Structural characterization of A-ZIF-8 layer
A facile spray method was used to prepare the A-ZIF-8 (A-ZIF-8) protective layer (Supplementary Fig. 1). ZIF-8 crystals are known to be structurally similar to zeolites, in which 2-methylimidazole (Hmim) donates a proton and couples with the Zn2+ node to form a ZnN4 tetrahedron30 (Fig. 1a). However, in the presence of CH3COO−, the copolymerization of Zn2+ with Hmim ligand was partially prevented by the steric hindrance due to the relatively large anion size of CH3COO−.31 As a result, the nucleation of orderly linked ZnN4 tetrahedra was disrupted, resulting in A-ZIF-8 structure (Fig. 1b). As seen in Fig. 1c, a continuous and compact A-ZIF-8 layer with no grain boundary was successfully coated on zinc foil. After increasing the heating time to 8 h, ZIF-8 grains with clear boundaries appeared on the zinc foil surface according to the field emission scanning electron microscopy (FESEM, Fig. 1d, Supplementary Fig. 2) and atomic force microscopy (AFM, Supplementary Fig. 3), suggesting that a longer heating time facilitates the generation of crystalline structure of ZIF-8. The thickness of the A-ZIF-8 layer was observed to be approximately 2.2 µm (Fig. 1e). Based on elemental mappings (Fig. 1f), all the elements were evenly distributed in the layer. The amounts of the C and N elements on Zn@A-ZIF-8 origin from the presence of the A-ZIF-8. The diffraction peaks of the C-ZIF-8 were consistent with those of the simulated ZIF-8 (Fig. 1g), indicating that the C-ZIF-8 had a crystalline structure. However, the loss of long-distance order as indicated by XRD patterns makes it difficult to characterize the amorphous structures.32 As shown in Fig. 1g and Supplementary Fig. 4, the A-ZIF-8 exhibits an absence of bragg peaks due to amorphous structure. The flat surface of the Zn@A-ZIF-8 anode is also confirmed by atomic force microscopy (AFM) (Fig. 1h, Supplementary Fig. 5).
Figure 1i showed that contact angles are 93.5°, 47.5°, and 45.5° for the zinc foil, Zn@C-ZIF-8, and Zn@A-ZIF-8, respectively, indicating that the hydrophilicity of the anode can be enhanced by the C/A-ZIF-8 modification. It was reported that enhancing the hydrophilicity of the metal anode facilitates contact between electrolyte ions and the metal anode, contributing to the Zn2+ flux regulation and uniform zinc deposition.33,34 As seen in Fig. 1j and Supplementary Fig. 6, the Zn@A-ZIF-8 has the largest ion transference number (i.e., 0.82) because of the absence of grain boundary. Brunauer-Emmett-Teller (BET) measurement indicates the surface area of A-ZIF-8 is as high as1265 m2 g− 1, and the main pore size distribution is 2–4 nm in diameter (Supplementary Fig. 7). Raman spectroscopy measurements were used to characterize the bond between Zn ions and the Hmim ligand. As shown in Supplementary Fig. 8, both Zn@C-ZIF-8 and Zn@A-ZIF-8 showed distinct peaks at 168 and 205 cm− 1, corresponding to the Zn = N bond. Besides, Fourier transform infrared spectroscopy (FTIR) spectra (Supplementary Fig. 9) revealed that the N-H…H stretch between 2400–3100 cm− 1 disappears completely for both A-ZIF-8 layer and C-ZIF-8 layer (compared to Hmim powders), suggesting that Hmim was deprotonated and incorporated into the ZIF-8 framework.31 The XPS spectra of samples (Supplementary Fig. 10) also verify the existence of C, N in both of the Zn@ZIF-8 and Zn@A-ZIF-8 sample. Moreover, the Zn 2p XPS spectra of the Zn foil, Zn@C-ZIF-8 and Zn@A-ZIF-8 samples (Fig. 1k) can be split into two peaks, corresponding to Zn 2p1/2 and Zn 2p3/2 of Zn0. Compared with Zn foil, both peaks for Zn@A-ZIF-8 sample shifted to higher binding energies, which might be ascribed to the fact that the metal Zn center is Zn2+ in the A-ZIF-8. High-resolution C 1s and N 1s spectra of the fabricated anodes are provided in the Supplementary Fig. 11, which indicate the chemical state of A-ZIF-8. The amorphous structure was further confirmed by extended X-ray adsorption fine structure (EXAFS) analysis (Fig. 1l). A-ZIF-8 exhibited a strong peak at ca. 1.6 Å for Zn-N coordination, which was similar to ZIF-8 powder (via chemical coprecipitation process in methanol, Supplementary Fig. 12). However, the number of N coordinated with Zn was decreased in the A-ZIF-8 structure compared with that in ZIF-8 powder (Supplementary Fig. 13, Supplementary Table 1), which lead to coordination defects between metal ions and organic ligands caused by the relatively large anion size of CH3COO− 31, thus forming a disordered amorphous structure.
