Kun
Zhang‡
a,
Yijia
Yuan‡
a,
Gang
Wang‡
b,
Fangzheng
Chen
a,
Li
Ma
c,
Chao
Wu
d,
Jia
Liu
a,
Bao
Zhang
e,
Chenglin
Li
a,
Hongtian
Liu
a,
Changan
Lu
a,
Xing
Li
f,
Shibo
Xi
d,
Keyu
Xie
*c,
Junhao
Lin
*bg and
Kian Ping
Loh
*a
aDepartment of Chemistry, National University of Singapore, Singapore, Singapore. E-mail: chmlohkp@nus.edu.sg
bDepartment of Physics and Shenzhen Key Laboratory of Advanced Quantum Functional Materials and Devices, Southern University of Science and Technology, Shenzhen, China. E-mail: linjh@sustech.edu.cn
cState Key Laboratory of Solidification Processing, Center for Nano Energy Materials, Northwestern Polytechnical University and Shaanxi Joint Laboratory of Graphene (NPU), Xi’an, China. E-mail: kyxie@nwpu.edu.cn
dInstitute of Chemical and Engineering Sciences, Agency of Science Technology and Research, 1 Pesek Road, Jurong Island, Singapore, Singapore
eSchool of Materials and Energy, University of Electronic Science and Technology of China, Chengdu, Sichuan, China
fDepartment of Chemistry, State Key Laboratory of Marine Pollution, City University of Hong Kong, Hong Kong, China
gQuantum Science Center of Guangdong-Hong Kong-Macao Greater Bay Area (Guangdong), Shenzhen, 518045, People's Republic of China
First published on 20th March 2025
Rechargeable aqueous zinc metal-based batteries present a promising alternative to conventional lithium-ion batteries due to their lower operating potentials, higher capacities, intrinsic safety, cost-effectiveness, and environmental sustainability. However, the use of aqueous electrolyte in zinc metal-based batteries presents its own unique set of challenges, which include the tendency for side reactions during discharge that encourages dendritic growth on Zn anodes, as well as sluggish kinetics caused by the large solvation shell of divalent Zn ions. Nanoporous materials can be deployed as coating on Zn anodes for enhancing both their performance and stability, particularly in addressing challenges associated with water reactivity and ion migration kinetics. In our study, we incorporated superhydrophobic fluorine chains into covalent organic frameworks (SPCOFs) to engineer nanochannels that facilitate efficient ion migration pathways. Molecular dynamics simulations demonstrate that these superhydrophobic fluorine chains significantly reduce interactions between the electrolyte and nanochannel walls, altering the confined electrolyte distribution. This modification enables rapid dehydration, reduces ion migration resistance, and promotes dense Zn deposition. The use of SPCOFs enable Zn batteries with exceptional stability, achieving over 5000 hours of runtime at high current densities and stable cycling across 800 cycles in full-cell configurations. This approach highlights the critical role of tailored nanochannel environments in advancing the functionality and durability of zinc metal-based batteries, offering a scalable and environmentally friendly alternative to traditional battery technologies.
Broader contextAs the global push for renewable energy gains momentum, there is a critical need for advancements in energy storage technologies that are both efficient and environmentally sustainable. Zinc-based batteries are stepping up as a formidable alternative to traditional lithium-ion batteries, lauded for their safety and cost-effectiveness. Yet, their broader application has been curtailed by persistent challenges such as dendritic growth and inefficient ion transport. Our research introduces a pioneering solution involving the use of superhydrophobic perfluoro chain-decorated covalent organic frameworks (SPCOFs). These frameworks are meticulously engineered to optimize ion migration pathways within zinc batteries, promoting rapid ion transport and consistent zinc deposition. This approach dramatically diminishes the resistance to ion migration, significantly boosting the battery's stability and longevity. Moreover, our study delves into the intricacies of ion interaction and distribution within sub-nanochannels, marking a notable first in this area of research. These enhancements could revolutionize zinc battery technology, positioning it as a cornerstone for future, more sustainable, and widely accessible energy storage systems. |
Generally, the migration of Zn ions in the electrolyte across the nanochannel involves a dehydration and migration process. The predominant ion migration mechanism in sub-nanochannels is characterized by a ‘sites hopping’ pattern,8,13 as shown in Fig. 1b. In this process, partially dehydrated ions and water molecules adsorb on specific zincophilic sites and traverse the one-dimensional channels. However, such a migration mode faces considerable flow frictions due to the strong interactions between the electrolyte and the channel, resulting in increased energy barriers and resistance experienced by the electrolyte ions.16–18 Moreover, the functional groups used in sub-nanochannels, such as low-density doping atoms or small groups like –CO, –C–F, –COOH, and –SO3, were low-density and small-functional-area, decrease efficiency in improving ion transport or suppressing of dendritic growth.8,12,13 Therefore, there is a need to explore effective channel functionalization approaches. Inspired by the ultrafast solution permeation observed in aquaporins,19 which is facilitated by their hydrophobic interior surface, recent studies have focused on enhancing rapid water permeation by modifying water–wall interactions within nanochannels. These modifications include creating superhydrophobic regions within nanotubes or constructing fluorous oligoamide nanorings.20,21 By reducing interfacial frictions between water and the channel through a dewetting effect, the transport of solution within the nanochannels can be accelerated.21 Although this approach holds promise for addressing the slow ion migration kinetics in AZBs, it remains unexplored.
