Cobalt loaded on concave hollow carbon octadecahedron for zinc–air batteries

The energy crisis and environmental pollution have put forward high requirements on renewable clean energy storage and conversion technologies.11. J. Wang, F. Xu, H. Jin, Y. Chen, Y. Wang, J. Wang, F. Xu, H. Y. Jin, Y. Q. Chen, and Y. Wang, Adv. Mater. 29, 1605838 (2017). https://doi.org/10.1002/adma.201605838 Rechargeable zinc–air batteries have, thus, received extensive attention.2–62. M. Wang, H. Zheng, Y. Dong, Z. Jin, S. Liao, W. Gong, X. Li, Z. Li, M. Xia, and J. Wang, Appl. Phys. Lett. 120, 261903 (2022). https://doi.org/10.1063/5.00887863. M. Wu, G. Zhang, N. Chen, Y. Hu, T. Regier, D. Rawach, and S. Sun, ACS Energy Lett. 6, 1153 (2021). https://doi.org/10.1021/acsenergylett.1c000374. A. Liu, X. Liang, X. Ren, W. Guan, and T. Ma, Electrochem. Energy Rev. 51, 112 (2021). https://doi.org/10.1007/S41918-021-00110-W5. Z. Zhang, B. Xi, X. Ma, W. Chen, J. Feng, and S. Xiong, SusMat 2, 114 (2022). https://doi.org/10.1002/sus2.536. H. Luo, B. Liu, Z. Yang, Y. Wan, and C. Zhong, Electrochem. Energy Rev. 51 5, 187 (2021). https://doi.org/10.1007/S41918-021-00107-5 However, their practical application is limited by the slow kinetics of the four-electron multi-step reaction in the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER).77. X. Zou and Y. Zhang, Chem. Soc. Rev. 44, 5148 (2015). https://doi.org/10.1039/C4CS00448E Traditional noble metal electrocatalysts (Ru, Ir, Pt, etc.) have high electrochemical activity.88. J. Song, S. Qiu, F. Hu, Y. Ding, S. Han, L. Li, H. Y. Chen, X. Han, C. Sun, and S. Peng, Adv. Funct. Mater. 31, 2100618 (2021). https://doi.org/10.1002/adfm.202100618 Therefore, it is quite challenging to develop non-noble metal electrocatalysts with comparable activity for ORR and OER in order to overcome the problems of high price and selectivity of noble metal catalysts.9,109. L. Li, B. Wang, G. Zhang, G. Yang, T. Yang, S. Yang, and S. Yang, Adv. Energy Mater. 10, 2001600 (2020). https://doi.org/10.1002/aenm.20200160010. J. Fu, R. Liang, G. Liu, A. Yu, Z. Bai, L. Yang, Z. Chen, J. Fu, R. Liang, G. Liu, A. Yu, Z. Chen, Z. Bai, and L. Yang, Adv. Mater. 31, 1805230 (2019). https://doi.org/10.1002/adma.201805230Metal organic framework (MOF) has great potential in catalysis as catalyst support or as catalyst itself owing to its porousness and large specific surface area.1111. Z. X. Cai, Z. L. Wang, Y. J. Xia, H. Lim, W. Zhou, A. Taniguchi, M. Ohtani, K. Kobiro, T. Fujita, and Y. Yamauchi, Angew. Chem., Int. Ed. 60, 4747 (2021). https://doi.org/10.1002/anie.202010618 ZIF-67 is a cobalt-based MOF with adjustable pore size and highly stable structure.12,1312. D. Fairen-Jimenez, S. A. Moggach, M. T. Wharmby, P. A. Wright, S. Parsons, and T. Düren, J. Am. Chem. Soc. 133, 8900 (2011). https://doi.org/10.1021/ja202154j13. J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen, and J. T. Hupp, Chem. Soc. Rev. 38, 1450 (2009). https://doi.org/10.1039/b807080f However, such advantages will be reduced in the preparation of catalyst, where high temperature treatment can cause the aggregation of cobalt atoms and the collapse of MOF structure.1414. I. Saana Amiinu, Z. Pu, X. Liu, K. Asare Owusu, H. Gabriela Rivera Monestel, F. Ofori Boakye, H. Zhang, S. Mu, I. S. Amiinu, Z. Pu, X. Liu, H. R. Monestel, F. O. Boakye, H. Zhang, S. Mu, and K. A. Owusu, Adv. Funct. Mater. 27, 1702300 (2017). https://doi.org/10.1002/adfm.201702300 Therefore, there is a great demand for a reasonable design of ZIF-67 from the perspective of maintaining geometric structure to realize high-efficiency catalysts.1515. I. Saana Amiinu, X. Liu, Z. Pu, W. Li, Q. Li, J. Zhang, H. Tang, H. Zhang, S. Mu, I. S. Amiinu, X. Liu, Z. Pu, W. Li, J. Zhang, H. Tang, H. Zhang, S. Mu, and Q. Li, Adv. Funct. Mater. 28, 1704638 (2018). https://doi.org/10.1002/adfm.201704638

