. Finally, it was dried at 60 °C to obtain β-MnO2. β-MnO2, acetylene black and polyvinylidene fluoride were mixed according to the mass ratio of 7[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1, and the appropriate amount of 1-methyl-2-pyrrolidinone was added to fully grind, then uniformly coated on the stainless steel mesh by doctor blade method, and the MnO2 cathode was prepared by vacuum drying at 60 °C for 12 h. Finally, the mass loading of active material was about 1.5 mg cm−2.

Material characterization

The morphology and microstructure of the samples were characterized by scanning electron microscope (SEM) (Zeiss Sigma 500, German) equipped with an energy-dispersive X-ray spectroscopy (EDS) detector. X-ray diffraction patterns (XRD) were recorded with Rigaku MiniFlex600 diffractometer equipment using Cu Kα radiation (k = 1.5418 Å) at 30 kV to analyze the crystal structure of the products. The surface composition and content of the electrodes were characterized by X-ray photoelectron spectroscopy (XPS) using Thermo Scientific K-Alpha photoelectron spectrometer with monochromatic Al Kα (1486.6 eV) radiation. The hydrophilicity difference of electrodes was measured by Chengde Dingsheng JY-82C video contact Angle tester. The surface defect degree of CC was characterized by Thermo Scientific DXR 3Xi Raman spectrometer. At room temperature, Fourier Transform Infrared Spectroscopy (FT-IR) was recorded by ALPHA II infrared spectrometer in the range of 4000–500 cm−1. Transmission electron microscopy (TEM) images were captured by Japan-JEOL-JEM 2100 F. The ultraviolet-visible (UV-vis) spectra were collected on UV-2700.

Electrochemical characterization

The cycle reversibility and stability of different scaffolds were studied by means of symmetric cells, which were assembled with NCC/NCC@CuNCs as the working electrode and Zn foil as the counter electrode. In order to evaluate the electroplating/stripping efficiency of the anodes, asymmetric batteries were assembled with NCC/NCC@CuNCs as the working electrode and Zn foil as the counter electrode. The full cells were assembled using an as-prepared β-MnO2 cathode and NCC@Zn/NCC@CuNCs@Zn anode, respectively. All the above components were packaged in the CR2032 coin-type battery case, using glass fiber as the diaphragm, 2 M ZnSO4 as the electrolyte for symmetric and asymmetric batteries, and 2 M ZnSO4 + 0.1 M MnSO4 as the electrolyte for full batteries. Using the LAND-CT2001A battery test system, the batteries were tested at 30 °C, including long cycle performance, nucleation overpotential, rate, CE and discharge specific capacity.

At room temperature, the electrochemical performance tests were recorded through electrochemical workstation (CHI760E). Electrochemical impedance spectroscopy (EIS) was measured at frequencies ranging from 0.01 Hz to 100 kHz. Cyclic voltammetry curves (CV) were collected at a scan rate of 0.1 mV s−1. The diffusion curves were tests via it curves at −150 mV overpotential. The hydrogen evolution reaction (HER) performance was measured by conducting linear sweep voltammetry (LSV) curves, where NCC@Zn or NCC@CuNCs@Zn was used as the working electrode and the Pt foil and Ag/AgCl electrode were used as the counterpart and the reference electrode, respectively. The corrosion property was also measured by means of LSV, the difference was that a two-electrode system was used, where NCC@Zn or NCC@CuNCs@Zn as the working electrode and Zn foil as the counter electrode.

Density functional theory (DFT) calculations

The binding energy of between Zn2+ and H2O molecule or GSH ligand respectively were calculated with DMol3 module by Materials Studio. The exchange–correlation functional was performed by generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE). The force convergence criteria, energy criteria for geometry optimization and energy criteria for SCF convergence were 0.002 Ha Å−1, 1.0 × 10−5 Ha and 1.0 × 10−6 Ha, respectively. The wave function was fitted by DNP basis set. The core electrons were processed by Effective Core Potential (ECP) method. The orbittal cutoff radius was 4.4 Å. The binding energy (Eb) were calculated as the following equation:
Ebinding = EABEAEB
where EAB represents the total energy of H2O or Zn2+ combined with GSH, EA is the energy of H2O or Zn2+, and EB is the energy of GSH.

Results and discussion

Preparation and characterization of GSH-CuNCs and different hosts

GSH-CuNCs were synthesized using a facile one-pot method. The UV-vis absorption spectrum of the obtained CuNCs solution exhibits an absorption peak at 250 nm (Fig. 2a), indicating the successful formation of CuNCs.39 The solution appeared light green under natural light and bright blue under UV light (inset in Fig. 2a). The dispersion state of the GSH-CuNCs in the aqueous solution was observed by TEM. The prepared GSH-CuNCs are uniformly distributed in a spherical shape (Fig. 2b).
image file: d4qi02605e-f2.tif
Fig. 2 (a) UV-vis absorption spectrum of CuNCs (inset: digital photos of the CuNCs solution under natural (the left) and UV light (the right)). (b) TEM image of CuNCs. (c) XPS high-resolution N 1s spectra of PCC and NCC. (d) XPS high-resolution Cu 2p spectrum of NCC@CuNCs. (e) TEM image of NCC@CuNCs. (f) UV-vis image of solution of NCC@CuNCs.

In order to improve the hydrophilicity and enhance the adsorption capacity of the CC substrate for water-soluble GSH-CuNCs, the PCC was hydrothermally treated with ammonia solution to obtain NCC.40–42

As shown in Fig. S2a (ESI), both PCC and NCC own smooth surfaces, indicating that the morphology of the CC remains essentially unchanged after N doping. The corresponding XRD spectra also confirm this finding (Fig. S2b, ESI). The elemental mapping images depict the uniform distribution of C and N elements in the NCC (Fig. S3, ESI), proving the successful N doping into the CC substrate. Meanwhile, the peak areas of the N 1s XPS spectra of CC were compared, as shown in Fig. 2c. NCC has a large N 1s peak area, illustrating its significantly higher N content (approximately 2.17 at%) than that of PCC. The high-resolution N 1s spectra of PCC and NCC can be deconvoluted into three main peaks at 397.9, 399.9, and 402.5 eV, corresponding to pyridinic N, pyrrole N, and graphite N, respectively (Fig. S4, ESI).43 The pyrrole N content of NCC is determined to be 67.1% (Table S1, ESI), elucidating the abundant defects in CC after N doping. This was also verified by Raman spectroscopy (Fig. S5, ESI), where the D peak (∼1350 cm−1) represents the defect degree of disordered C materials and the G peak (∼1580 cm−1) is the main characteristic peak of graphene owing to the in-plane vibration of C atoms; thus, the peak strength ratio (ID/IG) reflects the degree of structural disorder of C materials.44 The ID/IG ratios of PCC and NCC are calculated to be 1.04 and 1.14, respectively, demonstrating the increased disorder and number of defects in CC after N doping, which is conducive to anchoring CuNCs.

