Rechargeable zinc-air batteries have attracted much attention because of their high energy density, low cost, and good eco-friendliness [1-5]. However, the sluggish kinetics at the air cathode resulting from the high overpotential for oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) become the main challenge hindering the commercial application of ZABs [6-13]. Currently, Pt/Ru-based catalysts are still known as the most effective commercial electrocatalysts for ORR and OER [14-19]. Still, the high cost and low cyclic stability have prevented large-scale commercialization in metal-air batteries. Therefore, it is essential to develop economic and high-performance catalysts for scalable implementation of zinc-air batteries.
Recent studies highlighted that transitional metal-N-doped carbon (NC) nanohybrids (M-N-C, M = Fe, Co, Ni, etc.) might hold promise as substitutes for noble metal electrocatalysts in both acidic and alkaline mediums [20-24]. Nevertheless, these monofunctional non-noble transition metals as the catalytic materials leading to single efficient electrocatalytic activity, either the ORR or OER process, dramatically limiting their electrocatalytic properties [25, 26]. Until recently, many excellent strategies for improving electrocatalytic performance have been reported through the development of carbon-based or loaded catalysts based on alloys, carbides, oxides, and sulfides/selenides [27-30]. As it is well known, the particle size is critical for the use of nanoparticles in heterogeneous catalysis [31-35]. A reduction in particle size usually leads to a significant increase in reactivity due to the higher surface-to-volume ratio of the active species [36]. For the applicability of bifunctional catalytic activity, it is more desirable that the carbon substrate has a large number of tiny metal particles rather than a large size of individual metal particles. Therefore, achieving an efficient and stable electrocatalyst that prevents the severe aggregation of metal particles and disperses uniformly on the carbon substrate during the reaction is essential.
Hollow structure is a unique structural material, the interior provides a larger surface area for more electrochemically active sites, better access to the electrolyte, faster diffusion paths for ions and electrons, and has the advantage of being a perfect electrocatalyst [37-45]. The double-shelled hollow structure provides a larger surface area in terms of electrocatalysis compared to a single-shelled hollow structure, making better use of the internal space, and also increasing mechanical stability, with the middle layer supporting each other [38, 46-50]. Therefore, the researchers concluded that the double-shelled hollow catalysts are more stable in cycling and have higher catalytic activity per unit area [51, 52]. Usually, the layer-by-layer approach using rigid templates is an effective means of constructing complex hollow structures [53]. The rigid template method and the microemulsified soft template method are currently the most widely used methods for the preparation of double-shelled hollow materials [54-58]. However, the main obstacle to their practical application is that the process of synthesizing composite hollow catalysts is both laborious and complex. In addition, these techniques often face challenges such as coating removal processes, single chemical composition, and compatibility of the selected template with the target material. Not many direct and general synthesis techniques have been reported for preparing complex double-shelled hollow structures while achieving tunability of the nanocomposition and maintaining the structural compatibility of the different hollow shells. Therefore, there is an urgent need to find an effective and facile synthesis method to obtain double-shelled hollow nanomaterials with tunable composition and functionality. Herein, we report a general approach to obtaining nitrogen-doped double-shelled carbon nanocages (Co NPs/Fe SAs DSCNs) with atomically dispersed Fe single atoms (Fe SAs) and tiny Co nanoparticles (Co NPs) co-anchored through in situ polymerization and pyrolysis.
Starting from the core-shell structure ZIF-67@ZIF-8, the double-shelled porous MOF-polymer composite nanocontainers (Fe/ZIF-67@ZIF-8@PPy) were obtained by proton hydrogen etching from pyrrole release with solvothermal reaction, followed by pyrolysis. The as-obtained Co NPs/Fe SAs DSCNs demonstrated potential and superior performance as a highly efficient bifunctional electrocatalyst also in rechargeable zinc-air batteries. In this unique structure, N-doped carbon loaded with Co NPs derived from ZIF-67 served as the inner shell, in which the embedded tiny Co NPs endowed the catalyst with more active sites and thus effectively accelerated the OER process. The ZIF-8 derived outer shell of N-doped hollow carbon anchored to Fe single atoms facilitates electrocatalytic ORR efficiently. Meanwhile, the double-shelled hollow carbon nanoreactor facilitates the exposure of more active sites, enables facile mass transfer, prevents the aggregation and corrosion of metal atoms, and ensures the durability of the catalyst. These characters endow Co NPs/Fe SAs DSCNs with superior ORR/OER catalytic activities (half-wave potential (E1/2) of 0.84 V and overpotential (ηj=10) of 346 mV) and stability in alkaline electrolyte. The assembled liquid Zn-air battery based on Co NPs/Fe SAs DSCNs exhibited high open-circuit potential (OCP) (1.45 V) and power density (71.3 mW/cm2), as well as long-term cycling stability, even precious metal catalyst of Pt/C and RuO2, indicating promising application in energy storage devices.