Mechanical Properties Of A-ZIF-8 Layer
Due to the unique structure of amorphous MOFs, they could have better mechanical properties than crystalline MOFs. As shown in Supplementary Fig. 14, even though the prepared Zn@A-ZIF-8 was folded into various shapes, the integrity of the protect layer was still retained. No exfoliation or cracks could be observed on the A-ZIF-8 layer, which indicates a strong bonding between zinc substrate and the A-ZIF-8 protective layer. This strong bonding can be attributed to the formation of HMIM hydrogen bond (N-H…N) chains on Zn foil, resulting in a strong interfacial interaction. The binding energies of pristine and A-ZIF-8 on Zn (002) surface were calculated to evaluate the interaction strength between ZIF-8 and Zn using the forcite module with universal force field35 of Materials Studio 2017. First, bulk Zn (Supplementary Fig. 15a) and C-ZIF-8 units cells (Fig. 2a) were fully optimized. A-ZIF-8 (Fig. 2b) was obtained by Heating and annealing process based on pristine ZIF-8 supercell. All the partial atomic charges were defined by the QEQ methods. The Zn (002) surface structure (Supplementary Fig. S15b) was obtained by cleaving the optimized Zn, which contained five Zn layers. Then, one 12×7 supercell of Zn (002) surface (Supplementary Fig. S15c) was constructed serving as ZIF-8 adsorption substrates. One 2×2×2 supercell of ZIF-8 was placed on the Zn (002) surface to build ZIF-8/Zn (002) system (Fig. 2c). The lattice mismatch between ZIF-8 and Zn (002) surface was less than 2%. Placing the A-ZIF-8 on the Zn (002) surface, we built the A-ZIF-8/Zn (002) system (Fig. 2d). There is one vacuum layer of larger than 20 Å perpendicular to the surface.
After geometry optimization, the molecular dynamic simulations under NVT (T = 298 K) ensemble were carried out to fully relax the two binding systems, during which the bottom layer of Zn substructure was fixed. The Ewald scheme and atom-based cutoff method were applied to treat electrostatic and van der Waals (vdW) interactions with a cut-off value of 12.5 Å, respectively. Equations of motion were integrated with a time step of 1 fs. After obtaining stable adsorption configuration, the binding energy ΔE was calculated according to the following Eq. 36:
$$\Delta E={E_{total}} - {E_{Zn}} - {E_{ZIF - 8}}$$
1
Where Etotal is the total energy of binding system, EZn and EZIF−8 are the energies of the isolated Zn substrate and ZIF-8, respectively. Specifically, a weak interaction between crystalline ZIF-8 and Zn (002) with binding energy − 67.86 kcal/mol (Fig. 2e) was obtained, while the A-ZIF-8 shows a higher binding energy of-266.96 kcal/mol with abundant HMIM hydrogen bond (N-H…N) chains, suggesting the strong interaction between A-ZIF-8 layer and Zn substrate. Besides, nano-scratch tests were performed to quantitatively study the bonding force between zinc substrate and protective layer. In each scratch test, there are three stages: (I) a very small load was applied to the surface to trace and map the original topography of the sample surface; (II) Following the same path, but the specified normal load is applied; (III) a very small load is again applied to track and measure the residual surface deformation along the scratch path after tip unloading (elastic recovery).37. In the following results, these three stages are represented by blue, green, and orange curves, respectively, the area between green and orange curves corresponds to the degree of elastic recovery (the possibility of coating peeling).37 The results (Figs. 2f and 2g) showed that the A-ZIF-8 layer has a larger elastic recovery area than the C-ZIF-8 layer, which means the former has a stronger binding between zinc substrate and protective layer. The corresponding 3D optical image of a 100 mN ramp load nano-scratch test on Zn@A-ZIF-8 and Zn@C-ZIF-8 is shown in Supplementary Fig. 16.
