A three-dimensional flower-like Cu-based composite and its low-temperature calcination derivatives for efficient oxygen evolution reaction

Yifan LIU Zhan ZHANG Rongmei ZHU Ziming QIU Huan PANG

Citation:  Yifan LIU, Zhan ZHANG, Rongmei ZHU, Ziming QIU, Huan PANG. A three-dimensional flower-like Cu-based composite and its low-temperature calcination derivatives for efficient oxygen evolution reaction[J]. Chinese Journal of Inorganic Chemistry, 2024, 40(5): 979-990. doi: 10.11862/CJIC.20240008 shu

三维花状铜基复合物及其低温煅烧衍生物用于高效析氧反应

    通讯作者: 张沾, zhanzhang@yzu.edu.cn
    庞欢, panghuan@yzu.edu.cn
  • 基金项目:

    国家自然科学基金 U1904215

    江苏省省自然科学基金 BK20200044

    江苏省双创计划 (2020)30974

    扬州市绿扬金凤基金 YZLYJF2020PHD110

摘要: 采用一种简单易行的方法来制备铜基电化学催化剂, 该催化剂用于析氧反应。利用沸石咪唑酯骨架-67(ZIF-67)为前驱体, 通过铜离子刻蚀合成新型铜基复合物(命名为Cu-NF)。它具有特殊的三维花状形貌和中孔结构, 其形貌受铜离子与ZIF-67的质量比的影响。随后, Cu-NF经过低温煅烧处理, 得到了不同温度下的衍生物, 而煅烧过程并未改变其原始形貌。煅烧温度决定了活性物种的组成及含量, 并增强了材料的多孔性。其中, 300 ℃煅烧得到的产物Cu-NF-300具有最好的析氧反应性能: 在1.0 mol·L-1 KOH溶液中, 过电位低达347 mV, 塔菲尔斜率为93 mV·dec-1。该材料电化学性能的提高归因于其三维结构以及低温煅烧所带来的活化作用。

English

  • Electrocatalytic water splitting for hydrogen is one of the most promising strategies for the scalable production of clean hydrogen as sustainable energy[1-3]. However, the overall water splitting reaction is seriously restricted by the oxygen evolution reaction (OER) (4OH- = 2H2O + 4e- + O2) at the anode, ascribing to the sluggish four-electron transfer and the rigid O=O bond formation[4-6]. Currently, noble metal oxides, such as IrO2 and RuO2, are developed as the most outstanding commercial electrocatalysts for OER[7-9]. Nevertheless, to address the scarcity and high price of noble metals, transition-metal-based catalysts, such as Fe, Co, and Ni-based oxides or hydroxides, as alternatives, have been extensively explored in recent years[10-16]. In contrast, copper-based materials have attracted less attention, as a result of the inherent high-occupied anti-bonding state of the d orbital of Cu (3d104s1), which can hardly hybridize with the O2p orbital of oxygen intermediates during the absorption process of OER, despite of their cost-effective and highly-conductive merits[17-18].

    To enhance the efficiency and the structural stability of electrocatalysts for water oxidation, most researchers focus on developing new electrocatalysts or improving the performance of the already-known electrocatalysts[4, 18-19]. Effective strategies have been reported to optimize electronic structures, such as heteroatom doping, nanostructuring, coordination engineering, and metal valence variation[20-24], which determine the inherent activities of the active sites. In addition, reasonable engineering in microstructure endows the large surface areas between the catalysts and the electrolyte, which improves the exposure of the active sites, accelerating mass transport and electron transfer[25-26]. The widely reported strategies are constructing hierarchical morphology[27-28], designing surface defects[29-30], introducing porous structure[31-32], etc. Moreover, the configuration of three-dimensional (3D) architectures can be another choice for structural optimization due to the easy access of the electrolyte to the electrode surface, which shortens the diffusion path[15, 33-34]. 3D nanonetworks constructed by lower dimensional nanomaterials provide better conductivity, stability, and anti-aggregation properties in comparison to low-dimensional materials[35-36]. These advantages endow the 3D-structured nanomaterials with wide applications in energy conversion and storage[37-39]. However, it is difficult to design Cu-based 3D materials in safe, low-energy consuming, and high-yield synthetic strategies. Most of the reported synthetic processes are complicated or of low yield[40-41]. With reasonable utilization of these strategies above, the designed novel Cu-based materials are promising for low-cost efficient electrocatalysis of water[42].

    In this work, we reported a novel 3D Cu-based nanoflower-like composite with mesoporous structure (abbreviated as Cu-NF) using zeolitic imidazolate framework-67 (ZIF-67) as precursor through a facile excessive Cu2+ etching. The follow-up calcination of the Cu-NF composites at low temperatures under the N2 atmosphere obtained a series of Cu-NF derivatives with the reserved original shape (Scheme 1). It is worth mentioning that low-temperature calcination can be a simple way to enlarge the specific surface area and achieve improved porosity, by removing the coordinated moisture and partially decomposing the organic linkers[32, 43]. Specifically, the product obtained at 300 ℃ (abbreviated as Cu-NF-300) resulted in the best OER performance, with the lowest overpotential of 347 mV and the smallest Tafel slope of 93 mV·dec-1 at 10 mA·cm-2. The 3D superstructure showed advantages in facilitating mass transport and charge transfer. Moreover, the newly formed active species and the improved porosity caused by low-temperature treatment contribute to enhanced OER performance. This work provides a gentle and facile approach to preparing Cu-based materials for electrochemical applications.