After that, NCC was immersed in CuNCs solution to obtain NCC@CuNCs. XPS was performed to explain its successful construction and analyze the surface composition. The XPS spectra of NCC@CuNCs exhibit constituent element peaks of CuNCs (Fig. 2d; Fig. S6, ESI). In particular, the satellite peak at approximately 942 eV is absent in the Cu 2p XPS spectrum, which is consistent with the XPS results of previously reported CuNCs.39 The EDS elemental mapping analysis shows homogeneous distribution of C, N, S, O, and Cu elements on the NCC host (Fig. S7, ESI), illustrating that there is no obvious agglomeration of the GSH-CuNCs adsorbed on NCC. In order to further illustrate the successful anchoring of CuNCs to the NCC substrate, TEM analysis of NCC@CuNCs was conducted. As expected, small spotted CuNCs are uniformly distributed on the NCC fibers (Fig. 2e), and conspicuous lattice fringes are observed in the high-resolution TEM images. The lattice fringes with a lattice plane distance of 2.07 Å corresponds to the Cu (111) crystal plane (JCPDS 89-2838; Fig. S8, ESI).45 Subsequently, NCC@CuNCs was ultrasonically treated in ultrapure water. The UV-vis image of the solution exhibits CuNCs absorption peak (Fig. 2f). These characterizations indicate the successful preparation of the NCC and NCC@CuNCs substrates. What's more, the properties of the CuNCs anchored to the NCC are still maintained.

Zn plating/stripping performance of different substrates

The degree of wetting at the interface of the electrolyte and electrode indicates the uniformity of the charge distribution, which directly affects the Zn deposition on the anode surface.46 Therefore, we separately investigated the contact angles of 2 M ZnSO4 on the surfaces of the PCC, NCC, and NCC@CuNCs substrates (Fig. S9, ESI). NCC@CuNCs displays a contact angle of 112.13°, which is smaller than those of NCC (115.42°) and PCC (125.07°) in the initial state. Although the contact angles of three samples all mitigate after standing for 5 and 10 min respectively, NCC@CuNCs remains the smallest. This can be interpreted as the presence of CuNCs to improve electrode hydrophilicity and electrolyte accessibility, which facilitates a uniform Zn2+ flux and decreases the interfacial free energy between NCC@CuNCs and the electrolyte, thereby promoting the rapid diffusion and uniform nucleation of Zn2+ on the CC surface.47 In order to verify the stability of CuNCs on the substrate during the charge–discharge cycle, 5 mA h cm−2 Zn was first deposited on NCC@CuNCs substrate and then stripped at equal capacity, then the UV image of the solution after ultrasound maintains the characteristic absorption peak of CuNCs (Fig. S10, ESI), suggesting the stable CuNCs loading on the CC during the charge–discharge process.

CE is a crucial parameter for evaluating the reversibility of Zn anodes.24,48 We assembled Zn//NCC and Zn//NCC@CuNCs asymmetric cells to probe the Zn plating/stripping behavior on different substrates. During the CE test, Zn was plated on the CC hosts by discharging at various current densities (1, 3, and 5 mA cm−2) for 1 h and then stripped to a cutoff voltage of 0.5 V at the same current density for each cycle. As shown in Fig. 3a, the Zn//NCC@CuNCs cell displays an average CE of ∼99.40% after 1000 cycles at 1 mA cm−2 with an areal capacity of 1 mA h cm−2. This cycling life is approximately 20 times longer than that of the Zn//NCC cell which undergoes short circuit after only 50 cycles. Besides, NCC@CuNCs exhibits a lower nucleation overpotential (NOP) of 128.2 mV at 2 mA cm−2 compared to 358.5 mV of NCC (Fig. 3b), which can be attributed to the CuNCs that guarantee homogeneous Zn nucleation and deposition on the CC surface. Zn//NCC@CuNCs exhibits an initial CE of 89.50% as the current density is increased to 3 mA cm−2, outperforming Zn//NCC (86.90%). In addition, the Zn//NCC@CuNCs cell (110 cycles) has a longer lifespan than Zn//NCC (37 cycles) even at a current density of 5 mA cm−2 (Fig. 3c). The performance of the asymmetric cells is more competitive with the reported AZIB hosts (Table S2, ESI).20,21,24,26,49–52 The PCC was immersed in CuNCs solution to obtain PCC@CuNCs for demonstrating the need for a defect-rich host to anchor CuNCs. The cycle life of the corresponding asymmetric cells at three current densities is longer than that of the NCC but less than that of NCC@CuNCs (Fig. S11, ESI). This not only shows the advantages of CuNCs, but also emphasizes the importance of proper amount of anchoring to improve the performance of AZIBs.


image file: d4qi02605e-f3.tif
Fig. 3 Asymmetric cells (a) CE plots of Zn plating/stripping on NCC and NCC@CuNCs at 1 mA cm−2 and 1 mA h cm−2. (b) Nucleation overpotentials of Zn deposition on different current collectors at a current density of 2 mA cm−2 and areal capacity of 2 mA h cm−2. (c) CE plots of Zn plating/stripping on NCC and NCC@CuNCs electrodes at different current densities. SEM morphologies of (d) NCC@CuNCs and (e) NCC after the first cycle under different current densities. (f) Schematic diagrams of Zn deposition on the two kinds of substrates.

Furthermore, we fetched the electrodes after the first cycle under different current densities and areal capacities to explore the surface morphology difference of the substrates using SEM. As illustrated in Fig. 3d, the smooth CC fibers of the NCC@CuNCs indicate the almost complete stripping of Zn at different current densities owing to substrates enhanced cyclic reversibility. Conversely, there are numerous irregular Zn nanosheets stacked and exhibited a vertical distribution, forming bumps on the electrode surface (Fig. 3e).