The synthetic strategy of Co NPs/Fe SAs DSCNs is schematically illustrated in Scheme 1. The double-shelled porous MOF-polymer composite nanomaterials (Fe/ZIF-67@ZIF-8@PPy) obtained by a simple solvothermal method consisting of the inner shell of ZIF-67-derived Fe-Co-PPy and the outer shell of ZIF-8-derived Fe-ZIF-8 PPy, could be further transformed into Co NPs/Fe SAs DSCNs after annealing treatment. Firstly, the ZIF-67@ZIF-8 core-shell structure was fabricated by seeded epitaxial growth method due to ZIF-8 (a = b = c = 16.9910 Å) [59] and ZIF-67 (a = b = c = 16.9589 Å) [60] with similar topological structure and unit cell parameters, and the color change of the product visually indicates the successful growth of ZIF-8 on ZIF-67 crystal (Fig. S1 in Supporting information). X-ray diffraction (XRD) analysis showed that the resulting ZIF-67@ZIF-8 core-shell structure had correspondingly similar features and high crystallinity to ZIF-8 and ZIF-67 crystals (Fig. S1). Thereafter, the precursor ZIF-67@ZIF-8 was dispersed within a methanol solution followed by the introduction of pyrrole monomers, which could be pre-filled evenly in the pores of the MOF templates. By adding two Fe salts (Fe(acac)3 & FeCl3), the ZIF-67@ZIF-8 templates were gradually disassembled synchronously with the polymerization of pyrrole monomers within a methanol solution to form Fe/ZIF-67@ZIF-8@PPy. Since Fe(acac)3 has a large conjugated structure (9.7 Å) and cannot enter the pores of MOFs (3.2 Å) [61], and it mainly acts as a dopant and initiator to induce the polymerization of pyrrole monomers on the outer surface of MOFs during the reaction process [20, 62], and the proton H released from the polymerization of pyrrole monomers will preferentially etch the inner ZIF-67 nanostructures (due to the poorer stability of ZIF-67 in acidic solutions), the released cobalt and iron co-precipitate to form the outer shell of Fe-Co PPy (Figs. S5a and b in Supporting information). At the same time, FeCl3 not only triggers the polymerization of pyrrole on the surface of ZIF-8, but also diffuses into the pores of MOFs due to its small size, triggering the polymerization of pyrrole monomers to form the inner shell on the outer surface of ZIF-67 until the outer shell of polypyrrole is thicker and denser, which stops the diffusion of FeCl3 and pyrrole monomers to the inside, thus terminating the reaction, and finally forming the complete Fe/ZIF-67@ZIF-8@PPy composite nanostructure [20]. Subsequent to the annealing treatment under Ar, the Co NPs/Fe SAs DSCNs catalysts were obtained. Thus, the difficulty of forming multi-shelled hollow structures with different shell materials was effectively solved by in situ disassembling-polymerization pyrolysis.