This work prepared a porous hollow carbon octadecahedron by introducing g-C3N4 quantum dots (GQDs) during the synthesis process of ZIF-67. Through modulating the force balance of the framework structure in the heat treatment by g-C3N4, the polyhedral morphology of ZIF-67 can be retained to obtain a material with a rich microporous structure and large specific surface area, helping to expose more catalytic active sites for efficient electrocatalysis.

Figures 1(a) and S1 briefly introduces the synthesis of the Co-based nanobox (CNQD/CoNB) catalyst and GQDs [Figs. S2(c) and S2(d)], respectively. As shown in Fig. 1(a), the cobalt-based zeolite imidazole framework and GQDs are self-assembled by the surfactant-assisted method, followed by pyrolysis in an N2 atmosphere. After pyrolysis, an octadecahedron catalyst with concave facets is formed. It will be interesting to explain why the flat planes of the octadecahedron became concaves after pyrolysis. Figure 1(b) illustrates the structural changes during the thermal annealing process. The structure change is governed by several forces. On each surface, there are two forces in opposite directions, which are fstretch and fshrink. fstretch makes the surface expand, and fshrink makes the surface collapse, whereas Fstretch is the resultant force of two fstretch in the direction of the vertex of two adjacent surfaces. Determined by the included angle of the two surfaces, the value of Fstretch is 1.85 times that of fstretch. ZIF-67 shrinks greatly under fshrink during the pyrolysis. The size collapsed from 800 nm to about 100 nm (Figs. S4 and S5) due to the lack of fstretch during pyrolysis. The situation for the CNQD/CoNBs that wraps g-C3N4 inside the MOF structure is different. The shrink will happen due to the pyrolysis of ZIF-67, while fstretch comes out due to the expansion caused by the decomposition of g-C3N4 at the same time. For the CNQD/CoNBs, the size change before and after pyrolysis is not so serious, only from 1200 to 900 nm (Figs. S3 and S6). This is explained by the holding effect of GQDs. During the pyrolysis, the surface layer of CNQD/CoNBs will be graphitized into carbon layer, which has different thermal expansion property (TEP) with GQDs inside. According to the TGA analysis (Fig. S11), g-C3N4 is basically stable at the stage from 0 to about 600 °C, while the TEP of the ZIF-67 at this stage is higher than g-C3N4. Compared with pure ZIF-67, the thermal shrinkage of the CNQD/CoNBs is greatly reduced due to the lower TEP of g-C3N4. As the temperature exceeds 600 °C, the thermal decomposition of g-C3N4 accelerates rapidly, and the TEP is also increased significantly. In this second stage, the fshrink of the surface will be greater than the fstretch, and the stress on the edges and vertices of the polyhedron is larger than the surface, which will result in anisotropic shrinkage.1616. W. Xiang, J. Li, J. Ma, Z. Sheng, H. Lu, and S. Yang, New J. Chem. 45, 22787–22797 (2021). https://doi.org/10.1039/D1NJ03900H The concave morphology can activate the Co–Nx part, which is inaccessible in the non-concave particles, and increase the surface area of the particles, thus, expanding the reaction interface and increasing the reaction kinetics.1717. X. Wan, X. Liu, Y. Li, R. Yu, L. Zheng, W. Yan, H. Wang, M. Xu, and J. Shui, Nat. Catal. 23(2), 259 (2019). https://doi.org/10.1038/s41929-019-0237-3The typical field emission scanning electron microscope (SEM) images in Figs. 2(a) and S3 show that CNQD/CoNBs with uniform size distribution have been synthesized. As shown in Figs. 2(d), S4, and S5, the ZIF-67 remained a typical hexahedron structure after pyrolysis. In the formation of ZIF-67, the introduced GQDs act as a surfactant to guide the 12 edges of the hexahedral ZIF into 12 planes, resulting in the transformation of the hexahedron into a polyhedron with 18 planes (Fig. S6). After pyrolysis, all the flat facets of the octadecahedron were depressed into concaves [Figs. 2(a), 2(c), and S3]. To prove the effect