More impressively, the electric field at these bumps is unevenly distributed, resulting in a “tip effect”. Consequently, an increasing number of Zn are attracted to deposit where the current density is higher, aggravating Zn dendrite growth during subsequent deposition. Dendrite growth increases the contact area between the electrode and electrolyte, leading to serious side reactions that poses a low CE and eventual failure of Zn//NCC.53

Under a constant voltage of −150 mV, the Zn nucleation behavior on the surface of different substrates was further studied by chronoamperometry (CA). The deposition current density of the NCC@Zn electrode exhibits a monotonically increasing trend within 400 s owing to the two-dimensional diffusion of Zn2+ that formed Zn dendrites on the electrode surface. Meanwhile, the surface current of the NCC@CuNCs@Zn electrode increases in the first 50 s of Zn deposition and then remains constant, reflecting the three-dimensional diffusion characteristic of Zn2+. This manifests the uniform Zn deposition on the electrode surface without obvious Zn dendrite formation (Fig. S12, ESI).54Fig. 3f presents the nucleation and growth behaviors of Zn on the surfaces of two types of substrates. Due to the lack of nucleation sites, Zn deposition on NCC surface is uneven and covers with irregular dendrites after cycling. On the contrary, the presence of CuNCs induces uniform and dense deposition of Zn on the surface of the CC, resulting in dendrite-free NCC@CuNCs@Zn electrode.

Inhibited hydrogen evolution and corrosion side reactions by NCC@CuNCs substrate

The depth of discharge (DOD) is an important metric reflecting the performance of AZIBs, representing the Zn utilization in symmetric cells.55 A conventional thick Zn foil (thickness = 100 μm) is used to assemble a symmetric cell, which fails because of the short circuit after 80 h with the DOD of only 3.4% (eqn (1) and Fig. S13a, ESI), thereby significantly increasing the manufacturing cost and reducing the actual energy density of the battery. We selected a reasonable DOD of 40% (eqn (2), ESI) and used the predeposition method on NCC and NCC@CuNCs to prepare Zn metal anodes (the details are shown in the Experimental section), thus reducing the amount of excess Zn and improving the Zn utilization. The XRD patterns of NCC@Zn and NCC@CuNCs@Zn show obvious characteristic diffraction peaks of Zn respectively (Fig. S14, ESI), indicating the successful Zn deposition on CC. The long-term galvanostatic cycling test of the symmetric cell was carried out under 2 mA cm−2 and areal capacity of 2 mA h cm−2. Short circuit is noted on Zn//NCC after 70 h, which can be interpreted as the uncontrolled growth of Zn dendrites that eventually punctured the diaphragm. In comparison, Zn//NCC@CuNCs has a longer cycle life of over 400 h (Fig. 4a; Fig. S13b and S15, ESI). The polarization voltage of Zn//NCC fluctuates greatly in the initial cycle, immediately followed by a short circuit, while Zn//NCC@CuNCs maintains basically stable voltage profiles for nearly 300 h. Notably, the NCC@CuNCs@Zn anode exhibits better cycling performance than most previously reported anodes (Table S3, ESI).20,22,24,26,28,29,49–52,56–60 These results reveal the excellent cyclic reversibility of NCC@CuNCs@Zn, effectively inhibiting the Zn dendrites growth and side reactions under the guidance of the zincophilic Cu cores and polar functional groups. The long cycle life of symmetric cells based on the PCC@CuNCs substrate at current densities of 1 and 2 mA cm−2 is between that of NCC and NCC@CuNCs (Fig. S16, ESI) owing to the limited hydrophilicity of the electrodes, again highlighting the positive role of CuNCs in improving electrode dynamics.
image file: d4qi02605e-f4.tif
Fig. 4 (a) Cycling performance of symmetric cells Zn//NCC and Zn//NCC@CuNCs at 2 mA cm−2. (b) Rate test of NCC@Zn//Zn and NCC@CuNCs@Zn//Zn at different current densities. (c) LSV curves in 1 M aqueous Na2SO4 electrolyte. (d) Linear polarization curves of NCC@Zn and NCC@CuNCs@Zn. (e) The DFT calculations for binding energies of Zn2+ to H2O and GSH ligand. (f) Schematic diagram of Zn2+ transport and uniform deposition on the surface of NCC@CuNCs substrate.

The rate performance of a symmetric cell is the standard for measuring whether the uniform deposition of Zn2+ on an electrode during cycling. Therefore, symmetric cells were tested with a fixed areal capacity of 1 mA h cm−2 to investigate the Zn2+ deposition of the NCC and NCC@CuNCs substrates at current densities 1, 2, 4, 5, and 10 mA cm−2, respectively (Fig. 4b). According to the obtained voltage curves, NCC@CuNCs has a lower voltage hysteresis than NCC, which is more prominent at high current densities, indicating a lower Zn nucleation barrier on the NCC@CuNCs substrate and further revealing the improvement of the Zn electroplating/stripping kinetics. This conclusion is also confirmed by the smaller charge-transfer resistance (Rct) of Zn//NCC@CuNCs in the Nyquist plots of the impedances of the symmetric cells (Fig. S17, ESI).

In order to analyze the inhibitory effect of the NCC@Zn and NCC@CuNCs@Zn anodes on hydrogen evolution, LSV was performed at a scanning rate of 5 mV s−1 in a 1 M Na2SO4 aqueous solution. As there is no interference from the Zn deposition, the obtained current signal can accurately reflect the HER intensity. As shown in Fig. 4c, NCC@CuNCs@Zn (−1.189 V) exhibits a lower negative overpotential than NCC@Zn (−1.169 V) at a fixed current density of 10 mA cm−2, demonstrating the suppressed HER for NCC@CuNCs@Zn electrode.53 Inhibition of HER can reduce the OH formation near the anode, restricting the generation of byproducts at the electrode–electrolyte interface. Moreover, the corrosion behaviors of the substrates with and without loaded CuNCs in 2 M ZnSO4 electrolytes were detected using the Tafel curves at a sweep speed of 1 mV s−1. NCC@Zn exhibits a corrosion current density of 0.3238 mA cm−2, which is higher than that of NCC@CuNCs@Zn (0.3057 mA cm−2, (Fig. 4d)). A lower corrosion current density suggests a preferable corrosion resistance of the electrode.11 Hence, the relatively good corrosion resistance of NCC@CuNCs@Zn demonstrates the role of CuNCs introduction to inhibit the occurrence of side reactions.