|
Download:
|
| Scheme 1. Schematic illustration of the synthetic strategy of Co NPs/Fe SAs DSCNs. | |
As shown in the field-emission scanning electron microscopy (FESEM) and transmission electron microscopy (TEM) images (Figs. S2a and b in Supporting information), the obtained ZIF-67@ZIF-8 particles exhibited a uniform dodecahedron shape and an average particle size of about 800 nm. X-ray spectroscopy (EDS) elemental mapping images (Fig. S2c in Supporting information) further revealed that C and N were distributed over the entire sample, but Zn and Co were mainly located at the interior and outside of the sample, which suggested the successful synthesis of the core−shell structure of ZIF-67@ZIF-8. After polymerization, the prepared Fe/ZIF-67@ZIF-8@PPy with a rougher surface than that before polymerization inherited the dodecahedron-shaped structure of the ZIF-67@ZIF-8 templates, which confirmed the polymerization of pyrrole on the surface of ZIF-8 (Fig. S6 in Supporting information). The Fe/ZIF-67@ZIF-8@PPy exhibited a hollow structure with the part of the inner ZIF-67 core being much clearer hollow structure than the outer ZIF-8 part (Figs. S3a and b in Supporting information). Elemental mapping images showed that Fe, Co, Zn, C, and N were distributed throughout the sample (Fig. S3c in Supporting information). After the pyrolysis process, Figs. 1a and b clearly showed that Co NPs/Fe SAs DSCNs still maintain the original structure, double-shelled structures can be observed clearly, where the gaps between the outer and inner shells and the connections between them by interconnected carbon structures can be seen clearly in each particle, thus effectively mitigating the strains that may be generated during the calcination process leading to the collapse of the shells. A shell-in-shell structure can be discerned from the broken particles (Fig. 1b, inset). Such double-shelled characteristics can highly facilitate the mass transfer process during electrocatalysis [48, 52]. Within the carbon matrix, no Fe nanoparticles could be found in the high-resolution TEM (HRTEM) image. In contrast, atomically dispersed Fe sites were detected by using aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) in Figs. 1c and d. The bright dots marked with orange circles could be assigned to isolated metal atoms. HRTEM (Fig. 1e) observations revealed a set of visible lattice stripes with a crystal plane spacing of 0.24 nm, corresponding to the (111) crystal plane of Co NPs [29]. The cobalt embedded in the Co NPs/Fe SAs DSCNs had a smaller size of about 12 nm compared to Fe/ZIF-67@PPy-800, consistent with TEM images (Fig. 1a and Fig. S4 in Supporting information), suggesting that the confinement effect between the double C–N shell plays an important role in preventing the aggregation of metal atoms [35, 56]. HAADF-TEM image and the corresponding elemental mapping images demonstrated that Fe (yellow), Co (purple), and N elements (cyan) are homogeneously distributed throughout the entire region (Fig. 1f), due to metal atom migrated during the pyrolysis process. The above results demonstrated that in situ polymerization and pyrolysis routes successfully produce the double-shelled hollow carbon nanocontainers, which consisted of Fe atoms and tiny Co NPs sites co-existing in the carbon matrix.
|
Download:
|
| Fig. 1. (a) TEM image, (b) SEM image (Inset: orange dash area enlarged SEM image), (c) aberration-corrected STEM, (d) Enlarged aberration-corrected STEM image (orange cycles marked single atom), (e) HR-TEM image, (f) HR-TEM image and elemental mapping (Fe yellow, Co purple, C green, and N cyan) of Co NPs/Fe SAs DSCNs. | |
The chemical composition and coordination environment of Co NPs/Fe SAs DSCNs were investigated by X-ray photoelectron spectroscopy (XPS) (Fig. S7a and Table S1 in Supporting information). The high-resolution C 1s XPS spectrum presented three characteristic peaks, corresponding to C–C (284.6 eV), C–N (285.1 eV), and C–O/C═C (399.7 eV) (Fig. S7b in Supporting information). The presence of the C–N bond implied the successful N doping into the C matrix, and N is mainly derived from pyrrole and residual dimethylimidazole. Meanwhile, the N 1s XPS spectrum showed five moieties, including pyridinic N (398.5 eV), pyrrolic N (399.2 eV), Fe/Co-N (400.1 eV), graphitic N (401.1 eV), and oxidized N (403.8 eV) (Fig. 2a). The co-existence of Fe and Co in Co NPs/Fe SAs DSCNs was also confirmed by Fe 2p and Co 2p spectra (Figs. 2b and c). It is believed that such a coordination interaction between Fe/Co-N was propitious to improve the catalytic activity, which determines the superior ORR/OER kinetics [63]. In addition, the Co NPs/Fe SAs DSCNs demonstrated a Brunauer–Emmett–Teller (BET) surface area of 305.08 m2/g, as presented in Fig. 2d. Meanwhile, the pore diameter distribution curves (Fig. 2d, inset) further shows that the presence of mesoporous pores with a pore size of about 5.4 nm, which is conducive to the dispersion and mass transfer of metal atoms during the catalytic process. Powder XRD patterns (Fig. 2e) of Co NPs/Fe SAs DSCNs exhibited a broad diffraction peak at 22°, corresponding to the (002) plane of carbon, indicating complete carbonization of the Co NPs/Fe SAs DSCNs nanocages. Other planes of 44.3° and 51.3° were determined to be the (111) and (200) planes of Co metal NPs (PDF#15-0806) (Fig. S8 in Supporting information), respectively, with no characteristic peaks for the crystalline Fe species, consistent with HRTEM results [29]. The ratio of the D band to the G band intensity in the Raman spectrum of the different temperature catalysts contained two peaks at approximately 1345 and 1591 cm-1, exhibited an estimated intensity ratio (ID/IG) of 1.02, 1.04, and 1.06, respectively, revealing the carbonized structure of the obtained catalysts (Fig. 2f).