of GQDs, g-C3N4 nanosheets were introduced in the control experiment. The pyrolyzed morphology were shown in Figs. 2(f) and S7–S9. From enlarged SEM of Fig. S9(b), after pyrolysis, the g-C3N4 nanosheet induced morphology is similar with the pyrolyzed ZIF-67, and no octadecahedron was formed. This is a direct proof of the structure-directing effect of GQDs.The structure of the octadecahedron was further observed with transmission electron microscopy (TEM). From the TEM images in Figs. 2(b), S10, and the EDS images in Fig. 2(g), the inside of the octadecahedron is hollow. However, as shown in Figs. 2(e) and 2(h), the pyrolyzed ZIF-67 does not have the hollow structure. The formation of the hollow structure can be explained by the decomposition of g-C3N4, which is proved by the TGA analysis (Fig. S11). During the pyrolysis, g-C3N4 was totally decomposed to make the inside of the octadecahedron hollow, and the ammonia produced by its decomposition reacted with the shell to make it porous. The hollow and porous structure plays an important role in exposing active sites and transferring reactants and products.1818. L. Gao, X. Gao, P. Jiang, C. Zhang, H. Guo, Y. Cheng, L. Gao, X. Gao, P. Jiang, C. Zhang, H. Guo, and Y. Cheng, Small 18, 2105892 (2022). https://doi.org/10.1002/smll.202105892 As shown in the bottom inset of Fig. 2(b), the fringe spacing of 0.205 nm was indexed to the lattice fringes of Co (111). The corresponding SAED image [Fig. 2(b) inset, top] also shows the clear diffraction ring of Co (111). Comparing the uniform distribution of the Co elements in CNQD/CoNBs in Fig. 2(g), obvious aggregation is found for ZIF-67 in Figs. 2(e) and 2(h). This result indicates that GQDs can effectively prevent the agglomeration of Co nanoparticles caused by the collapse of the ZIF framework structure. Therefore, it is possible to obtain more single Co atoms anchored on the N-doped porous carbon in the form of Co–Nx active sites by pyrolysis, which is beneficial to exhibit an exceptional catalytic activity.The Fourier-transform infrared spectroscopy (FT-IR) measurements (Fig. S24) suggest the formation of abundant functional groups in CNQD/CoNBs precursor as supported by the presence of hydroxyl (O–H), C=N, C–C and C–H stretch bands, and NH bend-stretch. These functional groups facilitate the combination of Co ions with the quantum dots, which can prevent the agglomeration of Co ions. Different amounts of CTAB as surfactants enable MOF particles to grow into different crystal morphologies.1919. Y. Pan, D. Heryadi, F. Zhou, L. Zhao, G. Lestari, H. Su, and Z. Lai, CrystEngComm 13, 6937 (2011). https://doi.org/10.1039/c1ce05780d Materials with hexahedral structures as shown in Fig. S4 were prepared in the absence of quantum dots. However, the abundant functional groups of g-C3N4 quantum dots in solution can weaken the effect of CTAB on the crystal surface to a certain extent, so that six crystal planes (100) of MOF particles can grow until the 12 truncated edges (110) were formed (Fig. S6) to obtain the octadecahedron.2020. A. Schejn, L. Balan, V. Falk, L. Aranda, G. Medjahdi, and R. Schneider, CrystEngComm 16, 4493 (2014). https://doi.org/10.1039/C3CE42485EX-ray diffraction (XRD) characterization in Fig. 3(a) shows that only two obvious diffraction peaks at 44.2° and 51.3° were observed for the three samples (ZIF-67, CNN/CoNBs, and CNQD/CoNBs), corresponding to the (111) and (200) planes of the face-centered cubic phase metal Co, further indicating that the addition of GQDs had no effect on the formation of ZIF-67. Through utilizing Lorentzian functions, the first-order Raman spectrum is further fitted into five bands to further comprehend the structural evolution in the process of edge enrichment. In general, D-band and G-band represent disorder structure and the degree of graphitization in graphite materials, respectively. The D1 band is attributed to the defective crystallinity, and the D2 band is associated with the transformation of sp2 to