The DFT calculation are shown in Fig. 4e, the binding energies of GSH ligand at five functional active sites (Fig. S18, ESI) with Zn2+ are −11.99, −12.04, −13.75, −13.02 and −14.41 eV respectively, which are much higher than that of Zn2+–H2O (−4.67 eV), implying that the peripheral GSH ligand of CuNCs can regulate the solvation structure of Zn2+ to promote the removal of the solvation sheath and increase Zn2+ flux.61 Consequently, the transport and deposition principles of Zn2+ on the NCC@CuNCs surface are legitimately inferred combined with the above CE and symmetric cell test results (Fig. 4f). In AZIBs, Zn2+ tends to solvate six water molecules to form Zn(H2O)62+ as well as first needs to be desolvated during the electrochemical reaction; therefore, the solvation structure is unfavorable for the charge and discharge behavior. Moreover, the water molecules existing in Zn(H2O)62+, which have higher electrochemical activity than the free water molecules, are preferentially decomposed into H+ and OH at the electrode surface, resulting in the HER and formation of an irreversible Zn4SO4(OH)6·xH2O by-product.62 Obviously, the removal of the solvated sheath is needed to improve the battery performance. On the one hand, the peripheral protective ligands of the CuNCs anchored on NCC have several polar functional groups, such as –COOH, –NH2, and –CO–NH–, which have a strong affinity for Zn2+, thus reducing the desolvation energy barrier of Zn2+. On the other hand, the rich hydrogen bonding network in the ligands can keep detrimental water out (Fig. S19, ESI), thus ultimately mitigating the interface polarization and significantly inhibiting the occurrence of anode dendrites and side reactions.63–65 Besides, the coulombic attraction force between the −COOH and Zn2+ and the complexation between Zn2+ and –CO–NH– groups, which accelerate its transfer to the zincophilic Cu cores of the CuNCs, guarantee fast Zn2+ transport kinetics and finally achieve homogenous and dense Zn deposition on the CC surface through the effective induction of the zincophilic copper cores.57,66

Verification of the performance improvement of AZIBs by CuNCs-based bifunctional anode

NCC was immersed in GSH and CuSO4 aqueous solutions to obtain NCC@GSH and NCC@CuSO4 for the symmetric and asymmetric cell tests, respectively, to fully demonstrate the advantages of CuNCs in improving the performance of Zn anodes. The XPS spectra of the corresponding elements demonstrate the successful loading of GSH (Fig. 5a; Fig. S20, ESI) and Cu2+ (Fig. 5b; Fig. S21, ESI) onto the NCC substrate. The symmetric cells based on NCC@GSH (150 h) and NCC@CuSO4 (220 h) at a current density of 1 mA cm−2 and areal capacity of 1 mA h cm−2 show obvious voltage fluctuation during the cycle (Fig. 5c), which is caused by the unstable chemical environment on the electrode surface. The cycle life of both is better than that of NCC but less than that of NCC@CuNCs. This finding is consistent with the results obtained at a current density of 2 mA cm−2 (areal capacity of 2 mA h cm−2; Fig. 5d). Unsurprisingly, the performance of the Zn//NCC@GSH and Zn//NCC@CuSO4 asymmetric cells under the three current densities also basically verifies the above conclusions (Fig. 5e and f; Fig. S22, ESI). Hence, although a single organic functional group or zincophilic active site can improve the performance, the extent of improvement is still limited and only the collaboration of both can better get rid of the dilemma faced by Zn anode. The NCC scaffold loaded with CuNCs possessed the aforementioned dual-modification advantages, revealing its tremendous application potential for AZIBs.
image file: d4qi02605e-f5.tif
Fig. 5 XPS high-resolution (a) S 2p spectrum of NCC@GSH, (b) Cu 2p spectrum of NCC@CuSO4. Cycling performance of Zn//NCC@GSH and Zn//NCC@CuSO4 symmetric cells at (c) 1 mA cm−2, (d) 2 mA cm−2. CE diagrams of (e) Zn//NCC@GSH, (f) Zn//NCC@CuSO4 at 1 mA cm−2 and an areal capacity of 1 mA h cm−2.

Electrochemical performance of the full cell

To evaluate the practical application value of the anodes, the NCC@CuNCs@Zn anodes were coupled with β-MnO2 and 2 M ZnSO4 + 0.1 M MnSO4 as the electrolyte to assemble full cells. MnSO4 was added to prevent the disproportionation of the MnO2 cathode.67 The XRD pattern shows typical β-MnO2 peaks (PDF#24-0735; Fig. 6a), and the nanorod-like morphology of β-MnO2 is verified by SEM (Fig. S23, ESI). The XRD and SEM images of the anodes after 100 cycles at 1 A g−1 were compared to further demonstrate the ability of the NCC@CuNCs@Zn anode to inhibit side reactions during cycling. As shown in Fig. 6b, the Zn4SO4(OH)6·5H2O by-product on the electrode surface of NCC@Zn has a higher diffraction peak, as well as the massive irregular particles on the electrode surface can be clearly seen in the SEM image. In contrast, the NCC@CuNCs@Zn anode shows a weak by-product peak in the XRD spectra and flat anode surface after cycling, which adequately indicates that the NCC@CuNCs@Zn anode can inhibit the HER and the generation of by-products.
image file: d4qi02605e-f6.tif
Fig. 6 (a) XRD pattern of β-MnO2. (b) XRD and SEM images (NCC@CuNCs@Zn: upper right, NCC@Zn: lower right) of different electrodes after 100 cycles of the full cells. The EIS comparison diagrams of (c) NCC@Zn//MnO2, (d) NCC@CuNCs@Zn//MnO2 full cells before and after 60 cycles. (e) Cycling performance at 1 A g−1 of the full cells. (f) Rate performance test of the full cells.