|
Download:
|
| Fig. 2. High-resolution (a) N 1s, (b) Fe 2p, (c) Co 2p XPS spectra of Co NPs/Fe SAs DSCNs composites. (d) N2 adsorption-desorption isothermal curves of Co NPs/Fe SAs DSCNs. Inset: Pore size distributions of Co NPs/Fe SAs DSCNs. (e) XRD patterns of the products at different reaction stages. (f) Raman curves of samples Co NPs/Fe SAs DSCNs at different temperatures. | |
To estimate the intrinsic electrocatalytic activity of the Co NPs/Fe SAs DSCNs, the ORR and OER performance of Co NPs/Fe SAs DSCNs, in alkaline media was investigated using a typical three-electrode system. The counterparts Fe/ZIF-8@PPy-800, Fe/ZIF-67@PPy-800, commercial Pt/C, and RuO2 were also evaluated for comparison. First, the ORR behavior of the catalysts was evaluated in O2-saturated 0.1 mol/L KOH with a rotating disk electrode. No redox peak was visible in the cyclic voltammetry (CV) curves of the sample in N2-saturated KOH solution, whereas a well-defined cathodic peak was observed in the O2-saturated electrolyte, suggesting the potential ORR catalytic capability of the catalysts (Fig. S9 in Supporting information). The LSV curve tested at 1600 rpm of Co NPs/Fe SAs DSCNs exhibits a high ORR activity with a half-wave potential (E1/2) of 0.84 V versus RHE, superior to those of Fe/ZIF-8@PPy-800 (0.81 V), Fe/ZIF-67@PPy-800 (0.81 V), ZIF-67@ZIF-8-800 (0.78 V), and even comparable to commercial Pt/C (0.82 V) (Fig. 3a and Fig. S10 in Supporting information). Notably, compared to single-shelled Fe/ZIF-8@PPy-800 and Fe/ZIF-67@PPy-800, double-shelled structure Co NPs/Fe SAs DSCNs manifests outstanding catalytic performance, which was attributed to the double-shelled hollow structure endowing multiple metal sites, where the active sites in each layer could be utilized more efficiently [46]. The superior catalytic activity of Co NPs/Fe SAs DSCNs for ORR was further verified by the Tafel plots obtained from the polarization curves (Fig. 3b). The Tafel slope of Co NPs@Fe SA DSCNs was 88 mV/dec, which was lower than that of Fe/ZIF-8@PPy-800 (97 mV/dec), Fe/ZIF-67@PPy-800 (120 mV/dec), ZIF-67@ZIF-8-800 (128 mV/dec), and even similar to Pt/C (102 mV/dec), which suggests that Co NPs/Fe SAs DSCNs electrocatalysts with unique double-shelled hollow nanostructures have faster reaction kinetics. The electron transfer number (n) for ORR was determined from the LSV curves (Fig. 3c) based on the Koutechy-Levich (K-L) equation. The K-L plots for the Co NPs/Fe SAs DSCNs showed good linearity in the potential range of 0.75–0.3 V, and the electron transfer number (n) close to 4.0, indicating that involving the in situ OH- formation in an alkaline medium. In addition to the high ORR activity, the Co NPs/Fe SAs DSCNs exhibited remarkable long-term catalytic stability. After 14 h of continuous timed current, the ORR activity retention of Co NPs/Fe SAs DSCNs was much higher than that of Pt/C (89% vs. 55%, Fig. 3d). Compared with the Pt/C catalyst, the Co NPs/Fe SAs DSCNs catalyst exhibited better durability and higher resistance to the methanol crossover effect under 0.1 mol/L KOH oxygen saturation conditions (Fig. S11 in Supporting information).