sp33. M. Wu, G. Zhang, N. Chen, Y. Hu, T. Regier, D. Rawach, and S. Sun, ACS Energy Lett. 6, 1153 (2021). https://doi.org/10.1021/acsenergylett.1c00037 hybridization in carbon lattice. The ratio of ID/ID′ (Table S1) can distinguish the nature of defects including boundary, vacancy, and sp3 carbon. According to the results of the Raman spectra [Fig. 3(c) and Table S1], ID/IG of CNQD/CoNBs is the highest (1.36). This may be owing to the g-C3N4 decomposing inside the ZIF-67 to produce ammonia during the process of thermal annealing, and more defects and disorder are formed on the inside and outside surface of the carbon material. The specific surface area and pore size distribution of the catalyst were studied using N2 adsorption/desorption isotherm. CNQD/CoNBs have the largest specific surface area of 296.6 m2 g−1 [Figs. 3(b), S12a, and Table S2]. The inset images in Fig. 3(b) shows that the micropores with a pore size of 0.55 nm are dominant, which are much more abundant in CNQD/CoNBs than in other samples [Fig. S12(b)]. The abundant micropores may come from the corrosion of gas during the thermal decomposition of g-C3N4, which provides a larger surface for the active center.The XPS measurement was performed to determine the chemical composition. According to the survey spectra in Figs. S13 and S14, all the four samples contained the same elements. The total N content of CNQD/CoNBs is 10.1 at. %, which is the highest in all the samples as listed in Table S3. Figure 3(d) shows that the N1s peak is deconvoluted into three different peaks, which are, respectively, attributed to graphitized nitrogen, pyrrole nitrogen, and pyridine nitrogen. Figure 3(e) is the high-resolution XPS spectra of Co 2p peak. The peak at 778.4 eV is attributed to the metal Co, while the peak at 780.1 eV corresponds to Co–Nx, and the peak position of Co2+ is at 781.5 eV. In addition, compared with ZIF-67 (Fig. S14 and Table S4), the doping of GQDs promotes the formation of Co–Nx. The N 1s peak located at 399.7 eV is ascribed to the Co–Nx bond, which again confirms that Co–Nx is embedded in the N-doped carbon matrix. According to recent studies, both the charge transfer caused by N doping and the defect structure caused by N removal in carbon materials are beneficial to the catalytic performance in OER and ORR.21–2421. X. Li, Z. Su, Z. Zhao, Q. Cai, Y. Li, and J. Zhao, J. Colloid Interface Sci. 607, 1005 (2022). https://doi.org/10.1016/j.jcis.2021.09.04522. D. Guo, R. Shibuya, C. Akiba, S. Saji, T. Kondo, and J. Nakamura, Science 351, 361 (2016). https://doi.org/10.1126/science.aad083223. H. Bin Yang, J. Miao, S. F. Hung, J. Chen, H. B. Tao, X. Wang, L. Zhang, R. Chen, J. Gao, H. M. Chen, L. Dai, and B. Liu, Sci. Adv. 2, e1501122 (2016). https://doi.org/10.1126/sciadv.150112224. Y. Jia, L. Zhang, L. Zhuang, H. Liu, X. Yan, X. Wang, J. Liu, J. Wang, Y. Zheng, Z. Xiao, E. Taran, J. Chen, D. Yang, Z. Zhu, S. Wang, L. Dai, and X. Yao, Nat. Catal. 28(2), 688 (2019). https://doi.org/10.1038/s41929-019-0297-4 Moreover, it is reported that Co–Nx is an active site of ORR.2525. Y. Du, F. X. Ma, C. Y. Xu, J. Yu, D. Li, Y. Feng, and L. Zhen, Nano Energy 61, 533 (2019). https://doi.org/10.1016/j.nanoen.2019.05.001 The formation of dispersed Co active site (Co2+ and metal Co) is significantly improved after the GQDs were doped (Table S5), which may be oxidized to CoOOH during the reaction and become the main active sites for OER.2626. B. Zhang, X. Zheng, O. Voznyy, R. Comin, M. Bajdich, M. García-Melchor, L. Han, J. Xu, M. Liu, L. Zheng, F. P. G. De Arquer, C. T. Dinh, F. Fan, M. Yuan, E. Yassitepe, N. Chen, T. Regier, P. Liu, Y. Li, P. De Luna, A. Janmohamed, H. L. Xin, H. Yang, A. Vojvodic, and E. H. Sargent, Science 352, 333 (2016). https://doi.org/10.1126/science.aaf1525Synchrotron-based x-ray absorption spectroscopy (XAS) is conducted to study the differences of the catalysts. Figure 3(f) shows the normalized x-ray absorption near edge structure (XANES) spectra of Co K-edge, which help to identify