The CV measurements exhibit typical and identical β-MnO2 redox peaks (Fig. S24a, ESI), indicating that the existence of CuNCs does not affect the Zn//MnO2 redox reaction kinetics. Meanwhile, the charge/discharge curves also show similar charge and discharge platforms (Fig. S24b, ESI), in agreement with the CV curves.68 The Rct of the full cell was compared before and after 60 cycles. The Rct for NCC@Zn//MnO2 increases significantly after 60 cycles (Fig. 6c), whereas that of NCC@CuNCs@Zn//MnO2 increases from ∼50 Ω to ∼150 Ω after cycling (Fig. 6d). According to these results, it is not difficult to find that the NCC@CuNCs@Zn anode constructed based on CuNCs exhibits better Zn electroplating/stripping kinetics and electrochemical stability.69

The cycling performance of the full cells is shown in Fig. 6e. NCC@CuNCs@Zn//MnO2 possesses a higher specific capacity (61.4 mA h g−1) than NCC@Zn//MnO2 (13.9 mA h g−1) after approximately 700 cycles, and the capacity of NCC@Zn//MnO2 rapidly declines after 481 cycles. Fig. 6f displays the rate performance of the full cells. NCC@CuNCs@Zn//MnO2 delivers higher average specific capacities of 338.1, 367.0, 279.0, 186.0, 110.6, and 56.4 mA h g−1 at 0.1, 0.2, 0.5, 1.0, 2.0, and 5.0 A g−1, respectively. When the current density is reverted to 0.1 A g−1, NCC@CuNCs@Zn//MnO2 maintains a high specific capacity of 367.1 mA h g−1, whereas that of NCC@Zn//MnO2 sharply attenuates to 60.81 mA h g−1. Therefore, the GSH-CuNCs introduced during the modification of Zn-free anodes inhibit dendrite growth and side reactions, greatly improving the overall electrochemical performance of the full cells.

Conclusions

In conclusion, a novel three-dimensional dendrite-free Zn metal anode is constructed by anchoring CuNCs protected by organic ligands onto a hydrophilic CC substrate through simple immersion to effectively address the problems of hydrogen evolution and corrosion due to Zn dendrite growth. The polar functional groups (–COOH, –NH2, and –CO–NH–) on the protective ligands of the outer layer of the CuNCs are conducive to the dissolution of Zn(H2O)62+ and provide good zincophilic sites to accelerate the transport of Zn2+ to the cores of the CuNCs through coulombic attraction. Simultaneously, the zincophilic Cu cores can guide uniform Zn nucleation and deposition on the CC surface, thereby inhibiting dendrite growth. Compared with untreated NCC, NCC with CuNCs exhibits impressive cycling stability in the symmetric cell test and a CE of up to 99.40% in the asymmetric cell after 1000 cycles at 1 mA cm−2. In addition, the NCC@CuNCs@Zn//MnO2 full cell shows higher discharge capacity. Therefore, this study provides a new approach for modifying Zn anodes and confirms the development potential of metal NCs in AZIBs design.

Data availability

The data that supports the findings of this study are available in the article and ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 22209154), and Zhengzhou University.