In this study, we modified the nanochannels within COFs by incorporating superhydrophobic fluorine chains to construct SPCOFs. The introduction of fluorine effectively reduced the surface energy of the nanochannels, tailoring the interactions between the electrolyte solution and the channel walls to facilitate rapid dehydration. Through molecular dynamics (MD) simulations, we further observed a unique ion migration behavior in SPCOFs: the dehydrated Zn solvent sheath detached from the channel walls and occupied larger central areas within the unwetted nanochannels. This structural adjustment is crucial for reducing resistance to ion migration. SPCOF-modified Zn anode (SPCOF@Zn) exhibited exceptional stability and kinetics, achieving an impressive runtime of 5000 hours at a high current density of 10 mA cm−2. Furthermore, when the SPCOF@Zn anode was incorporated into a Zn||ZVO full cell, stable cycling was maintained for over 500 cycles and up to 800 cycles at low N/P ratios of approximately 2 and 4, respectively. These results highlight the significant potential of this innovative approach for practical battery applications.
The Debye length of zinc ions in a 2 M ZnSO4 electrolyte can be calculated as follows:
![]() | (1) |
Fig. 1 illustrates three zinc anode scenarios: (a) a bare Zn anode, (b) a Zn anode coated with nanochanneled porous layers under 5 nm in diameter (∼2 times Debye length), and (c) a similar setup with low-surface-energy channels. In Fig. 1a, migration of Zn ions in the bare Zn anode is influenced by electromigration, this usually contributes to a multitude of issues such as uncontrolled deposition, electrode corrosion and HER, which promotes zinc dendrite formation. Fig. 1b demonstrates the dehydration of Zn ions using sub-5 nm nanochannels in the porous layers, where 5 nm is close to the Debye screening length in 2 M ZnSO4 solution. The migration of zinc ions occurs via a 'site hopping' mechanism, in which the interaction between electrolyte and partially wetted nanochannel imparts resistance to the flow.16–18Fig. 1c illustrates the case where a low-surface-energy wall (via inner fluorinated organic chains) reducing the electrolyte-channel interaction, thus promoting fast ion migration and dense zinc deposition.
Both SPCOF and SHCOF displayed excellent crystallinity and an antiparallel-stacked architecture, as confirmed by powder X-ray diffraction (PXRD) analyses (Fig. 2e and Fig. S5 and S6, ESI†). The PXRD patterns for SPCOF revealed peaks at 3.66°, 6.35°, and 7.20°, correlating to the (100), (110), and (200) planes, respectively. Similarly, SHCOF exhibited peaks at 3.52°, 6.12°, and 7.08°, corresponding to these planes. Theoretical simulations of eclipsed, staggered, and antiparallel stacking configurations were conducted, with the antiparallel structures showing greater congruence with the experimental PXRD result. SPCOF is characterized by antiparallel stacking within a P3/C1 space group and refined cell parameters of a = 29.40 Å, b = 29.40 Å, c = 9.37 Å, α = γ = 90°, and β = 120°. In comparison, SHCOF, also exhibiting antiparallel stacking, possesses a slightly tilted unit cell within the same space group, with parameters of a = 30.14 Å, b = 30.14 Å, c = 7.37 Å, α = γ = 90°, and β = 120°.