|
Download:
|
| Fig. 3. Electrocatalytic performance of ORR and OER of the catalysts. (a) ORR polarization curves of the as-prepared catalysts in O2-saturated 0.1 mol/L KOH at 1600 rpm. (b) Tafel plot for ORR. (c) ORR polarization curves of Co NPs/Fe SAs DSCNs at different scan rates (Inset: corresponding K–L plot). (d) Chronoamperoometric measurements at 0.7 V vs. RHE. (e) OER polarization curves of all samples in 1.0 mol/L KOH. (f) Tafel plot for OER. (g) Electrochemical impedance spectroscopy plots at 1.20 V vs. RHE for OER. (h) Chronoamperoometric measurements at 1.6 V vs. RHE. (i) OER potential at a current density of 10 mA/cm2 (Ej=10), ORR half-wave potential (E1/2), and their difference (ΔE) of prepared catalyst. | |
The Co NPs/Fe SAs DSCNs not only showed high electrocatalytic activity for ORR but also exhibited excellent electrocatalytic activity for OER. The OER activities of different catalysts were further evaluated in 1 mol/L KOH electrolyte. The overpotential of Co NPs/Fe SAs DSCNs at a current density of 10 mA/cm2 was only 346 mV. It was superior to Fe/ZIF-67@PPy-800 (400 mV), ZIF-67@ZIF-8-800 (410 mV), and even similar to that of RuO2 (280 mV) under the same test conditions (Fig. 3e). It indicates that the Co NPs/Fe SAs DSCNs samples have excellent OER electrocatalytic activity. It is noteworthy that the OER performance of the catalyst Fe/ZIF-8@PPy-800 without Co element is very weak, which suggests that Co NPs are the main active sites for OER [29]. In addition, the comparison of catalysts Fe/ZIF-67@PPy-800 and ZIF-67@ZIF-8-800 demonstrated that the unique double-shelled hollow nanocavity could provide more exposed electrically active sites for the chemisorption of reactants and subsequent reactions, resulting in excellent OER catalytic activity in alkaline media [47]. The OER electrocatalytic kinetics of different samples was evaluated by the slope of the Tafel curve. According to the data shown in Fig. 3f, the corresponding Tafel slope of Co NPs/Fe SAs DSCNs was fitted to 107 mV/dec, which is significantly smaller than that of Fe/ZIF-67@PPy-800 (122 mV/dec), ZIF-67@ZIF-800 (146 mV/dec), and comparable to RuO2 (85 mV/dec), showing outstanding OER kinetics. This reveals that Co NPs/Fe SAs DSCNs possesses better electron transport synergies and can accelerate the OER-related proton transport process. Also, it can be seen from the Nyquist plot (Fig. 3g) that Co NPs/Fe SAs DSCNs have smaller semicircle diameters compared to the other catalysts, indicating that Co NPs/Fe SAs DSCNs catalysts have the highest charge transfer efficiency during the electrochemical process. The Tafel and Rct results collectively demonstrated that the Co NPs/Fe SAs DSCNs ability to catalyze oxygen evolution, suggesting that it can be used as a promising bifunctional electrode material for rechargeable metal-air batteries. The electrochemical double-shelled capacitance (Cdl), which is proportional to the electrochemically active surface area (ECSA), was further determined by measuring cyclic voltammetric curves in the potential range of 1.074–1.174 V in the absence of redox processes. The Co NPs/Fe SAs DSCNs had higher Cdl compared to Fe/ZIF-8@PPy-800, Fe/ZIF-8@PPy-800, and ZIF-67@ZIF-8-800 (Figs. S12 and S13 in Supporting information), which further indicates that the catalysts have higher electrochemically active area and specific surface area, which are more favorable for the catalytic reaction. In addition, the as-prepared Co NPs/Fe SAs DSCNs demonstrates the stability of electrochemical performance for OER, even outperforming the commercial RuO2 (Fig. 3h and Fig. S14 in Supporting information). The oxygen electrode activity (ΔE) (the difference between the ORR half-wave potential and the OER potential at 10 mA/cm2) is an important measure of the bifunctional catalytic activity of the ORR and OER. As shown in Fig. 3i, the ΔE value of Co NPs/Fe SAs DSCNs catalysts was 0.73 V, which was comparable to that of Pt/C-RuO2 (0.68 V) and far superior to the other prepared catalyst materials, and compared with other reported high-performance bifunctional catalysts (Table S2 in Supporting information), the Co NPs/Fe SAs DSCNs catalysts were excellent, which further illustrates that the synthesized Co NPs/Fe SAs DSCNs materials have efficient bifunctional catalytic activity.