References

  1. X. Qian, L. Li, Y. Wang, Z. Tian, H. Zhong, W. Chen, T. Chen and J. Fu, Versatile metallo-supramolecular polymeric interphase for highly reversible zinc metal anodes, Energy Storage Mater., 2023, 58, 204–214 CrossRef.
  2. C. Li, Q. Shao, K. Luo, Y. Gao, W. Zhao, N. Cao, S. Du, X. Jin, P. Zou and X. Zang, A lean-zinc and zincophilic anode for highly reversible zinc metal batteries, Adv. Funct. Mater., 2023, 33, 2305204 CrossRef CAS.
  3. X. Yang, Z. Dong, G. Weng, Y. Su, J. Huang, H. Chai, Y. Zhang, K. Wu, J. B. Baek, J. Sun, D. Chao, H. Liu, S. Dou and C. Wu, Crystallographic manipulation strategies toward reversible Zn anode with orientational deposition, Adv. Energy Mater., 2024, 14, 2401293 CrossRef CAS.
  4. J. N. Hao, B. Li, X. L. Li, X. H. Zeng, S. L. Zhang, F. H. Yang, S. L. Liu, D. Li, C. Wu and Z. P. Guo, An In-depth study of Zn metal surface chemistry for advanced aqueous Zn-ion batteries, Adv. Mater., 2020, 32, 2003021 CrossRef CAS PubMed.
  5. X. Jin, L. Song, C. Dai, Y. Xiao, Y. Han, X. Li, Y. Wang, J. Zhang, Y. Zhao, Z. Zhang, N. Chen, L. Jiang and L. Qu, A flexible aqueous zinc–iodine microbattery with unprecedented energy density, Adv. Mater., 2022, 34, 2109450 CrossRef CAS PubMed.
  6. Q. Zhang, J. Luan, Y. Tang, X. Ji and H. Wang, Interfacial design of dendrite-free zinc anodes for aqueous zinc-ion batteries, Angew. Chem., Int. Ed., 2020, 59, 13180–13191 CrossRef CAS PubMed.
  7. Z. Wang, J. Huang, Z. Guo, X. Dong, Y. Liu, Y. Wang and Y. Xia, A Metal-organic framework host for highly reversible dendrite-free zinc metal anodes, Joule, 2019, 3, 1289–1300 CrossRef CAS.
  8. L. Kang, M. Cui, F. Jiang, Y. Gao, H. Luo, J. Liu, W. Liang and C. Zhi, Nanoporous CaCO3 coatings enabled uniform Zn stripping/plating for long-life zinc rechargeable aqueous batteries, Adv. Energy Mater., 2018, 8, 1801090 CrossRef.
  9. J. Hao, X. Li, S. Zhang, F. Yang, X. Zeng, S. Zhang, G. Bo, C. Wang and Z. Guo, Designing dendrite-free zinc anodes for advanced aqueous zinc batteries, Adv. Funct. Mater., 2020, 30, 2001263 CrossRef CAS.
  10. J. Heo, Y. E. Hwang, G. Doo, J. Jung, K. Shin, D. Y. Koh and H. T. Kim, Modulation of solvation structure and electrode work function by an ultrathin layer of polymer of intrinsic microporosity in zinc ion batteries, Small, 2022, 18, 2201163 CrossRef CAS PubMed.
  11. S. Li, J. Fu, G. Miao, S. Wang, W. Zhao, Z. Wu, Y. Zhang and X. Yang, Toward planar and dendrite-free Zn electrodepositions by regulating Sn-crystal textured surface, Adv. Mater., 2021, 33, 2008424 CrossRef CAS PubMed.
  12. G. Sun, M. Zhou, X.-Y. Dong, S.-Q. Zang and L. Qu, An efficient and versatile biopolishing strategy to construct high performance zinc anode, Nano Res., 2022, 15, 5081–5088 CrossRef CAS.
  13. F. Zhang, J.-W. Qian, W.-X. Dong, Y.-F. Qu, K. Chen, J. Chen, Y.-F. Cui and L.-F. Chen, Integration of confinement crosslinking and in situ grafting for constructing artificial interphases toward stabilized zinc anodes, Energy Environ. Sci., 2024, 17, 7258–7270 RSC.
  14. X. Zhao, Y. Wang, C. Huang, Y. Gao, M. Huang, Y. Ding, X. Wang, Z. Si, D. Zhou and F. Kang, Tetraphenylporphyrin–based chelating ligand additive as a molecular sieving interfacial barrier toward durable aqueous zinc metal batteries, Angew. Chem., Int. Ed., 2023, 62, e202312193 CrossRef CAS PubMed.
  15. Y. Li, J. Cheng, D. Zhao, X. Chen, G. Sun, S. Qiao, W. Zhang and Q. Zhu, Inhibiting zinc dendrites and side reactions enabled by solvation structure regulation and facile de-solvation process, Energy Storage Mater., 2023, 63, 102997 CrossRef.
  16. Z. Cheng, K. Wang, J. Fu, F. Mo, P. Lu, J. Gao, D. Ho, B. Li and H. Hu, Texture exposure of unconventional (101)Zn facet: enabling dendrite-free Zn deposition on metallic zinc anodes, Adv. Energy Mater., 2024, 14, 2304003 CrossRef CAS.
  17. X. Jin, G. Lai, X. Xiu, L. Song, X. Li, C. Dai, M. Li, Z. Quan, B. Tang, G. Shao, Z. Zhang, F. Liu, L. Qu and Z. Zhou, Solvent polarity-induced regulation of cation solvation sheaths for high-voltage zinc-based batteries with a 1.94 V discharge platform, Angew. Chem., Int. Ed., 2024, e202418682 Search PubMed.
  18. Z. Luo, L. Ren, Y. Chen, Y. Zhao, Y. Huyan, Z. Hou and J.-G. Wang, Regulating the interface chemistry of separator to normalize zinc deposition for long lifespan Zn batteries, Chem. Eng. J., 2024, 481, 148448 CrossRef CAS.
  19. Y. Zhang, G. Yang, M. L. Lehmann, C. Wu, L. Zhao, T. Saito, Y. Liang, J. Nanda and Y. Yao, Separator effect on zinc electrodeposition behavior and its implication for zinc battery lifetime, Nano Lett., 2021, 21, 10446–10452 CrossRef CAS PubMed.
  20. X. Li, M. Li, K. Luo, Y. Hou, P. Li, Q. Yang, Z. Huang, G. Liang, Z. Chen, S. Du, Q. Huang and C. Zhi, Lattice matching and halogen regulation for synergistically induced uniform zinc electrodeposition by halogenated Ti3C2 MXenes, ACS Nano, 2022, 16, 813–822 CrossRef CAS PubMed.
  21. Q. Jian, Z. Guo, L. Zhang, M. Wu and T. Zhao, A hierarchical porous tin host for dendrite-free, highly reversible zinc anodes, Chem. Eng. J., 2021, 425, 130643 CrossRef CAS.
  22. Q. H. Cao, H. Gao, Y. Gao, J. Yang, C. Li, J. Pu, J. J. Du, J. Y. Yang, D. M. Cai, Z. H. Pan, C. Guan and W. Huang, Regulating dendrite-free zinc deposition by 3D zincopilic nitrogen-doped vertical graphene for high-performance flexible Zn-ion batteries, Adv. Funct. Mater., 2021, 31, 2103922 CrossRef CAS.
  23. J. Zhou, W. Ma, Y. Mei, F. Wu, C. Xie, K. Wang, L. Zheng, L. Li and R. Chen, Constructing stable zinc-metal anodes by synergizing hydrophobic host with zincophilic interface for aqueous zinc ions batteries, Small Methods, 2024, 2301411 CrossRef PubMed.