    Scheme 1

    Scheme 1.  Synthetic approach of the Cu-NF composites and their derivatives

    Materials contained cobalt nitrate hexahydrate (Co(NO3)2·6H2O, 99.99%, Aladdin) and 2-methylimidazole (C4H6N2, 98%, Energy Chemical Co., Ltd.). Besides, copper(Ⅱ) nitrate trihydrate (Cu(NO3)2·3H2O, 99%), potassium hydroxide (KOH, 95%), ethanol (CH3CH2OH, AR) and methanol (CH3OH, AR) were purchased from Sinopharm Chemical Reagent. Nafion solution (5%) was from Sigma-Aldrich. All the chemicals were used without further purification. The water used in all experiments was prepared by passing through an ultra-pure purification system (Aqua Solutions).

    The ZIF-67 was synthesized according to our previous work[44]. Co(NO3)2·6H2O (3.64 g) and 2-methylimidazole (2-MIm) (4.106 7 g) were dissolved in 100 mL methanol, respectively. Then the two solutions were mixed and sonicated in an ultrasonic cleaner for 15 min. Then the purple product was separated by centrifugation and washed for 2-3 times with methanol. After that, the product was finally dried at 25 ℃ for 6 h.

    0.1 g of ZIF-67 and 0.2 g of Cu(NO3)2·3H2O were dispersed in 50 and 12.5 mL ethanol by ultrasonication, respectively. The two solutions were mixed and stirred for 12 h. Then the product was separated by centrifugation and washed with methanol. After that, the product was finally dried at 25 ℃.

    The procedures of synthesis of CuCo-NS and Cu-NP were identical to that of Cu-NF, except that the mass ratio of Cu(NO3)2·3H2O to ZIF-67 was 1∶1 or 3∶1.

    The Cu-NF composites were heat treated at different temperatures (200, 250, 300, 350, and 400 ℃) under an N2 atmosphere with a heating rate of 2 ℃·min-1 and retained for 2 h. The samples were named Cu-NP-T (T ℃ was the calcination temperature).

    The morphological characteristics of the sample were observed using a field emission scanning electron microscope (SEM) at an acceleration voltage of 5.0 kV (Zeiss_Supra55). The transmission electron microscopy (TEM) images, the selected area electron diffraction (SAED) images, and the energy dispersive spectroscopy (EDS) mapping images were captured on a Tecnai G2 F30 transmission electron microscope (300 kV). The X-ray diffraction (XRD) patterns were examined on a Bruker D8 Advanced X-ray diffractometer (Cu radiation: λ=0.154 18 nm, 40 kV, 40 mA, 2θ=5°-80°). The Fourier transform infrared spectrometer (FTIR) measurement was investigated on a Tensor 27. The inductively coupled plasma optical emission spectroscopy (ICP-OES) measurements were performed using a simultaneous ICP spectrometer (Optima 7300 DV, Perkin Elmer) equipped with a solid-state detector. The X-ray photoelectron spectroscopy (XPS) measurement was carried out using an axis ultra X-ray photoelectron spectrometer (Kratos Analytical Ltd., UK) equipped with Al source (=1 486.6 eV). The thermal gravimetric analysis (TGA) was carried out with a Pyris 1 TGA (Perkin Elmer, America) scientific instrument and heated from room temperature to 800 ℃ with a ramp rate of 5 ℃·min-1 in nitrogen gas flow.

    All the electrochemical measurements of OER were performed in a three-electrode system at CHI760e Electrochemical Station (CH Instruments, Shanghai, China). Electrochemical measurements were carried out using Hg/HgO electrode as reference electrode, graphite carbon rod as counter electrode, and glassy carbon electrode (GCE) (diameter: 3.0 mm) as the working electrode in 1.0 mol·L-1 KOH electrolyte. 5 mg of the samples (Cu-NF-T) was dispersed in 1 mL ethanol and ultrasonicated for 10 min. Then, 100 μL of 5% Nafion solution was added to the prepared solution, followed by sonicating for another 10 min. Then, 5 μL of the above suspension was taken with a pipette and placed dropwise onto the dried GCE surface. The GCE was dried at room temperature and the catalyst loading was about 0.3 mg·cm-2.