The Zn-air battery performance was investigated using an assembled liquid Zn-air battery device with Co NPs/Fe SAs DSCNs as oxygen electrode catalysts. As shown in Fig. 4a, a schematic diagram of the Zn-air battery device was constructed with a Zn pole plate as the anode, carbon paper containing catalyst as the cathode of the battery, and a 6.0 mol/L KOH/0.2 mol/L Zn(Ac)2 mixture solution as the electrolyte. As shown in Fig. 4b, the Co NPs/Fe SAs DSCNs-catalyzed cell exhibited an open-circuit voltage of 1.45 V, which was higher than that of 1.40 V with the Pt/C-RuO2 catalysts. The charge/discharge polarization curves in Fig. 4c further indicated that the charge/discharge voltage gap of Co NPs/Fe SAs DSCNs also shows a narrower voltage gap than that of Pt/C-RuO2 at the same current density, which indicates that the Co NPs/Fe SAs DSCNs have a better performance. Its power density result in Fig. 4d shows that the battery of Co NPs/Fe SAs DSCNs had a power density of 71.3 mW/cm2, which is better than the Pt/C-RuO2 (57.9 mW/cm2) and other catalysts (Table S3 in Supporting information). Moreover, the cycling rechargeability for the secondary battery was further studied at a current density of 2 mA/cm2 with a duration of 30 min per cycle. After 300 cycles of testing, the charging-discharging potential gap of Co NPs/Fe SAs DSCNs without increase due to the downhill of the discharging potential, suggesting its long-lasting durability of the Co NPs/Fe SAs DSCNs cathode than the Pt/C-RuO2 (Fig. 4e). As a demonstration, two cells connected in series can easily light up the LED beads for several hours (Fig. S15 in Supporting information). These battery parameters of Co NPs/Fe SAs DSCNs based Zn-air batteries prove to have the potential to serve as clean energy storage systems.
|
Download:
|
| Fig. 4. Electrochemical performance of rechargeable zinc–air battery. (a) Schematics for the fabricated Zn-air cells under alkaline electrolyte. (b) The circuit voltage (OCV) curve of Co NPs/Fe SAs DSCNs and Pt/C-RuO2-based Zn–air batteries (inset shows the photograph of the Co NPs/Fe SAs DSCNs-based Zn–air battery with an OCP of 1.45 V). (c) Charging-discharging polarization plots. (d) Discharge polarization curves and corresponding power density curves. (e) Charge–discharge stabilities at a current density of 2 mA/cm2. | |
In summary, a facile in situ disassembly-polymerization-pyrolysis strategy is proposed to prepare N-doped double-shelled hollow carbon nanocages (Co NPs/Fe SAs DSCNs) embedded Fe SAs and tiny Co NPs binary sites. Benefiting from the high intrinsic activity of multiple metal sites well encapsulated in a double-shelled hollow carbon matrix where the active sites in each layer could be utilized more efficiently, the obtained Co NPs/Fe SAs DSCNs catalyst can serve as an efficient, economical, and durable bifunctional catalyst for ORR and OER in an alkaline environment. Moreover, zinc-air batteries utilizing Co NPs/Fe SAs DSCNs as catalysts demonstrated excellent long-term stability and charge/discharge reversibility, indicating significant potential for practical applications. These findings will provide guidance for further rational design and construction of double-shelled hollow structures with multiple active sites and a deeper understanding of synergistic effects in energy-related catalytic applications.
Declaration of competing interestThe authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
CRediT authorship contribution statementCongcong Wang: Writing – original draft, Methodology, Investigation, Data curation, Conceptualization. Kai Zhang: Writing – review & editing, Funding acquisition. Bai Yang: Supervision, Formal analysis.
AcknowledgmentThis work was supported by the National Natural Science Foundation of China (NSFC, No. 21774045).
Supplementary materialsSupplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110538.