the valence states of Co species. The curve of CNQD/CoNBs is basically the same as that of the metal Co reference sample, and the slight forward deviation may be attributed to the increase in Co–N bond content. The corresponding FT-EXAFS in R space [Fig. 3(g)] demonstrates the radial distribution function (RDF) of central atoms. The Co–Co peaks, Co–O peaks, and Co–N peaks were observed obviously in the reference materials Co foil, Co oxide, and Co phthalocyanine, while the peaks at R = 1.91 and 2.49 Å in CNQD/CoNBs are Co–N and Co–Co backscattering, respectively. The fitting result of CNQD/CoNBs EXFAS spectrum [Fig. 3(h) and Table S6] revealed that Co atom was coordinated with half N atoms and nine Co atom. Compared with the coordination structure of 12 Co atoms in Co foil, the coordination number of Co–Co in CNQD/CoNBs decreases to 9.39, indicating abundant vacancies and defects, which may be caused by the better maintenance of the framework structure of the material and the reduction of the content of Co nanoparticles. To confirm the adjacent scattering atomic species properties of the metal atoms, wavelet transform (WT) is further conducted to separate backscattering atoms owing to its powerful resolution in both R-space and K-space. As shown in Figs. 3(i)–3(k), the WT plot of CNQD/CoNBs show the maximum intensity at 3.1 Å−1 in K-space, which is higher than the Co–N (2.3 Å−1) in CoPC and slightly lower than Co–Co (3.3 Å−1) in Co foil. This again proves the two-bonding state of Co–Co and Co–N in CNQD/CoNBs.To understand fundamental aspects of O2 adsorption (reduction) and desorption (oxidation) at the Co nanoparticles-CoN4 composite site with four different geometric configurations, the relevant models were conceptually established in the theoretical study to represent the distinctly active sites in the catalysts, as shown in Figs. 4(a)–4(d). From the calculated density of states (DOS) [Fig. 4(e)], one can see that the materials manifest metallic characters with a zero-bandgap, and the d band center of the surface-side structure (Ed = −1.776 V) and plane-center structure (Ed = −1.767 V) can be seen to fall in the intermediate between those of surface-center (Ed = −1.799 V) and plane-side structure (Ed = −1.739 V). As for the bifunctional catalyst, the material should have a more moderate band energy, so it can have a more appropriate energy barrier in the opposite process of O2 adsorption and desorption for better catalytic activity in ORR and OER. Compared with the base plane structure of Co nanoparticles located in the center of CoN4, the curved substrate structure of Co nanoparticles located at the side of CoN4 is obviously more likely to form from the perspective of probability, so the content may be higher. In Fig. 4(f), the surface side structure also has a relatively moderate adsorption energy of O2 (2.3 eV), compared with 2.83 eV of the plane center structure, 2.00 eV of the plane side structure and 2.48 eV of the surface center structure. At the same time, materials with curved structure will have more active sites exposed due to their geometric configuration and the integrity of their own frame structure, so they have better catalytic activity.To prove the effect of adding GQDs on electrochemical performance of materials, a series of samples was first evaluated for OER in an alkaline medium. CNQD/CoNBs exhibit the lowest overpotential [1.36 [email protected] mA/cm2, Fig. S15(a)], suggesting that the addition of GQDs can effectively promote the OER performance. CNQD/CoNBs show the smallest Tafel slope of 140 mV/dec [Fig. S15(b)], suggesting more rapid OER kinetic process. As shown in Fig. S15(c), the OER stability of CNQD/CoNBs is superior than commercial RuO2. The rates of O2 production were monitored by the water drainage method to estimate the Faradaic efficiency (Fig. S16), and the calculated value was 96.5%. In general, CNQD/CoNBs exhibit better OER catalytic activity than most reported electrocatalysts [Fig. S25, Table S7 and S8).