    The synthesis of the Cu-NF composites was achieved using ZIF-67 as a precursor through a facile Cu2+ etching strategy, as shown in Scheme 1. ZIF-67 particles were first synthesized according to the previous report[44]. The as-prepared ZIF-67 showed a typical dodecahedral morphology with an average size of 1 mm and smooth surface, characterized by SEM (Fig.S1, Supporting information). Afterward, Cu(NO3)2 was employed as a chemical etching agent to reset the metal arrangement and reconstruct the metal-ligand framework. For such a synthesis, the mass ratio of Cu2+ to ZIF-67 was found highly critical to control the morphology of the product. The morphology and the structure of the attained materials from different mass ratios were characterized by SEM (Fig. 1a-1c), as well as the morphology variations with the increase of reaction time (Fig.S2-S4). As shown in Fig.S2b, a moderate Cu2+ to ZIF-67 mass ratio (1∶1) caused insufficient etching of ZIF-67 at first, leading to the formation of nanosheets on the surface with the basic ZIF-67 core intact. With time passing, the nanosheets covered the core gradually (Fig.S2). As displayed in Fig. 1a, the product obtained in 12 h was composed of nanosheets (CuCo-NS). Since the mass ratio increased to 2∶1, excessive Cu2+ contributed to the morphological deformation of the dodecahedral ZIF-67 (1 h), followed by its disappearance and the subsequent appearance of the 3D flower-like Cu-based composites composed of thin nanosheets (6 h) (Fig.S3). The flowers became bigger and the petals were denser (Fig.S3c). A high yield was obtained in 12 h, which was considered optimal for the subsequent studies (Cu-NF, Fig. 1b). Further increasing the mass ratio to 3∶1 (Fig.S4), the polyhedron collapsed more rapidly, reconstructed the piles of nanoplates (Cu-NP, Fig. 1c).

    Figure 1

    Figure 1.  SEM images of (a) CuCo-NS, (b) Cu-NF, and (c) Cu-NP composites; HRTEM images and SAED pattern (Inset) of (d-f) the Cu-NF composites; (g) HAADF-STEM image of the Cu-NF composites and their corresponding elemental mapping images

    Particularly, excessive Cu2+ etching (mCu2+mZIF-67=2∶1) reaction went through the decomposition of ZIF-67 and the reconstruction of a novel Cu-based material. These could be evidenced both by the color change from purple to turquoise and the phenomenon of the solid disappearance and reappearance (Fig.S5). During the process, the hydrolysis of Cu2+ caused the generation of H+, which etches ZIF-67 severely, followed by the collapse of the framework[45]. Subsequently, concentrated Cu2+ and 2-MIm reconstructed a new network by self-assembly, together with the hydrolysis, forming Cu-LDHs (Cu-based layered double hydroxides) and other compounds on the surface. Hence, an increased Cu2+ ratio (3∶1) caused a more rapid decomposition reaction. The solid disappeared within 15 min and reoccurred in 2 h. In contrast, when a moderate mass ratio of Cu2+ to ZIF-67 (1∶1) was applied, there was no significant color change in the solution and only the color of the product changed to grey-purple.

    In light of the Cu-NF composites, Fig. 1b displayed that as-fabricated Cu-NF composites were uniformly 3D flower structures with an average diameter of 2 mm, comprised of thin nanosheets. The high-resolution TEM (HRTEM) images and the SAED pattern in Fig. 1d-1f exhibited the polycrystalline feature of the Cu-NF composites. The high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image and elemental mapping images of the Cu-NF composites in Fig. 1g, verified the homogeneous distribution of Cu, C, N, and O elements throughout the entire network. Nevertheless, little Co was detected in the composites. Meanwhile, the metal contents were also quantified by ICP-OES and relevant data were listed in Table S1. The results indicated that Cu element accounted for 97.64% and Co only for 2.36%, reconfirming the composites based on Cu, which agreed with the results from elemental mapping images. In addition, the results from element analysis (CNH) informed that the contents of C, N, and H in the organic part were 26.73%, 15.51%, and 3.25%, respectively, which were precise the proportions of C, N, and H in 2-MIm, indicating the existence of 2-MIm in the Cu-NF composites. The contents of Cu and Co elements in CuCo-NS and Cu-NP were also examined by ICP-OES and listed in Table S1. The N2 adsorption-desorption isotherms of the Cu-NF composites (Fig.S6) analyzed that the specific surface area was 41 m2·g-1, which was calculated by the Barrett-Joyner-Halenda (BET) method. Accordingly, the pore-size distribution curve (Fig.S6) of the composites indicated its mesoporous feature with an average pore diameter of 25.70 nm.

    To further investigate the microstructure and composition of the Cu-NF composites, XRD was utilized to investigate the prepared materials (Fig. 2a and 2b). As displayed in Fig. 2a, after the addition of Cu2+ for 30 min, the XRD pattern of the as-prepared ZIF-67 stayed unchanged. Since the solid reoccurred in 6 h, the new groups of peaks centered at 12.6°, 25.8°, and 34.1° appeared, suggesting the generation of a novel composite. With the passage of the reaction time (12 h), these peaks became intensified. In comparison with the standard cards, shown in Fig. 2b, the dominant peaks at 12.6°, 25.8°, and 34.1° refer to (001), (004) and (121) facets of gerhardtite (PDF No.14-0687), the peaks at 13.3°, 35.1° referring to the (001) and (201) facets of copper hydroxide hydrate (PDF No.42-0637) and those at 12.1°, 24.5°, and 33.9° can be indexed to (003), (006) and (012) planes of hydrotalcite-like LDHs[14, 26, 46], which reconfirmed the polycrystalline feature of the Cu-NF composites by the results from HRTEM. In addition, the XRD patterns of CuCo-NS and Cu-NP were also collected (Fig.S7). The CuCo-NS pattern had not only all the characteristic diffraction peaks that ZIF-67 has but also the peaks indicating the LDH construction. These suggest the coexisted ZIF-67 and LDH structure, reconfirming the results from SEM. By contrast, the Cu-NP XRD pattern demonstrates the well-crystalline state of Cu-LDH.