The structural integrity and successful synthesis of SPCOFs were rigorously validated through Fourier-transform infrared spectroscopy (FT-IR) and solid-state nuclear magnetic resonance (NMR) spectroscopy (Fig. 2f and Fig. S7 and S8, ESI†). The FT-IR spectra featured a distinct peak at approximately 1155 cm−1, characteristic of C–F bond vibrations,25 affirming the successful integration of perfluorohexyloxy chains into the COF structure. Further confirmation was provided by solid-state cross-polarization/magic-angle-spinning (CP/MAS) 13C NMR spectroscopy, where peaks around 140 ppm indicated the carbon atoms involved in forming hydrazone linkages. Additional peaks ranging from 108.4 to 145.9 ppm were ascribed to the carbons in the perfluoro-chains and aromatic carbons within the COF framework. Notably, compared to SHCOF, SPCOF exhibited new peaks at higher fields, corresponding to the carbons on the perfluoro-chains.
The high crystallinity of SPCOF and SHCOF were validated using low-dose cryogenic transmission electron microscopy (cryo-TEM). Specifically, SPCOF and SHCOF were exfoliated using ethanol-assisted liquid sonication, and the thin flakes obtained were dispersed onto lacey carbon films. These samples were subsequently frozen rapidly at liquid nitrogen temperatures (∼77 K) for TEM analysis. The low-magnification cryo-TEM images shown in Fig. S9a and S10a (ESI†) display the characteristic morphology of large SPCOF and SHCOF nanosheets obtained by ultrasonic exfoliation. The high-resolution TEM results in Fig. S7b and S8b (ESI†) indicate that SPCOF and SHCOF share similar structural characteristics, with their nanosheets typically stacking in random orientations. This is reflected in the FFT patterns, which show sharp, concentric ring patterns, suggesting that these nanosheets have the same crystal structure but stack with isotropic orientations. The measured first-order reflection ring spacings for SPCOF and SHCOF are 0.39 nm−1 and 0.38 nm−1, respectively, consistent with the (100) series crystal planes determined by PXRD.
Although SPCOF and SHCOF show similar structure skeleton in the FFT analysis, the slight structural characteristic caused by the superhydrophobic perfluorohexyl chains can be distinguished in high-resolution TEM (HRTEM) images in real space. As shown in the selected area HRTEM image in Fig. 3a, well-aligned single-crystal regions were observed in certain small areas of SPCOF, captured at approximately 80 K with a cumulative electron dose below 100 e− Å−2. The high-resolution micrograph along the [001] axis indicate that the nanopores in SPCOF are arranged in a triangular pattern. The (100) facets are marked by orange lines and have an interplanar spacing of 2.55 nm, corresponding to the reflection spot indicated by a blue circle in the FFT (inset). The magnified HRTEM image of SPCOF and the overlaid atomic structure model in Fig. 3b provide a clearer depiction of the pore structure formed by interlaced carbon atom chains. Simulated HRTEM image using the SPCOF model at an under focus value of 750 nm matches well with the experimental one (Fig. 3c and Fig. S11, ESI†). As a comparison, the experimental HRTEM images of SHCOF, shown in Fig. 3d–f and Fig. S12 (ESI†), show structural pore periodicity similar to SPCOF. However, SHCOF exhibits a slightly larger interplanar spacing of 2.6 nm and a pore diameter of approximately 1.7 nm, while 1.4 nm in SPCOF. This is strong structural evidence that the perfluorohexyl chains have modify the morphology of the pores, consistent with our previous explanations. In addition, low-temperature electron energy loss spectroscopy (cryo-EELS) confirmed the presence of perfluorohexyl chains in SPCOF, as evidenced by the characteristic edges of C, N, O, and F (the fingerprint of perfluorohexyl chains), as shown in Fig. 3g with uniform distribution Fig. S9c (ESI†). In contrast, the EELS spectrum and elemental maps of SHCOF demonstrate the absence of F signal (Fig. 3h and Fig. S10c, ESI†). The HRTEM-derived pore diameter of SPCOF agrees well with that determined by BET tests. The smaller pore size in SPCOF is due to the perfluorohexyloxy chains integrated within the COF matrix, resulting in a smaller BET surface area relative to SHCOF. Both COFs manifested pore sizes below the Debye length of Zn ions in the electrolyte.