The ORR activity of the samples was further explored. The polarization curves in Fig. S15(d) shows that CNQD/CoNBs have superior onset potential (Eonset = 0.96 V vs RHE) and half-wave potential (E1/2 = 0.86 V vs RHE) than commercial Pt/C samples. In addition, CNQD/CoNBs have the lowest Tafel slope [Fig. S15(e), 51 mV/dec], indicating the fastest reaction kinetics. Compared with Pt/C, CNQD/CoNBs show a better stability in Fig. S15(f). The post-reaction catalysts were characterized. The results of XRD and TEM shows that the structure and crystal phase of the samples are basically the same before and after the reaction (Figs. S17 and S18). XPS results of the materials after OER show the disappearance of metal Co peaks and the formation of Co in high valence state, which may be amorphous oxide production (Fig. S19).

Correspondingly, Nyquist plots shows the smallest arc radius (Fig. S20), indicating that CNQD/CoNBs have the lowest charge transfer resistance. Furthermore, CNQD/CoNBs also exhibit the largest electrochemical double-layer capacitance (Cdl = 70.9 mF/cm2, Fig. S21 and S22). The electrochemical active surface area (ECSA) is calculated from the Cdl after dividing it by the specific area capacitance (Cs) of 0.04 mF cm−2.2727. H. S. Jadhav, A. Roy, B. Z. Desalegan, and J. G. Seo, Sustainable Energy Fuels 4, 312 (2019). https://doi.org/10.1039/c9se00700h As shown in Fig. S23, the specific activity trend showed that the intrinsic activity of CNQD/CoNBs is superior to other samples.The overall electrocatalytic performances are assessed through the potential difference between Ej10 of OER and E1/2 of ORR in Fig. 5(a). In comparison with reported bifunctional ORR and OER catalysts28,2928. L. Liu, Y. Wang, F. Yan, C. Zhu, B. Geng, Y. Chen, and S. lei Chou, Small Methods 4, 1–9 (2020). https://doi.org/10.1002/smtd.20190057129. C. Zhu, Z. Yin, W. Lai, Y. Sun, L. Liu, X. Zhang, Y. Chen, and S. L. Chou, Adv. Energy Mater. 8, 1–12 (2018). https://doi.org/10.1002/aenm.201802327 CNQD/CoNBs have the smallest ΔE of 0.50 V [Fig. 5(b), Tables S6 and S7], indicating that it has good reversible oxygen reaction activity and could be a good candidate as air electrode for zinc–air batteries.So far, many cobalt-based catalysts for zinc–air batteries have been reported.30–3230. L. Gao, M. Zhang, H. Zhang, and Z. Zhang, J. Power Sources 450, 227577 (2020). https://doi.org/10.1016/j.jpowsour.2019.22757731. J. Tan, T. Thomas, J. Liu, L. Yang, L. Pan, R. Cao, H. Shen, J. Wang, J. Liu, and M. Yang, Chem. Eng. J. 395, 125151 (2020). https://doi.org/10.1016/j.cej.2020.12515132. C. Guan, A. Sumboja, H. Wu, W. Ren, X. Liu, H. Zhang, Z. Liu, C. Cheng, S. J. Pennycook, J. Wang, C. Guan, H. J. Wu, X. M. Liu, H. Zhang, S. J. Pennycook, J. Wang, A. Sumboja, Z. L. Liu, W. N. Ren, and C. W. Cheng, Adv. Mater. 29, 1704117 (2017). https://doi.org/10.1002/adma.201704117 The zinc–air battery was assembled with CNQD/CoNBs as the catalyst [Fig. 5(c)]. The power density of the battery drawn in Fig. 5(e) can reach 210 mW/cm2 and the open circuit voltage is basically stable at 1.49 V [Fig. 5(d)]. In addition, after 80 h cycle test, the voltage gap of CNQD/CoNBs-based batteries did not change significantly, indicating its good stability as shown in Fig. 5(f). The above-mentioned analysis confirmed the excellent performance of CNQD/CoNBs as the OER/ORR electrocatalysts rechargeable zinc–air devices.