    Figure 2

    Figure 2.  (a) XRD patterns of the samples from the reaction mixture taken at different times during the reaction; (b) XRD patterns, (c) FTIR spectrum, (d) full XPS spectrum and XPS high-resolution spectra of (e) C1s, (f) Cu2p, (g) O1s, and (h) N1s of the Cu-NF composites

    Moreover, FTIR spectroscopy was utilized to further demonstrate the functional groups that existed in the prepared Cu-NF composites. As shown in Fig. 2c, a broad peak centered at 3 461 cm-1 is associated with O—H stretching, while the sharp strong peaks at 1 384 and 1 420 cm-1 indicate the vibration and stretching modes of nitrate[47-48]. The absorption of nitrate to balance the charge is typical in LDH structure[48]. The peaks at 1 560 and 1 146 cm-1 are related to the C=N and C—N stretching vibration, verifying the presence of 2-MIm[49-50]. The absorption peaks below 750 cm-1 are the vibration modes of Cu—O[47-48].

    In addition, XPS was performed to study the valance states of chemical elements in the Cu-NF composites (Fig. 2d-2h). Fig. 2d reconfirmed the coexistence of Cu, C, N, and O elements. The peak of Co was barely visible, indicating little Co existed, matching the results aforementioned. In Fig. 2f, the high-resolution Cu2p spectrum with two dominant groups of peaks could be noticed. The main peaks at 934.8 and 954.7 eV are indexed to the 2p3/2 and 2p1/2, along with their satellite peaks at 942.5 and 962.6 eV, which are the characteristic of the Cu2+ species[51-52]. The peaks of O1s at 531.4 and 532.2 eV displayed in Fig. 2g are ascribed to M—O (M=metal) and hydroxyl groups[53], confirming the Cu—OH structure. The peaks of N1s (Fig. 2h) are assigned to three nitrogen configurations, including pyridinic N (400.1 eV), pyrrolic N (398.9 eV) and N in nitrite (407.0 eV)[54], and those of C1s (Fig. 2e) are referring to C—C and N—C[44, 46], verifying the existence of 2-MIm.

    To modulate the microstructure of the Cu-NF composites, low-temperature calcination was conducted. Depending on the results of TGA under the N2 atmosphere (Fig. 3a), the weight loss of the Cu-NF composites suffered several steps before collapse at 400 ℃, which involved the removal of the moisture and the organic ligands. Therefore, calcination temperature was rationally chosen between 200 and 400 ℃. The SEM images in Fig. 4a1-4e1 showed that the 3D flower-like outlines composed of thin nanosheets were maintained after calcined at different temperatures, except for Cu-NF-400. Fig. 4d1 and 4e1 displayed the thicker petals with smaller sizes (around 1 mm), indicating that the high-temperature treatment led to structure collapse and material agglomeration. Meanwhile, the TEM images exhibited distinct nanoparticles produced on the nanosheets after raising the temperature to 300 ℃ or above (Fig. 4c2-4e2), which could be the metallic Cu particles produced, being further confirmed in subsequent XRD results. The XRD patterns in Fig. 3b disclosed the significant changes in the chemical components and their contents among the prepared samples with the increase of the calcination temperature and Fig.S8-S12 detailed their chemical compositions. Additionally, the sharper peaks of Cu-NF-350 and Cu-NF-400 demonstrate their stronger crystallinity, as can be seen in Fig. 3b. Cu-NF-200 shows the less complicated component (Fig.S8) and lower crystallinity than Cu-NF, because of the removal of the water absorbed physically and coordinated in the network. Then, Cu2+ was reduced to Cu2+1O (Cu2O with copper excess defects), together with the removal of the coordinated water and NO3- at 250 ℃ (Fig.S9), and further reduced to metallic Cu at 400 ℃ (Fig.S12). Particularly, Cu-NF-300 was composed of crystalline nanoparticles of Cu2+1O and metallic Cu (Fig.S10), while Cu-NF-350 had similar constituents with higher content of Cu (Fig.S13). These results matched well with the TEM images in Fig. 4a2-4e2.

    Figure 3

    Figure 3.  (a) TGA curve of the Cu-NF composites; (b) XRD patterns, (c) FTIR spectra of the Cu-NF composites and their derivatives and (d) Raman spectra

    Figure 4

    Figure 4.  SEM images of (a1) Cu-NF-200, (b1) Cu-NF-250, (c1) Cu-NF-300, (d1) Cu-NF-350, (e1) Cu-NF-400; TEM images of (a2) Cu-NF-200, (b2) Cu-NF-250, (c2) Cu-NF-300, (d2) Cu-NF-350, (e2) Cu-NF-400

    The FTIR spectra of the prepared samples exhibit that the peak at 1 146 cm-1 relates to the C—N bond[55] became weaker and vanished after calcined at 350 and 400 ℃ (Fig. 3c), depicting the decomposition of the ligand from 350 ℃ and a certain degree of carbonization at 350 ℃ and above. Together with the results above, the existence of N in Cu-NF-300 demonstrates the composites to be N-doped Cu2+1O and Cu nanocomposites, which contributes to enhanced OER performance. Furthermore, the Raman spectra of all samples were recorded to identify the structure of the carbon framework (Fig. 3d). The represent D and G peaks at 1 350 and 1 580 cm-1 display the degree of the defect and disorder of the carbon in the materials, respectively[56]. All the materials possess different degrees of surface defects after heat treatment.