Anti-corrosion And Dendrite Suppression Properties Of A-ZIF-8 Layer
To explore the anti-corrosion behavior of the A-ZIF-8 layer, Zn foil and Zn@A-ZIF-8 were soaked in a 2 M ZnSO4 solution for 3 days. As shown in Fig. 3a, the bright Zn foil turned dark grey, suggesting that corrosion reaction took place. In contrast, the Zn@A-ZIF-8 changed slightly, indicating that the A-ZIF-8 as a protective layer took effect. By examining their micromorphology, it was observed that Zn@A-ZIF-8 retains a flat surface, while Zn foil has a large number of regular hexagonal by-products deposited on surface (Fig. 3a). In order to identify the by-products, XRD analysis was conducted, showing some new peaks appearing at 8–25° that are consistent with the characteristic peaks of (Zn(OH)2)3(ZnSO4)(H2O)3 (PDF#78–0247) (Fig. 3b). With the protection of amorphous MOFs, no obvious peaks can be identified in basic zinc sulfate. The anti-corrosion properties were further examined through the linear polarization curves (Fig. 3c). Compared with the Zn foil (-1.02 V), the lower corrosion potential (Vcorr) of the Zn@A-ZIF-8 (-1.003 V) and Zn@C-ZIF-8 (-1.007 V) can be achieved.
The corrosion current density (Icorr) of the Zn@A-ZIF-8 and Zn@C-ZIF-8 anodes are 3.8 mA cm− 2 and 4.1 mA cm− 2 respectively, which are much lower than the Zn foil (7.5 mA cm− 2). The higher corrosion positive potential and lower corrosion current of the Zn@A-ZIF-8 anode indicate the less tendency of corrosion reaction (dissolved O2-induced passivation).33,38 The deposition and growth of Zn are studied by chronoamperometry (CA) at an overpotential of -150 mV using three-anode system (Fig. 3d). The current variation at a constant overpotential can reflect nucleation, process, and surface changes of anode surface. 19,39,40 At the overpotential of -150 mV, the current density keeps increasing over 200 s for the Zn foil, suggesting a long and rampant 2D diffusion process and rough deposition propagation. The absorbed ions laterally diffuse along the surface to find the most energetically favorable sites for charge transfer.13 Additionally, compared with zinc foil, the diffusion mode for the Zn@A-ZIF-8 anode evolves to initial 2D diffusion in the first 30 s, followed by a continuous and stable 3D diffusion, suggesting the restricted zinc ions tend to nucleate in the vicinity of initial nucleation sites rather than thermodynamically favorable adsorption sites with a low energy barrier on the anode.41 In addition, hydrogen evolution reaction (HER) was studied by the linear sweep voltammetry (LSV) test on different anode. As shown in Fig. 3e, the lowest onset potential (i.e., -1.79 V) was noted for the Zn@A-ZIF-8, indicating the best suppression effect on HER. As shown in Fig. 3f, the symmetric cells based on the Zn@A-ZIF-8 anode exhibit a lower nucleation overpotential (NOP) (32 mV) at a current density of 1 mA cm− 2 during the initial process, which is much smaller than Zn foil anode (65 mV). The lower NOP means a lower nucleation barrier, beneficial for a relatively uniform metal plating process.4,42
Moreover, the microscopic morphologies of the Zn foil anode and Zn@A-ZIF-8 anode after 50 cycles at 1 mA cm− 2 are probed by SEM. As shown in the surface and cross-sectional SEM images (Fig. 3g) showed that the Zn@A-ZIF-8 anode maintains a smooth surface, while irregularly shaped and disordered Zn dendrites are observed on the surface of Zn foil. Furthermore, the in-situ optical visualization of 40 minutes Zn deposition at a current density of 5 mA cm− 2 is recorded in Fig. 3h. After an initial 5 min of deposition, nonuniform protuberances appeared on the Zn foil surfaces, and grew into needle-like dendrites afterwards. After 40 min electroplating, the dendrites on the Zn foil evolved into a mossy and rough surface, which would enhance the local electric field on dendrite area, therefore, accelerating dendrite growth and increasing the risk of short circuit. During the electroplating process a large number of bubbles were generated on the Zn foil surface due to hydrogen evolution. By contrast, a uniform Zn deposition layer formed and grew on the Zn@A-ZIF-8 anode through the whole electroplating process. Piecing together all the learned information, the functional mechanism of the A-ZIF-8 as protective layer was proposed, as shown in Fig. 4, the hydrogen evolution, corrosion, and dendrites of zinc anode can be perfectly solved by our ultra strong binding and grain boundary free A-ZIF-8 coating.