| [1] |
S.S. Shinde, J.Y. Jung, N.K. Wagh, et al., Nat. Energy 6 (2021) 592-604. DOI:10.1038/s41560-021-00807-8 |
| [2] |
S. Han, X. Hu, J. Wang, et al., Adv. Energy Mater. 8 (2018) 1800955. |
| [3] |
Y. Xue, Y. Guo, Q. Zhang, et al., Nano Micro Lett. 14 (2022) 162. |
| [4] |
L. Yan, Y. Xu, P. Chen, et al., Adv. Mater. 32 (2020) 2003313. |
| [5] |
F. Yang, J. Xie, X. Liu, et al., Small 17 (2020) 2007085. |
| [6] |
Y. Jiao, Y. Zheng, M. Jaroniec, S.Z. Qiao, Chem. Soc. Rev. 44 (2015) 2060-2086. |
| [7] |
J. Yi, P. Liang, X. Liu, et al., Energy Environ. Sci. 11 (2018) 3075-3095. DOI:10.1039/c8ee01991f |
| [8] |
Y. Li, M. Gong, Y. Liang, et al., Nat. Commun. 4 (2013) 1805. DOI:10.3892/or.2013.2333 |
| [9] |
Y.P. Deng, Y. Jiang, R. Liang, et al., Nat. Commun. 11 (2020) 1952. |
| [10] |
X. Wu, C. Tan, C. He, et al., J. Power Sources 520 (2022) 230891. |
| [11] |
A. Yu, W. Long, L. Zhu, et al., Chin. Chem. Lett. 34 (2023) 107860. |
| [12] |
Y. Wu, M. Jing, J. Li, et al., Chin. Chem. Lett. 35 (2024) 109269. |
| [13] |
H. Wang, L. Cao, Y. Feng, et al., Chin. Chem. Lett. 34 (2023) 107601. |
| [14] |
C. Lafforgue, F. Maillard, V. Martin, et al., ACS Catal. 9 (2019) 5613-5622. DOI:10.1021/acscatal.9b00439 |
| [15] |
M. Ming, Y. Zhang, C. He, et al., Small 15 (2019) 1903057. |
| [16] |
P. Tan, B. Chen, H. Xu, et al., Energy Environ. Sci. 10 (2017) 2056-2080. |
| [17] |
X. Zheng, P. Cui, Y. Qian, et al., Angew. Chem. Int. Ed. 59 (2020) 14533-14540. DOI:10.1002/anie.202005241 |
| [18] |
X. Han, X. Wu, C. Zhong, et al., Nano Energy 31 (2017) 541-550. |
| [19] |
D. He, K. Cheng, T. Peng, et al., J. Mater. Chem. A 1 (2013) 2126-2132. |
| [20] |
C. Wang, Y. Chen, M. Zhong, et al., J. Mater. Chem. A 9 (2021) 22095-22101. DOI:10.1039/d1ta05039g |
| [21] |
L. Jiao, R. Zhang, G. Wan, et al., Nat. Commun. 11 (2020) 2831. |
| [22] |
Y. Chen, L. Shen, C. Wang, et al., Appl. Catal. B 274 (2020) 119112. |
| [23] |
Y. Chen, L. Shen, C. Wang, et al., Chem. Eng. J. 430 (2022) 132632. |
| [24] |
Y. Chen, S. Ji, Y. Wang, et al., Angew. Chem. Int. Ed. 56 (2017) 6937-6941. DOI:10.1002/anie.201702473 |
| [25] |
Y. Chen, M. Wang, S. Xiang, et al., ACS Sustain. Chem. Eng. 7 (2019) 10912-10919. DOI:10.1021/acssuschemeng.9b01789 |
| [26] |
X. Gong, J. Zhu, J. Li, et al., Adv. Funct. Mater. 31 (2020) 2008085. |
| [27] |
K. Elumeeva, M.A. Kazakova, D.M. Morales, et al., ChemSusChem 11 (2018) 1204-1214. DOI:10.1002/cssc.201702381 |
| [28] |
Y. Qiao, P. Yuan, Y. Hu, et al., Adv. Mater. 30 (2018) 1804504. |
| [29] |
N.K. Wagh, D.H. Kim, S.H. Kim, et al., ACS Nano 15 (2021) 14683-14696. DOI:10.1021/acsnano.1c04471 |
| [30] |
P. Li, F. Qiang, X. Tan, et al., Appl. Catal. B 340 (2024) 123231. |
| [31] |
Y. Lykhach, S.M. Kozlov, T. Skála, et al., Nat. Mater. 15 (2015) 284-288. |
| [32] |
W. Chen, J. Ji, X. Feng, et al., J. Am. Chem. Soc. 136 (2014) 16736-16739. DOI:10.1021/ja509778y |