    The electrochemical properties of the prepared samples as electrocatalysts for OER were evaluated in 1 mol·L-1 KOH solution by using a standard three-electrode system. In the linear sweep voltammetry (LSV) curves of Fig. 5a, Cu-NF-300 showed the earliest current response at 1.577 V (vs RHE) to reach the current density of 10 mA·cm-2 among all the prepared samples, namely a low overpotential of 347 mV (Fig. 5b). The other catalysts Cu-NF-250 and Cu-NF-350 were just inferior to it, requiring overpotentials of 368 and 377 mV, respectively. Meanwhile, the current of Cu-NF-300 increased most sharply, suggesting the best OER activity of Cu-NF-300. Accordingly, at a current density of 20 mA·cm-2, the overpotential of Cu-NF-300 was the lowest as well (Fig. 5b). The Tafel slopes derived from the LSV curves were calculated to evaluate the kinetics of the OER performance for all the samples (Fig. 5c). It is well known that a lower Tafel slope means a slower increase of the overpotential with the current density, demonstrating superior OER kinetics of the electrocatalyst. Cu-NF-300 showed the expected lowest Tafel slope with only 93 mV·dec-1, demonstrating the most favorable kinetics. The Tafel slope of other samples followed exactly the trend of their overpotentials. In addition, the electrochemical active surface areas (ECSAs) were estimated via the electrochemical double-layer capacitance (Cdl) using the cyclic voltammograms (CV) method in the non-Faradic region at different scan rates (Fig.S13). As shown in Fig. 5d, the Cdl values of Cu-NF-300, Cu-NF-350, and Cu-NF-400 were 6.71, 5.16, and 5.2 mF·cm-2, suggesting larger active surface areas than the others, with Cu-NF-300 as the largest one. The larger ECSA of Cu-NF-300 than Cu-NF-200 and Cu-NF-250 may be attributed to the improved porosity, resulting in a larger specific surface area. Nevertheless, compared to Cu-NF-350 and Cu-NF-400, the increased ECSA may be ascribed to more structure defects at the molecular level. Nyquist plots in Fig. 5e demonstrated the electrochemical impedance of the as-prepared electrocatalysts. The observation of sample Cu-NF-300 with a steeper line in low-frequency regions depicted the least charge transfer resistance among all these samples and matched well with the smallest Tafel slope.

    Figure 5

    Figure 5.  (a) LSV curves, (b) overpotentials at the specific current densities, (c) Tafel plots, (d) ECSAs based on CV curves (Fig.S13) at different scan rates, and (e) Nyquist plots of impedance spectroscopy analysis of the as-prepared electrocatalysts in 1.0 mol·L-1 KOH; (f) LSV curves of Cu-NF-300

    Inset in (e): fitted equivalent circuit model for the electrochemical impedance tests.

    The reasons for the better OER performance of Cu-NF-300 than Cu-NF-250, and Cu-NF-350, can be analyzed by combining the variation of microstructures and chemical compositions. With the temperature increasing from 250 to 300 ℃, metallic Cu formed in Cu-NF-300 acts as a new active center and improves the conductivity of the material, leading to better OER activity than Cu-NF-250. In the meantime, the increased heat treatment leads to the elimination of the coordinated ligands, which increases the porosity for better mass transfer. Further increasing the temperature to 350 ℃, the content ratio of metallic Cu to Cu2+1O increases in Cu-NF-350 compared to Cu-NF-300, leading the better conductivity. However, the stronger crystallinity of Cu-NF-350 results in lower surface defect, and metal agglomeration decreases the active centers, further reducing its OER activity. Overall, the chemical composition and the microstructure of Cu-NF-300 have a positive influence on the electrocatalytic process. Afterward, the durability property of the Cu-NF-300 was evaluated using a continuous cyclic voltammetry sweep for 1 000 cycles (Fig. 5f) and the polarization curve was found to be close to the initial one, only 29 mV decay of the overpotential at 10 mA·cm-2 after 1 000 cycles. In the long-time stability test via chronopotentiometry, a slight overpotential increase was observed after the test up to 15 h (Fig.S14), exhibiting relatively satisfactory stability of Cu-NF-300 in an alkaline medium.

    In summary, we prepared a novel Cu-based composite through a facile chemical etching strategy using ZIF-67 as a precursor, by controlling the mass ratio of Cu2+ to ZIF-67 to regulate the morphology. The as-obtained Cu-NF composites had a 3D flower-like architecture and mesoporous channels. After low-temperature calcination, a series of derivatives with the reserved original morphology have been obtained and applied as OER electrocatalysts. Particularly, Cu-NF-300 showed the best OER performance among all these samples, with a low overpotential of 347 mV and a small Tafel slope of 93 mV·dec-1 at 10 mA·cm-2. Except for the active species, the microstructure has a significant influence on the electrocatalyst for OER efficiency. The designed 3D superstructure and special microstructure obtained from calcination facilitate the transportation of the materials. Our results provide a new simple strategy for design Cu-based electrochemical materials.