Electrochemical performances of Zn@A-ZIF-8 anode.
The electrochemical behavior of Zn foil anode and Zn@A-ZIF-8 anode based symmetric cells were evaluated at different current density (Figs. 5a, 5c, Supplementary Fig. 17). At a current density of 1 mA cm− 2 (Fig. 5a), the Zn foil symmetric cell showed abrupt failure after 120 h. Impressively, benefiting from the A-ZIF-8 layer, the Zn@A-ZIF-8 symmetric cell achieved an ultra-long cycle life, about ten months (7000 h), which is 12 times higher than the cycle life of crystalline MOF protective layer (Fig. 5b). At a high current density of 10 mA cm− 2, the Zn@A-ZIF-8 symmetric cells still exhibit high stability with a prolonged cycling life of over 5500 cycles (1200 h, Fig. 5c). While the Zn foil symmetric cell suffers severe polarization after 105 h at 10 mA cm− 2. The performance of Zn@A-ZIF-8 anode is compared with that of other surface-modified Zn anodes reported in recent studies (Fig. 5d). The rate performance of symmetric cells using the Zn foil anode and Zn@A-ZIF-8 anode are shown in Fig. 5e. Compared with the Zn foil anode, the Zn@A-ZIF-8 anode displays a stable voltage profile with a lower voltage hysteresis. When the current density increases from 1 to 10 mA cm− 2, the cells using the Zn@A-ZIF-8 anode exhibit a steadily increasing hysteresis of 36, 43, 45, 46, and 52 mV, respectively (Fig. 5f).
Besides, when the current densities increased to 20 mA cm− 2 (51% Zn utilization, Fig. 5g), the Zn@A-ZIF-8 symmetric cells can keep cycling for over 150 h, indicating that A-ZIF-8 layer has excellent tolerance to volume change of the Zn anodes. Electrochemical impedance spectroscopy (EIS) tests were performed before and after 50 cycles (Fig. 5h). The charge transfer resistance (RCT) of the Zn@A-ZIF-8 anode was slightly lower than that of the Zn foil anode before cycling. After 50 cycles, the RCT was significantly reduced due to the infiltration of electrolytes during the cycling process, indicating the fast charge transfer and Zn plating/stripping kinetics.
Coulombic efficiency (CE) is an important parameter to evaluate the reversibility of a cell anode during electrochemical process. In this work, the prepared Zn//Ti and Zn//Ti@A-ZIF-8 cells were studied during the galvanizing/stripping process. As shown in Fig. 6a, at a current density of 3 mA cm− 2, the average CE of Zn//Ti@A-ZIF-8 cells reached 99.5% after running over 1000 cycles. By contrast, the Zn//Ti cell obtained an average CE of 79.2% and showed dramatic fluctuations after running 80 cycles, suggesting that the galvanizing/stripping process on Ti foil encounters instability issue. The initial voltage gap of Zn//Ti@A-ZIF-8 cells is 51mV, much smaller than that of Zn//Ti cells (i.e., 115mV) as exhibited in Fig. 6b and Supplementary Fig. 18. It indicates that the A-ZIF-8 layer can effectively reduce the energy barrier for Zn nucleation/dissolution and suppress dendrite growth on Ti foil. The Zn@A-ZIF-8 anode was further investigated in full cell, where homemade α-MnO2 was used as cathode. Before the test, the structure and morphology of the cathode material were observed by X-ray diffractometer and scanning electron microscope as shown in Supplementary Fig. 20. Zn//MnO2 full cells were tested in an electrolyte consisting of 2 M ZnSO4 and 0.2 M MnSO4. The cyclic voltammetry (CV) curves of ZIBs based on the Zn foil anode and the Zn@A-ZIF-8 anode are exhibited in Fig. 6c. The typical redox peaks with high reversibility can be observed for different anodes, suggesting the presence of A-ZIF-8 layer does not affect the electrochemical kinetics of Zn plating/stripping. The AC impedance spectrum of