  24. T. Chen, Y. Wang, Y. Yang, F. Huang, M. Zhu, B. T. W. Ang and J. M. Xue, Heterometallic seed-mediated zinc deposition on inkjet printed silver nanoparticles toward foldable and heat-resistant zinc batteries, Adv. Funct. Mater., 2021, 31, 2101607 CrossRef CAS.
  25. Y. J. Zhang, G. Y. Wang, F. F. Yu, G. Xu, Z. Li, M. Zhu, Z. J. Yue, M. H. Wu, H. K. Liu, S. X. Dou and C. Wu, Highly reversible and dendrite-free Zn electrodeposition enabled by a thin metallic interfacial layer in aqueous batteries, Chem. Eng. J., 2021, 416, 128062 CrossRef CAS.
  26. Y. X. Zeng, X. Y. Zhang, R. F. Qin, X. Q. Liu, P. P. Fang, D. Z. Zheng, Y. X. Tong and X. H. Lu, Dendrite-free zinc deposition induced by multifunctional cnt frameworks for stable flexible Zn-ion batteries, Adv. Mater., 2019, 31, 1903675 CrossRef PubMed.
  27. L. Zeng, J. He, C. Yang, D. Luo, H. Yu, H. He and C. Zhang, Direct 3D printing of stress-released Zn powder anodes toward flexible dendrite-free Zn batteries, Energy Storage Mater., 2023, 54, 469–477 CrossRef.
  28. M. Zhou, G. Sun and S.-Q. Zang, Uniform zinc deposition on O,N-dual functionalized carbon cloth current collector, J. Energy Chem., 2022, 69, 76–83 CrossRef CAS.
  29. Y. Li, Z. X. Tan, Y. S. Liang, Y. Xiao, D. D. Cen, Y. L. Liu and Y. R. Liang, Amine-functionalized carbon cloth host for dendrite-free Zn metal anodes, ACS Appl. Energy Mater., 2021, 4, 4482–4488 CrossRef CAS.
  30. Y. Qian, C. Meng, J. He and X. Dong, A lightweight 3D Zn@Cu nanosheets@activated carbon cloth as long-life anode with large capacity for flexible zinc ion batteries, J. Power Sources, 2020, 480, 228871 CrossRef CAS.
  31. J. Cao, D. Zhang, X. Zhang, Z. Zeng, J. Qin and Y. Huang, Strategies on regulating Zn2+ solvation structure for dendrites-free and side reactions-suppressed zinc-ion batteries, Energy Environ. Sci., 2022, 15, 499–528 RSC.
  32. L. Liu and A. Corma, Metal catalysts for heterogeneous catalysis: from single atoms to nanoclusters and nanoparticles, Chem. Rev., 2018, 118, 4981–5079 CrossRef CAS PubMed.
  33. Y. Xu, Q.-G. Dong, J.-P. Dong, H. Zhang, B. Li, R. Wang and S.-Q. Zang, Assembly of copper-clusters into a framework: enhancing the structural stability and photocatalytic HER performance, Chem. Commun., 2023, 59, 3067–3070 RSC.
  34. L. J. Liu, Z. Y. Wang, Z. Y. Wang, R. Wang, S. Q. Zang and T. C. W. Mak, Mediating CO2 electroreduction activity and selectivity over atomically precise copper clusters, Angew. Chem., Int. Ed., 2022, 61, e202205626 CrossRef CAS PubMed.
  35. X.-K. Wan, X.-L. Cheng, Q. Tang, Y.-Z. Han, G. Hu, D.-E. Jiang and Q.-M. Wang, Atomically precise bimetallic Zu19Cu30 nanocluster with an icosidodecahedral Cu30 shell and an alkynyl–Cu interface, J. Am. Chem. Soc., 2017, 139, 9451–9454 CrossRef CAS PubMed.
  36. Y. Zhou, W. Gu, R. Wang, W. Zhu, Z. Hu, W. Fei, S. Zhuang, J. Li, H. Deng, N. Xia, J. He and Z. Wu, Controlled sequential doping of metal nanocluster, Nano Lett., 2024, 24, 2226–2233 CrossRef CAS PubMed.
  37. Z. Wu, L.-F. Wang, X.-F. Liu, R.-W. Huang, R. Wang, G. Sun and S.-Q. Zang, Regulating electrochemical performance of Cu7S4 electrodes via ligand engineering in copper cluster precursors, Nano Res., 2024, 17, 9746–9755 CrossRef CAS.
  38. Y. Wang, X.-K. Wan, L. Ren, H. Su, G. Li, S. Malola, S. Lin, Z. Tang, H. Häkkinen, B. K. Teo, Q.-M. Wang and N. Zheng, Atomically precise alkynyl-protected metal nanoclusters as a model catalyst: observation of promoting effect of surface ligands on catalysis by metal nanoparticles, J. Am. Chem. Soc., 2016, 138, 3278–3281 CrossRef CAS PubMed.
  39. X. Hu, W. Wang and Y. Huang, Copper nanocluster-based fluorescent probe for sensitive and selective detection of Hg2+ in water and food stuff, Talanta, 2016, 154, 409–415 CrossRef CAS PubMed.
  40. L.-F. Chen, Z.-H. Huang, H.-W. Liang, W.-T. Yao, Z.-Y. Yu and S.-H. Yu, Flexible all-solid-state high-power supercapacitor fabricated with nitrogen-doped carbon nanofiber electrode material derived from bacterial cellulose, Energy Environ. Sci., 2013, 6, 3331–3338 RSC.
  41. X. Wang, H. Pan, Q. Lin, H. Wu, S. Jia and Y. Shi, One-step synthesis of nitrogen-doped hydrophilic mesoporous carbons from chitosan-based triconstituent system for drug release, Nanoscale Res. Lett., 2019, 14, 259 CrossRef PubMed.
  42. L. Wang, G. Fan, J. Liu, L. Zhang, M. Yu, Z. Yan and F. Cheng, Selective nitrogen doping on carbon cloth to enhance the performance of zinc anode, Chin. Chem. Lett., 2021, 32, 1095–1100 CrossRef CAS.
  43. Y. C. Liang, Y. Y. Wang, H. W. Mi, L. N. Sun, D. T. Ma, H. W. Li, C. X. He and P. X. Zhang, Functionalized carbon nanofiber interlayer towards dendrite-free, Zn-ion batteries, Chem. Eng. J., 2021, 425, 131862 CrossRef CAS.
  44. J. Wang, H. Zhang, L. Yang, S. Zhang, X. Han and W. Hu, In situ implanting 3D carbon network reinforced zinc composite by powder metallurgy for highly reversible Zn-based battery anodes, Angew. Chem., Int. Ed., 2024, 63, e202318149 CrossRef CAS PubMed.
  45. X. Jia, J. Li and E. Wang, Cu nanoclusters with aggregation induced emission enhancement, Small, 2013, 9, 3873–3879 CrossRef CAS PubMed.
  46. Q. Meng, X. Jin, N. Chen, A. Zhou, H. Wang, N. Zhang, Z. Song, Y. Huang, L. Li, F. Wu and R. Chen, Interface engineering with dynamics-mechanics coupling for highly reactive and reversible aqueous zinc-ion batteries, Adv. Sci., 2023, 10, 2306656 CrossRef CAS PubMed.
  47. J. Ding, Y. Liu, S. Huang, X. Wang, J. Yang, L. Wang, M. Xue, X. Zhang and J. Chen, In situ construction of a multifunctional quasi-gel layer for long-life aqueous zinc metal anodes, ACS Appl. Mater. Interfaces, 2021, 13, 29746–29754 CrossRef CAS PubMed.