In summary, GQDs are doped during the preparation of ZIF-67 to form CNQD/CoNBs octadecahedron with 18 concaved facets. The GQDs has a significant effect on this structural transition from hexahedron to octadecahedron. The formation of the hollow structure is caused by the decomposition of internal g-C3N4 during the pyrolysis, which may contribute to the uniform dispersion of Co nanoparticles. CNQD/CoNBs exhibit excellent oxygen electrocatalytic activity and durability in alkaline solution, with an OER potential of only 1.36 V at 10 mA/cm2 and the half-wave potential of ORR at 0.86 V. Based on CNQD/CoNBs as the air electrode catalyst, the rechargeable zinc–air battery showed better performance than batteries with precious metal catalysts in power density and cycle stability. Therefore, this work provides a method for constructing multifunctional and durable carbon-based materials for rechargeable zinc–air battery.

See the supplementary material for experimental details, additional SEM and TEM images, XRD patterns, TGA analysis, XPS spectra, electrochemical results, and FT-IR spectra.

This work was supported by the National Natural Science Foundation of China (Nos. 22075102, 22005120, 21576301, 51973244, and 22209056); the Fundamental Research Funds for the Central Universities (Nos. 21620329 and 21621007); the Postdoctoral Research Foundation of China (No. 2020M673071); the Special Funds for the Cultivation of Guangdong College Students' Scientific and Technological Innovation (No. pdjh2022b0065); the Guangdong Basic and Applied Basic Research Foundation (Nos. 2022A1515010523 and 2023A1515010270); the Science and Technology Planning Project of Guangzhou (Nos. 201605030008 and 202201010360), and the National Innovation and Entrepreneurship Training Program for Undergraduate (No. 202110559044).

Conflict of Interest

The authors have no conflicts to disclose.

Author Contributions

Zhiyi Chen and Yanting Ye contributed equally to this work.

Zhiyi Chen: Data curation (equal); Formal analysis (equal); Investigation (equal); Software (equal); Writing – original draft (lead). Xiang Yu: Data curation (equal); Formal analysis (supporting); Funding acquisition (equal); Resources (equal); Validation (equal). Hui Meng: Funding acquisition (equal); Resources (equal); Validation (equal); Writing – review & editing (equal). Yanting Ye: Formal analysis (equal); Software (equal); Writing – original draft (supporting). Xiulan Li: Data curation (supporting); Software (equal). Longfu Li: Data curation (supporting); Formal analysis (supporting). Muzi Yang: Formal analysis (supporting); Resources (supporting). Jian Chen: Formal analysis (supporting); Resources (supporting). Fangyan Xie: Data curation (supporting); Formal analysis (supporting); Resources (supporting). Yanshuo Jin: Supervision (supporting); Validation (supporting). Nan Wang: Funding acquisition (equal); Investigation (equal); Resources (equal); Supervision (equal); Writing – original draft (supporting); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