AZIBs is shown in Fig. 6e. The RCT value of the Zn@A-ZIF-8//MnO2 cell (167.8 Ω) is lower than that of the Zn foil //MnO2 cell (557.7 Ω), suggesting a fast charge transfer of Zn@A-ZIF-8 //MnO2 cell, which is in good agreement with the results from a symmetric cell. In addition, the long-term cycling stability of full cell was tested at a current density of 1 A g− 1. The Zn@A-ZIF-8//MnO2 cell exhibits much better cycling stability with a capacity retention rate of 98.9% over 2000 cycles than the MnO2//Zn cell (30.8%) as shown in Fig. 6f, which agrees well with the galvanostatic charge/discharge curves (Fig. 6d and Supplementary Fig. 21). The elevated electrochemical properties of the Zn@A-ZIF-8//MnO2 cell could be attributed to the high reversible properties of the Zn@A-ZIF-8 anode, which possesses a stable cyclic life without Zn dendrites. Furthermore, the morphology changes of the Zn foil anode and Zn@A-ZIF-8 anode after 1000 cycles in the full cell were investigated by SEM (Supplementary Fig. 22). The Zn@A-ZIF-8 anode shows a flat and dense surface after 1000 cycles, while a rough surface with lots of protrusions on Zn foil was observed.
Flexible symmetrical batteries based on Zn foil and Zn@A-ZIF-8 anode were assembled, and their cycle performances were tested at different bending angle (Fig. 7a). The lifespan of the Zn foil symmetric cell varies with various bending angles (Supplementary Fig. 23). A 400 h survival time was achieved for unfolded; a 150 h survival time was stabilized for folded 90 degrees, and only a 50 h survival time was retained for folded 180 degrees. Locally high electric field intensity promotes the migration of zinc ions towards the folded area, resulting in an uneven zinc deposition.43 The SEM images (Supplementary Fig. 24b) are also consistent with the optical image (Supplementary Fig. 23b). Because more Zn is deposited onto the folded aera. For the Zn symmetric cells based on Zn@A-ZIF-8 anode, the Zn symmetric cells demonstrate extraordinary cyclability for more than 800 h regardless of the bending angles (Fig. 7a).
The highly adhesive A-ZIF-8 layer can eliminate local high electric field during folding and prohibit Zn dendrite piecing in the curve area (Fig. 7b). The SEM images of 800 h cycled Zn@A-ZIF-8 anode shows uniform and flat zinc deposition even in the folded aera (Supplementary Fig. 24d). Besides, the A-ZIF-8 layer could be deposited on other metal substrates such as Al, Cu, and Mg by the same procedure (Fig. 7c). Cross-sectional SEM images suggest good adhesion of the A-ZIF-8 layer to these substrates (Fig. 7d), indicating a broad potential application as protection layers for metal anodes. The Mg@A-ZIF-8//Mg@A-ZIF-8 and Al@A-ZIF-8//Al@A-ZIF-8 symmetric batteries were assembled to explore the application of A-ZIF-8 layers in other battery systems. Compared with that of bare Mg anode (2.1V, 25 h), the Mg@A-ZIF-8 anode achieved a lower stripping overpotential of 1.05 V and a higher cycle life (250 h, Fig. 7e). This is because the A-ZIF-8 layer eliminates nonselective inter crystalline diffusion of electrolyte and realize selective Mg2+ transport. Different from the rough and loose surface layer on the Mg foil anode (Supplementary Fig. 25), the Mg@A-ZIF-8 anode displayed a flat surface after 50 cycles (Fig. 7f, Supplementary Fig. 26), suggesting that the MOF membrane can stabilize the Mg metal anode during the plating/stripping cycles, possibly attributed to the MOF channels. Similarly, as shown in Supplementary Fig. 27, the Al@A-ZIF-8//Al@A-ZIF-8 symmetrical battery exhibits 300 h cycle life, which is higher than the Al foil//Al foil symmetrical battery (120 h).