X-ray absorption near-edge structure (XANES) and Raman spectroscopy were utilized to examine the functionality of perfluorohexyl chains in SPCOFs and the Zn2+ solvation configuration in cycled electrodes (Fig. 4b and Fig. S20, ESI†). Pre-cycling electrodes were encased within battery cases sealed with Kapton tape to prevent water vaporization (Fig. S21, ESI†). XANES analysis revealed a shift towards lower energy levels in the electrolytes within COFs compared to pristine ZnSO4, indicating inhibited electron transfer from Zn to oxygen in water, which was supported by k3-weighted extended X-ray absorption fine structure (EXAFS) analysis (Fig. 4c). The primary Zn–O peak at 1.66 Å showed a slight redshift, confirming reduced oxygen coordination around Zn2+.27,28 The weak Zn–O interactions in the SPCOF layers reflects the desolvation of hydrated Zn ions, as corroborated by redshifted Raman peaks around 990 cm−1, indicative of lower water content in Zn solvation structures.29,30 MD simulations provided atomic-level insights into the structure of the solvation shell around the Zn ions, with radial distribution function (RDF) profiles and coordination numbers detailed in Fig. 4d and e and Fig. S22 and S23, ESI.† Confined electrolytes in SPCOF showed a pronounced peak at approximately 2.1 Å, indicating desolvation of Zn2+ from water, with significant reductions in coordination number and binding energy, enhancing the dehydration propensity of electrolytes within SPCOF channels.31,32 Electrochemical stability windows (ESW) evaluated in Fig. S24 (ESI†) showed that the SPCOF layer improved the ESW of ZnSO4 electrolyte to 3.1 V, significantly inhibiting side reactions like HER and OER, thus enhancing zinc battery cycle stability.
Fresistance = Fviscous + Ffriction | (2) |
![]() | (3) |
The reduction in viscous forces (Fviscous) of the electrolyte in SPCOF is reflected by a decrease in the statistical binding energy between the COF skeleton and the electrolyte species (Zn2+, SO42+, and H2O), detailed in Fig. 5g. The binding energies within the SPCOF-confined electrolyte are 1.24, 0.21, and 1.10 kcal mol−1, respectively, which are significantly lower than those observed in SHCOF, facilitating lower resistane ion migration. The potential mean force (PMF) analysis for Zn ions traversing few-layered COFs shows that they face a lower free energy barrier in SPCOF than in SHCOF, aided by fluorine chains that guide Zn ion migration along lower resistance pathways (Fig. 5h and i). This effect is further supported by tracking the mean square displacements (MSD) of Zn2+, SO42−, and H2O within various COF channels (Fig. 5j and Fig. S28, ESI†), revealing notably higher diffusion rates in SPCOF layers, thereby demonstrating the structural and chemical properties that enhance mobility.
To evaluate the utilization and sustainability of the SPCOF@Zn anode, nucleation overpotentials on Ti foil with various modifications were measured. The SPCOF@Ti electrode showed a nucleation overpotential of 29 mV, lower than bare Ti (46 mV) and SHCOF@Ti (35 mV), as shown in Fig. 6d. In contrast, the bare Zn-Ti cell demonstrated a low average Coulombic efficiency (ACE) of 94.07% and failed before reaching 100 cycles at a current density of 1 mA cm−2 and capacity of 1 mA h cm−2 (Fig. 6e). The introduction of SHCOF interfaces extended the lifespan to over 500 cycles with 98.72% ACE, while SPCOF interphases improved longevity to 1000 cycles with an ACE of 99.5%. Additionally, the SPCOF-enhanced battery exhibits a reduced overpotential of 51.7 mV and stable cycling for over 2000 cycles with an ACE of 99.8% at a higher current density of 10 mA cm−2 (Fig. S32, ESI†). The homogeneous deposition of Zn facilitated by the SPCOF interphase was observed in situ by optical microscopy, SEM, chronoamperometry, XRD, and XPS analyses. The SPCOF@Zn anode displayed ultra-dense and uniform zinc deposition, favoring the crystallization of Zn films with faceted Zn (002) plane (Fig. S33–S42, ESI†).