    CRediT authorship contribution statement: Liu Yifan: Methodology, Conceptualization, Data curation, Formal analysis; Zhang Zhan: Conceptualization, Methodology, Writing-original draft, Formal analysis, Supervision, Funding acquisition; Zhu Rongmei: Methodology, Validation; Qiu Ziming: Formal analysis, Validation; Pang Huan: Funding acquisition, Writing-review & editing, Supervision.

    Declaration of Competing Interest: The 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.

    Acknowledgements: This work was supported by the National Natural Science Foundation of China (Grant No.U1904215), Natural Science Foundation of Jiangsu Province (Grant No.BK20200044), Jiangsu Shuangchuang Project (Grant No.(2020)30974) and Lvyangjinfeng Talent Program of Yangzhou (Grant No.YZLYJF2020PHD110).


    Supporting information is available at http://www.wjhxxb.cn
    1. [1]

      Chatenet M, Pollet B G, Dekel D R, Dionigi F, Deseure J, Millet P, Braatz R D, Bazant M Z, Eikerling M, Staffell I, Balcombe P, Shao-Horn Y, Schäfer H. Water electrolysis: From textbook knowledge to the latest scientific strategies and industrial developments[J]. Chem. Soc. Rev., 2022, 51(11):  4583-4762. doi: 10.1039/D0CS01079K

    2. [2]

      Liang Z B, Zhao R, Qiu T J, Zou R Q, Xu Q. Metal-organic framework-derived materials for electrochemical energy applications[J]. EnergyChem, 2019, 1(1):  100001. doi: 10.1016/j.enchem.2019.100001

    3. [3]

      Song J J, Wei C, Huang Z F, Liu C T, Zeng L, Wang X, Xu Z C J. A review on fundamentals for designing oxygen evolution electrocatalysts[J]. Chem. Soc. Rev., 2020, 49(7):  2196-2214. doi: 10.1039/C9CS00607A

    4. [4]

      Masa J, Andronescu C, Schuhmann W. Electrocatalysis as the nexus for sustainable renewable energy: The gordian knot of activity, stability, and selectivity[J]. Angew. Chem. Int. Ed., 2020, 59(36):  15298-15312. doi: 10.1002/anie.202007672

    5. [5]

      Zhang K X, Zou R Q. Advanced transition metal-based OER electrocatalysts: current status, opportunities, and challenges[J]. Small, 2021, 17(37):  2100129. doi: 10.1002/smll.202100129

    6. [6]

      Xu Y M, Fan K C, Zou Y, Fu H Q, Dong M Y, Dou Y H, Wang Y, Chen S, Yin H J, Al-Mamun M, Liu P R, Zhao H J. Rational design of metal oxide catalysts for electrocatalytic water splitting[J]. Nanoscale, 2021, 13(48):  20324-20353. doi: 10.1039/D1NR06285A

    7. [7]

      Hong C B, Li X F, Wei W B, Wu X T, Zhu Q L. Nano-engineering of Ru-based hierarchical porous nanoreactors for highly efficient pH-universal overall water splitting[J]. Appl. Catal. B-Environ., 2021, 294:  120230. doi: 10.1016/j.apcatb.2021.120230

    8. [8]

      Yao Q, Huang B L, Zhang N, Sun M Z, Shao Q, Huang X Q. Channel-rich RuCu nanosheets for pH-universal overall water splitting electrocatalysis[J]. Angew. Chem. Int. Ed., 2019, 58(39):  13983-13988. doi: 10.1002/anie.201908092

    9. [9]

      Sardar K, Ball S C, Sharman J D B, Thompsett D, Fisher J M, Smith R A P, Biswas P K, Lees M R, Kashtiban R J, Sloan J, Walton R I. Bismuth iridium oxide oxygen evolution catalyst from hydrothermal synthesis[J]. Chem. Mater., 2012, 24(21):  4192-4200. doi: 10.1021/cm302468b

    10. [10]

      Hao J, Wu K L, Lyu C J, Yang Y Q, Wu H J, Liu J J, Liu N Y, Lau W M, Zheng J L. Recent advances in interface engineering of Fe/Co/Ni-based heterostructure electrocatalysts for water splitting[J]. Mater. Horizons, 2023, 10(7):  2312-2342. doi: 10.1039/D3MH00366C

    11. [11]

      Wei Y, Zheng M B, Zhu W, Zhang Y, Hu W H, Pang H. Preparation of hierarchical hollow CoFe Prussian blue analogues and its heat-treatment derivatives for the electrocatalyst of oxygen evolution reaction[J]. J. Colloid Interface Sci., 2023, 631:  8-16. doi: 10.1016/j.jcis.2022.11.014

    12. [12]

      Hu L Y, Tian L L, Ding X, Wang X, Wang X S, Qin Y, Gu W L, Shi L, Zhu C Z. p-d hybridization in CoFe LDH nanoflowers for efficient oxygen evolution electrocatalysis[J]. Inorg. Chem. Front., 2022, 9(20):  5296-5304. doi: 10.1039/D2QI01688E

    13. [13]

      Zhang H, Meng G, Wei T R, Ding J Y, Liu Q, Luo J, Liu X J. Co doping promotes the alkaline overall seawater electrolysis performance over MnPSe3 nanosheets[J]. Chem. Commun., 2023, 59:  12144-12147. doi: 10.1039/D3CC03434H

    14. [14]