  48. K. Zhao, G. Fan, J. Liu, F. Liu, J. Li, X. Zhou, Y. Ni, M. Yu, Y. M. Zhang, H. Su, Q. Liu and F. Cheng, Boosting the kinetics and stability of Zn anodes in aqueous electrolytes with supramolecular cyclodextrin additives, J. Am. Chem. Soc., 2022, 144, 11129–11137 CrossRef CAS PubMed.
  49. X. Jiang, Y. Lam, W. Li, S. Jiang and H. Jia, Self-healing composite anode induced by liquid-metal interlayer for flexible Zn ion storage application, Compos. Commun., 2024, 45, 101796 CrossRef.
  50. P. X. Sun, Z. Cao, Y. X. Zeng, W. W. Xie, N. W. Li, D. Luan, S. Yang, L. Yu and X. W. D. Lou, Formation of super-assembled TiOx/Zn/N-doped carbon inverse opal towards dendrite-free Zn anodes, Angew. Chem., Int. Ed., 2022, 61, e202115649 CrossRef CAS PubMed.
  51. J.-H. Wang, L.-F. Chen, W.-X. Dong, K. Zhang, Y.-F. Qu, J.-W. Qian and S.-H. Yu, Three-dimensional zinc-seeded carbon nanofiber architectures as lightweight and flexible hosts for a highly reversible zinc metal anode, ACS Nano, 2023, 17, 19087–19097 CrossRef CAS PubMed.
  52. S. Kumar, H. Yoon, H. Park, G. Park, S. Suh and H.-J. Kim, A dendrite-free anode for stable aqueous rechargeable zinc-ion batteries, J. Ind. Eng. Chem., 2022, 108, 321–327 CrossRef CAS.
  53. H. Yan, S. Li, J. Zhong and B. Li, An electrochemical perspective of aqueous zinc metal anode, Nano-Micro Lett., 2024, 16, 15 CrossRef CAS PubMed.
  54. Y. Song, Y. Liu, S. Luo, Y. Yang, F. Chen, M. Wang, L. Guo, S. Chen and Z. Wei, Blocking the dendrite-growth of Zn anode by constructing Ti4O7 interfacial layer in aqueous zinc-ion batteries, Adv. Funct. Mater., 2024, 34, 2316070 CrossRef CAS.
  55. X. Zhang, L. Zhang, X. Jia, W. Song and Y. Liu, Design strategies for aqueous zinc metal batteries with high zinc utilization: from metal anodes to anode-free structures, Nano-Micro Lett., 2024, 16, 75 CrossRef CAS PubMed.
  56. Z. Jiang, S. Zhai, L. Shui, Y. Shi, X. Chen, G. Wang and F. Chen, Dendrite-free Zn anode supported with 3D carbon nanofiber skeleton towards stable zinc ion batteries, J. Colloid Interface Sci., 2022, 623, 1181–1189 CrossRef CAS.
  57. Q. Zhang, J. Y. Luan, L. Fu, S. G. Wu, Y. G. Tang, X. B. Ji and H. Y. Wang, The three-dimensional dendrite-free zinc anode on a copper mesh with a zinc-oriented polyacrylamide electrolyte additive, Angew. Chem., Int. Ed., 2019, 58, 15841–15847 CrossRef CAS PubMed.
  58. Z. Kang, C. L. Wu, L. B. Dong, W. B. Liu, J. Mou, J. W. Zhang, Z. W. Chang, B. Z. Jiang, G. X. Wang, F. Y. Kang and C. J. Xu, 3D porous copper skeleton supported zinc anode toward high capacity and long cycle life zinc ion batteries, ACS Sustainable Chem. Eng., 2019, 7, 3364–3371 CrossRef CAS.
  59. Q. Zhang, J. Y. Luan, X. B. Huang, L. Zhu, Y. G. Tang, X. B. Ji and H. Y. Wang, Simultaneously regulating the ion distribution and electric field to achieve dendrite-free Zn anode, Small, 2020, 16, 2000929 CrossRef CAS PubMed.
  60. M. Zhou, Z. Wu, R. Wang, G. Sun and S.-Q. Zang, An in situ reduction strategy toward dendrite-free Zn anodes, Sci. China Mater., 2023, 66, 1757–1766 CrossRef CAS.
  61. R.-H. Wang, W. Wang, Y.-Z. Zhang, W. Hu, L. Yue, J.-H. Ni, W.-Q. Zhang, G. Pei, S. Yang and L.-F. Chen, Photoexcitation–enhanced high-ionic conductivity in polymer electrolytes for flexible, all-solid-state lithium-metal batteries operating at room temperature, Angew. Chem., Int. Ed., 2024, e202417605 Search PubMed.
  62. F. Yang, J. A. Yuwono, J. Hao, J. Long, L. Yuan, Y. Wang, S. Liu, Y. Fan, S. Zhao, K. Davey and Z. Guo, Understanding H2 evolution electrochemistry to minimize solvated water impact on zinc-anode performance, Adv. Mater., 2022, 34, 2206754 CrossRef CAS PubMed.
  63. L. Hong, X. Wu, Y. S. Liu, C. Yu, Y. Liu, K. Sun, C. Shen, W. Huang, Y. Zhou, J. S. Chen and K. X. Wang, Self–adapting and self-healing hydrogel interface with fast Zn2+ transport kinetics for highly reversible zn anodes, Adv. Funct. Mater., 2023, 33, 2300952 CrossRef CAS.
  64. Z. Zhao, J. Zhao, Z. Hu, J. Li, J. Li, Y. Zhang, C. Wang and G. Cui, Long-life and deeply rechargeable aqueous Zn anodes enabled by a multifunctional brightener-inspired interphase, Energy Environ. Sci., 2019, 12, 1938–1949 RSC.
  65. D. Xu, Z. Wang, C. Liu, H. Li, F. Ouyang, B. Chen, W. Li, X. Ren, L. Bai, Z. Chang, A. Pan and H. Zhou, Water catchers within sub-nano channels promote step-by-step zinc-ion dehydration enable highly efficient aqueous zinc-metal batteries, Adv. Mater., 2024, 36, 2403765 CrossRef CAS PubMed.
  66. H. Dong, J. Li, J. Guo, F. Lai, F. Zhao, Y. Jiao, D. J. L. Brett, T. Liu, G. He and I. P. Parkin, Insights on flexible zinc-ion batteries from lab research to commercialization, Adv. Mater., 2021, 33, 2007548 CrossRef CAS PubMed.
  67. H. Pan, Y. Shao, P. Yan, Y. Cheng, K. S. Han, Z. Nie, C. Wang, J. Yang, X. Li, P. Bhattacharya, K. T. Mueller and J. Liu, Reversible aqueous zinc/manganese oxide energy storage from conversion reactions, Nat. Energy, 2016, 1, 16039 CrossRef CAS.
  68. J. Cui, Z. Li, A. Xu, J. Li and M. Shao, Confinement of zinc salt in ultrathin heterogeneous film to stabilize zinc metal anode, Small, 2021, 17, 2100722 CrossRef CAS PubMed.
  69. Z. Cao, Y. B. Zhu, K. Chen, Q. Wang, Y. Li, X. Xing, J. Ru, L. G. Meng, J. Shu, N. Shpigel and L. F. Chen, Super–stretchable and high-energy micro-pseudocapacitors based on mxene embedded Ag nanoparticles, Adv. Mater., 2024, 36, 2401271 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4qi02605e

This journal is © the Partner Organisations 2025
Click here to see how this site uses Cookies. View our privacy policy here.