Long-term galvanostatic cycling experiments were conducted on Zn||Zn symmetric cells to assess the stability of Zn anodes. The bare Zn cell, operating at a current density of 1 mA cm−2 and a capacity of 1 mA h cm−2, failed after just 330 hours due to a short-circuit (Fig. 6f). In contrast, a Zn battery equipped with SHCOF exceeded 700 hours of operation. Remarkably, introducing the SPCOF layer further enhanced performance, achieving over 5000 hours of stable cycling, demonstrating the SPCOF layer's capability to stabilize the Zn-electrolyte interface and effectively guide Zn deposition. The decrease in overpotential with cycles can be attributed to the increased surface area of the Zn electrode, which occurs during repeated cycling.40 Notably, the cumulative capacity enabled by the SPCOF@Zn symmetric cell are much higher than most of the previously reported values from Zn electrodes based on different modification strategies (Fig. 6g, see Table S1 for details, ESI†). Even under increased current densities of 10 mA cm−2, the SPCOF@Zn anode maintained this exceptional longevity and stability, whereas the bare Zn anode lasted less than 80 hours, showing significant voltage hysteresis (Fig. S43, ESI†). Additional tests at a high depth of discharge (50%) using thin Zn foils (20 μm) revealed that SPCOF@Zn anodes greatly improved cycling stability, enduring 350 hours, confirming the SPCOF@Zn electrode's suitability for prolonged lifespan applications under low N/P ratio conditions (Fig. S44, ESI†).
Using a high-loading ZVO cathode, the SPCOF@Zn||ZVO battery demonstrated exceptional cycling stability under rigorous conditions, sustaining over 800 and 500 cycles at N/P ratios of approximately 4 and 2, respectively (Fig. 7a and Fig. S48, ESI†), which outperforms most of the previous reported AZBs (Fig. 7b, see Table S2 (ESI†) for details). Note that variations in the initial capacity of high-loading Zn||ZVO full cells are due to slight differences in loading between cells. Additionally, the gradually increasing capacity during the initial hundreds of cycles can be attributed to improved wetting, which enhances ionic transport and electrode utilization. The scalability of the SPCOF@Zn anode was further proven in larger configurations. A 3 × 4 cm2 single-layer pouch cell maintained stable cycling for over 200 cycles at a current density of 4 mA cm−2 (0.2 A g−1), while the 7 × 8 cm2 SPCOF@Zn||ZVO pouch cells achieved capacities of 0.3 A h and 0.55 A h by employing one-layer and two-layer at a current density of 4 mA cm−2, respectively. These results highlight the enhanced performance and versatility of the SPCOF@Zn anode.
Besides aqueous Zn battery, the SPCOF modification layer also can be used as an anode modifier in non-aqueous Li batteries. SPCOF significantly enhances Li deposition, resulting in a denser and smoother morphology (Fig. S51a–c, ESI†). Additionally, the SPCOF@Li electrode shows superior performance compared to the bare Li anode, featuring a reduced nucleation overpotential of 28 mV, a higher average coulombic efficiency of 99.97%, and an extended lifespan of over 2000 hours for the Li metal anode (Fig. S51d–f, ESI†). To further demonstrate its practical application in lithium-based battery systems, we also evaluated the performance of Li–S and Li–CO2 batteries incorporating the SPCOF modification layer (Fig. S52, ESI†). For the Li–S battery, the SPCOF layer facilitated improved cycling stability over 500 cycles, with a capacity decay of just 0.066% per cycle. In the Li–CO2 battery, the modification layer enhanced energy efficiency, reducing overpotentials to 0.72 V, 1.10 V, 1.38 V, and 1.61 V at current densities of 100 mA g−1, 200 mA g−1, 500 mA g−1, and 1000 mA g−1, respectively. Additionally, Na anode with the SPCOF modification layer demonstrated improved cycling stability, as shown in Fig. S53 (ESI†). These results highlight the versatility of the SPCOF layer as an anode modifier for wide-ranging energy storage system operating in both aqueous and non-aqueous electrolytes.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ee00132c |
‡ These authors contributed equally: Zhang Kun, Yuan Yijia and Wang Gang. |
This journal is © The Royal Society of Chemistry 2025 |