      Zhang X, Yan F, Ma X Z, Zhu C L, Wang Y, Xie Y, Chou S L, Huang Y J, Chen Y J. Regulation of morphology and electronic structure of FeCoNi layered double hydroxides for highly active and stable water oxidization catalysts[J]. Adv. Energy Mater., 2021, 11(48):  2102141. doi: 10.1002/aenm.202102141

    15. [15]

      Huang Y, Zhang S L, Lu X F, Wu Z P, Luan D Y, Lou X W. Trimetallic spinel NiCo2-xFexO4 nanoboxes for highly efficient electrocatalytic oxygen evolution[J]. Angew. Chem. Int. Ed., 2021, 60(21):  11841-11846. doi: 10.1002/anie.202103058

    16. [16]

      Huang W G, Liu Q F, Zhou Z W, Li Y S, Ling Y J, Wang Y, Tu Y C, Wang B B, Zhou X H, Deng D H, Yang B, Yang Y, Liu Z, Bao X H, Yang F. Tuning the activities of cuprous oxide nanostructures via the oxide-metal interaction[J]. Nat. Commun., 2020, 11(1):  2312. doi: 10.1038/s41467-020-15965-8

    17. [17]

      Shi H, Zhou Y T, Yao R Q, Wan W B, Ge X, Zhang W, Wen Z, Lang X Y, Zheng W T, Jiang Q. Spontaneously separated intermetallic Co3Mo from nanoporous copper as versatile electrocatalysts for highly efficient water splitting[J]. Nat. Commun., 2020, 11(1):  2940. doi: 10.1038/s41467-020-16769-6

    18. [18]

      盛子鸣, 陶友荣, 许路路, 杨鹏, 王文欣, 吴兴才. 简易电沉积法制备CoFe-P催化剂用于高效析氧反应[J]. 无机化学学报, 2023,39,(7): 1325-1337. SHENG Z M, TAO Y R, XU L L, YANG P, WANG W X, WU X C. CoFe-P catalyst prepared by a facile electrodeposition for high-efficient oxygen evolution reaction[J]. Chinese J. Inorg. Chem., 2023, 39(7):  1325-1337.

    19. [19]

      秦春玲, 陈爽, GomaaHassanien, ShenashenMohamed A., El-SaftySherif A., 刘倩, 安翠华, 刘熙俊, 邓齐波, 胡宁. 外加物理场调控二维材料的HER和OER性能[J]. 物理化学学报, 2024,40,(9): 2307059. QIN C L, CHEN S, GOMMA H, SHENASHEN M A, EI-SAFTY S A, LIU Q, AN C H, LIU X J, DENG Q B, HU N. Regulating HER and OER performances of 2D materials by the external physical fields[J]. Acta Phys.-Chim. Sin., 2024, 40(9):  2307059.

    20. [20]

      Zheng X B, Yang J R, Xu Z F, Wang Q S, Wu J B, Zhang E H, Dou S X, Sun W P, Wang D S, Li Y D. Ru-Co pair sites catalyst boosts the energetics for the oxygen evolution reaction[J]. Angew. Chem. Int. Ed., 2022, 61(32):  e202205946. doi: 10.1002/anie.202205946

    21. [21]

      Zhao G Q, Li P, Cheng N Y, Dou S X, Sun W P. An Ir/Ni(OH)2 heterostructured electrocatalyst for the oxygen evolution reaction: breaking the scaling relation, stabilizing iridium(Ⅴ), and beyond[J]. Adv. Mater., 2020, 32(24):  2000872. doi: 10.1002/adma.202000872

    22. [22]

      Li S, Chen B B, Wang Y, Ye M Y, van Aken P A, Cheng C, Thomas A. Oxygen-evolving catalytic atoms on metal carbides[J]. Nat. Mater., 2021, 20(9):  1240-1247. doi: 10.1038/s41563-021-01006-2

    23. [23]

      Tian X L, Zhao X, Su Y Q, Wang L J, Wang H M, Dang D, Chi B, Liu H F, Hensen E J M, Lou X W, Xia B Y. Engineering bunched Pt-Ni alloy nanocages for efficient oxygen reduction in practical fuel cells[J]. Science, 2019, 366(6467):  850-856. doi: 10.1126/science.aaw7493

    24. [24]

      Zhang Q, Lian K, Liu Q, Qi G C, Zhang S S, Luo J, Liu X J. High entropy alloy nanoparticles as efficient catalysts for alkaline overall seawater splitting and Zn-air batteries[J]. J. Colloid Interface Sci., 2023, 646:  844-854. doi: 10.1016/j.jcis.2023.05.074

    25. [25]

      Yu L, Deng D H, Bao X H. Chain mail for catalysts[J]. Angew. Chem. Int. Ed., 2020, 59(36):  15294-15297. doi: 10.1002/anie.202007604

    26. [26]

      Li M, Li Y B, Cu Q, Li Y, Li H Y, Li Z H, Li M, Liao H, Li G, Li G R, Wang X. Hollow and hierarchical CuCo-LDH nanocatalyst for boosting sulfur electrochemistry in Li-S batteries[J]. Mater. Adv., 2023, 4:  0032.