REFERENCES

1. J. Wang, F. Xu, H. Jin, Y. Chen, Y. Wang, J. Wang, F. Xu, H. Y. Jin, Y. Q. Chen, and Y. Wang, Adv. Mater. 29, 1605838 (2017). https://doi.org/10.1002/adma.201605838, Google ScholarCrossref, ISI2. M. Wang, H. Zheng, Y. Dong, Z. Jin, S. Liao, W. Gong, X. Li, Z. Li, M. Xia, and J. Wang, Appl. Phys. Lett. 120, 261903 (2022). https://doi.org/10.1063/5.0088786, Google ScholarScitation3. M. Wu, G. Zhang, N. Chen, Y. Hu, T. Regier, D. Rawach, and S. Sun, ACS Energy Lett. 6, 1153 (2021). https://doi.org/10.1021/acsenergylett.1c00037, Google ScholarCrossref4. A. Liu, X. Liang, X. Ren, W. Guan, and T. Ma, Electrochem. Energy Rev. 51, 112 (2021). https://doi.org/10.1007/S41918-021-00110-W, Google ScholarCrossref5. Z. Zhang, B. Xi, X. Ma, W. Chen, J. Feng, and S. Xiong, SusMat 2, 114 (2022). https://doi.org/10.1002/sus2.53, Google ScholarCrossref6. H. Luo, B. Liu, Z. Yang, Y. Wan, and C. Zhong, Electrochem. Energy Rev. 51 5, 187 (2021). https://doi.org/10.1007/S41918-021-00107-5, Google ScholarCrossref7. X. Zou and Y. Zhang, Chem. Soc. Rev. 44, 5148 (2015). https://doi.org/10.1039/C4CS00448E, Google ScholarCrossref8. J. Song, S. Qiu, F. Hu, Y. Ding, S. Han, L. Li, H. Y. Chen, X. Han, C. Sun, and S. Peng, Adv. Funct. Mater. 31, 2100618 (2021). https://doi.org/10.1002/adfm.202100618, Google ScholarCrossref9. L. Li, B. Wang, G. Zhang, G. Yang, T. Yang, S. Yang, and S. Yang, Adv. Energy Mater. 10, 2001600 (2020). https://doi.org/10.1002/aenm.202001600, Google ScholarCrossref10. J. Fu, R. Liang, G. Liu, A. Yu, Z. Bai, L. Yang, Z. Chen, J. Fu, R. Liang, G. Liu, A. Yu, Z. Chen, Z. Bai, and L. Yang, Adv. Mater. 31, 1805230 (2019). https://doi.org/10.1002/adma.201805230, Google ScholarCrossref11. Z. X. Cai, Z. L. Wang, Y. J. Xia, H. Lim, W. Zhou, A. Taniguchi, M. Ohtani, K. Kobiro, T. Fujita, and Y. Yamauchi, Angew. Chem., Int. Ed. 60, 4747 (2021). https://doi.org/10.1002/anie.202010618, Google ScholarCrossref12. D. Fairen-Jimenez, S. A. Moggach, M. T. Wharmby, P. A. Wright, S. Parsons, and T. Düren, J. Am. Chem. Soc. 133, 8900 (2011). https://doi.org/10.1021/ja202154j, Google ScholarCrossref, ISI13. J. Lee, O. K. Farha, J. Roberts, K. A. Scheidt, S. T. Nguyen, and J. T. Hupp, Chem. Soc. Rev. 38, 1450 (2009). https://doi.org/10.1039/b807080f, Google ScholarCrossref, ISI14. I. Saana Amiinu, Z. Pu, X. Liu, K. Asare Owusu, H. Gabriela Rivera Monestel, F. Ofori Boakye, H. Zhang, S. Mu, I. S. Amiinu, Z. Pu, X. Liu, H. R. Monestel, F. O. Boakye, H. Zhang, S. Mu, and K. A. Owusu, Adv. Funct. Mater. 27, 1702300 (2017). https://doi.org/10.1002/adfm.201702300, Google ScholarCrossref15. I. Saana Amiinu, X. Liu, Z. Pu, W. Li, Q. Li, J. Zhang, H. Tang, H. Zhang, S. Mu, I. S. Amiinu, X. Liu, Z. Pu, W. Li, J. Zhang, H. Tang, H. Zhang, S. Mu, and Q. Li, Adv. Funct. Mater. 28, 1704638 (2018). https://doi.org/10.1002/adfm.201704638, Google ScholarCrossref16. W. Xiang, J. Li, J. Ma, Z. Sheng, H. Lu, and S. Yang, New J. Chem. 45, 22787–22797 (2021). https://doi.org/10.1039/D1NJ03900H, Google ScholarCrossref17. X. Wan, X. Liu, Y. Li, R. Yu, L. Zheng, W. Yan, H. Wang, M. Xu, and J. Shui, Nat. Catal. 23(2), 259 (2019). https://doi.org/10.1038/s41929-019-0237-3, Google ScholarCrossref18. L. Gao, X. Gao, P. Jiang, C. Zhang, H. Guo, Y. Cheng, L. Gao, X. Gao, P. Jiang, C. Zhang, H. Guo, and Y. Cheng, Small 18, 2105892 (2022). https://doi.org/10.1002/smll.202105892, Google ScholarCrossref19. Y. Pan, D. Heryadi, F. Zhou, L. Zhao, G. Lestari, H. Su, and Z. Lai, CrystEngComm 13, 6937 (2011). https://doi.org/10.1039/c1ce05780d, Google ScholarCrossref20. A. Schejn, L. Balan, V. Falk, L. Aranda, G. Medjahdi, and R. Schneider, CrystEngComm 16, 4493 (2014). https://doi.org/10.1039/C3CE42485E, Google ScholarCrossref21. X. Li, Z. Su, Z. Zhao, Q. Cai, Y. Li, and J. Zhao, J. Colloid Interface Sci. 607, 1005 (2022). https://doi.org/10.1016/j.jcis.2021.09.045, Google ScholarCrossref

留言 (0)

沒有登入
gif