| [33] |
K. Murata, Y. Mahara, J. Ohyama, et al., Angew. Chem. Int. Ed. 56 (2017) 15993-15997. DOI:10.1002/anie.201709124 |
| [34] |
H. Wang, J. Lu, Chin. J. Chem. 38 (2020) 1422-1444. DOI:10.1002/cjoc.202000205 |
| [35] |
H. Chen, K. Shen, Q. Mao, et al., ACS Catal. 8 (2018) 1417-1426. DOI:10.1021/acscatal.7b03270 |
| [36] |
P. Liu, R. Qin, G. Fu, N. Zheng, J. Am. Chem. Soc. 139 (2017) 2122-2131. DOI:10.1021/jacs.6b10978 |
| [37] |
L. Yu, H.B. Wu, X.W.D. Lou, Acc. Chem. Res. 50 (2017) 293-301. DOI:10.1021/acs.accounts.6b00480 |
| [38] |
L. Yu, H. Hu, H.B. Wu, X.W. Lou, Adv. Mater. 29 (2017) 1604563. |
| [39] |
B.Y. Guan, L. Yu, J. Li, X.W. Lou, Sci. Adv. 2 (2016) e1501554. |
| [40] |
B.Y. Guan, L. Yu, X.W. Lou, Adv. Mater. 28 (2016) 9596-9601. DOI:10.1002/adma.201603622 |
| [41] |
K. Huang, Y. Sun, Y. Zhang, et al., Adv. Mater. 31 (2018) 1801430. |
| [42] |
S. Zhan, X. Chen, B. Xu, et al., Nano Today 47 (2022) 101626. |
| [43] |
K. Wang, L. Xu, J. Wang, et al., Green Energy Environ. 8 (2023) 1154-1160. |
| [44] |
X. Chen, N. Yang, Y. Wang, et al., Adv. Mater. 34 (2021) 2107400. |
| [45] |
X. Chen, S. Liu, N. Yang, et al., EcoMat 5 (2023) e12348. |
| [46] |
Z. Huang, H. Pan, W. Yang, et al., ACS Nano 12 (2018) 208-216. DOI:10.1021/acsnano.7b05832 |
| [47] |
J. Zhang, L. Yu, Y. Chen, et al., Adv. Mater. 32 (2020) 1906432. |
| [48] |
P. Zhang, B.Y. Guan, L. Yu, X.W. Lou, Angew. Chem. Int. Ed. 56 (2017) 7141-7145. DOI:10.1002/anie.201702649 |
| [49] |
H. Hu, B.Y. Guan, X.W. Lou, Chem 1 (2016) 102-113. |
| [50] |
Z.S.Z.Y.Y.N.W. Dan, Chem. J. Chin. Univ. 42 (2021) 333-348. |
| [51] |
X.Y. Yu, L. Yu, L. Shen, et al., Adv. Funct. Mater. 24 (2014) 7440-7446. DOI:10.1002/adfm.201402560 |
| [52] |
H. Hu, B. Guan, B. Xia, X.W. Lou, J. Am. Chem. Soc. 137 (2015) 5590-5595. DOI:10.1021/jacs.5b02465 |
| [53] |
X. Chen, P. Li, J. Wang, et al., Nano Res. 15 (2022) 4117-4123. DOI:10.1007/s12274-021-4063-y |
| [54] |
L. Ren, K. Sun, Y. Wang, et al., Adv. Mater. 36 (2024) 2310547. |
| [55] |
L. Li, B. Chen, Z. Zhuang, et al., Chem. Eng. J. 399 (2020) 125647. |
| [56] |
X. Liu, L. Gong, L. Wang, et al., Nano Lett. 24 (2023) 592-600. |
| [57] |
G. Peng, J. Zhou, J. Tao, et al., Nano Res. 17 (2023) 3164-3174. |
| [58] |
J. Cui, Y. Wang, A. Postma, et al., Adv. Funct. Mater. 20 (2010) 1625-1631. DOI:10.1002/adfm.201000209 |
| [59] |
Z.N.K.S. Park, A.P. Coˆte´, J.Y. Choi, et al., Proc. Nat. Acad. Sci. U. S. A. 103 (2006) 10186-10191. DOI:10.1073/pnas.0602439103 |
| [60] |
A.P.R. Banerjee, B. Wang, C. Knobler, et al., Science 319 (2008) 939-943. DOI:10.1126/science.1152516 |
| [61] |
Y.J. Chen, S.F. Ji, Y.G. Wang, et al., Angew. Chem. Int. Ed. 56 (2017) 6937-6941. DOI:10.1002/anie.201702473 |
| [62] |
Y. Ye, F. Cai, H. Li, et al., Nano Energy 38 (2017) 281-289. |
| [63] |
Z. Wang, X. Jin, C. Zhu, et al., Adv. Mater. 33 (2021) 2104718. |
2025, Vol. 36 