    27. [27]

      Li X R, Li Y P, Wang C L, Xue H G, Pang H, Xu Q. A 3D hierarchical electrocatalyst: Core-shell Cu@Cu(OH)2 nanorods/MOF octahedra supported on N-doped carbon for oxygen evolution reaction[J]. Nano Res., 2023, 16(5):  8012-8017. doi: 10.1007/s12274-023-5389-4

    28. [28]

      Lee J, Suh Y, Dubey P P, Barako M T, Won Y. Capillary wicking in hierarchically textured copper nanowire arrays[J]. ACS Appl. Mate. Interfaces, 2019, 11(1):  1546-1554. doi: 10.1021/acsami.8b14955

    29. [29]

      Yan D F, Xia C F, Zhang W J, Hu Q, He C X, Xia B Y, Wang S Y. Cation defect engineering of transition metal electrocatalysts for oxygen evolution reaction[J]. Adv. Energy Mater., 2022, 12(45):  2202317. doi: 10.1002/aenm.202202317

    30. [30]

      Fan F F, Huang Q L, Rani K K, Peng X L, Liu X T, Wang L M, Yang Z Y, Huang D J, Devasenathipathy R, Chen D H, Fan Y H, Chen W. Interface and doping engineering of Co-based electrocatalysts for enhanced oxygen reduction and evolution reactions[J]. Chem. Eng. J., 2023, 470:  144380. doi: 10.1016/j.cej.2023.144380

    31. [31]

      Nam D H, Bushuyev O S, Li J, De Luna P, Seifitokaldani A, Dinh C T, de Arquer F P G, Wang Y H, Liang Z Q, Proppe A H, Tan C S, Todorović P, Shekhah O, Gabardo C M, Jo J W, Choi J M, Choi M J, Baek S W, Kim J, Sinton D, Kelley S O, Eddaoudi M, Sargent E H. Metal-organic frameworks mediate Cu coordination for selective CO2 electroreduction[J]. J. Am. Chem. Soc., 2018, 140(36):  11378-11386. doi: 10.1021/jacs.8b06407

    32. [32]

      Liu X G, Geng P B, Zhang G X, Zhang Y, Dou F, Pang H. Synthesis of 2D Co-MOF nanosheets and low-temperature calcination activation for lithium-ion batteries[J]. Inorg. Chem., 2023, 62(16):  6527-6536. doi: 10.1021/acs.inorgchem.3c00783

    33. [33]

      Gao Y D, Qiu Z M, Lu Y, Zhou H J, Zhu R M, Liu Z, Pang H. Rational design and general synthesis of high-entropy metallic ammonium phosphate superstructures assembled by nanosheets[J]. Inorg. Chem., 2023, 62(8):  3669-3678. doi: 10.1021/acs.inorgchem.3c00038

    34. [34]

      Yun Q B, Lu Q P, Zhang X, Tan C L, Zhang H. Three-dimensional architectures constructed from transition-metal dichalcogenide nanomaterials for electrochemical energy storage and conversion[J]. Angew. Chem. Int. Ed., 2018, 57(3):  626-646. doi: 10.1002/anie.201706426

    35. [35]

      Lv J W, Gao X Q, Han B, Zhu Y F, Hou K, Tang Z Y. Self-assembled inorganic chiral superstructures[J]. Nat. Rev. Chem., 2022, 6(2):  125-145. doi: 10.1038/s41570-021-00350-w

    36. [36]

      Zheng S S, Zheng Y, Xue H G, Pang H. Ultrathin nickel terephthalate nanosheet three-dimensional aggregates with disordered layers for highly efficient overall urea electrolysis[J]. Chem. Eng. J., 2020, 395:  125166. doi: 10.1016/j.cej.2020.125166

    37. [37]

      Zeng G, He Y C, Ma D D, Luo S W, Zhou S H, Cao C S, Li X F, Wu X T, Liao H G, Zhu Q L. Reconstruction of Ultrahigh-aspect-ratio crystalline bismuth-organic hybrid nanobelts for selective electrocatalytic CO2 reduction to formate[J]. Adv. Funct. Mater., 2022, 32(30):  2201125. doi: 10.1002/adfm.202201125

    38. [38]

      Li L H, Liu X J, Wang J M, Liu R, Liu Y R, Wang C L, Yang W X, Feng X, Wang B. Atomically dispersed Co in a cross-channel hierarchical carbon-based electrocatalyst for high-performance oxygen reduction in Zn-air batteries[J]. J. Mater. Chem. A, 2022, 10(36):  18723-18729. doi: 10.1039/D2TA05777H

    39. [39]

      Xiong X L, You C, Liu Z A, Asiri A M, Sun X P. Co-doped CuO nanoarray: An efficient oxygen evolution reaction electrocatalyst with enhanced activity[J]. ACS Sustain. Chem. Eng., 2018, 6:  2883-2887. doi: 10.1021/acssuschemeng.7b03752

    40. [40]

      Wang Y, Wang S Q, Liu D Y, Zhou L, Du R, Li T T, Miao T T, Qian J J, Hu Y, Huang S M. Normal-pulse-voltage-assisted in situ fabrication of graphene-wrapped MOF-derived CuO nanoflowers for water oxidation[J]. Chem. Commun., 2020, 56:  8750-8753. doi: 10.1039/D0CC03132A

    41. [41]

      Usman A, Xiong F, Aftab W, Qin M L, Zou R Q. Emerging Solid-to-solid phase-change materials for thermal-energy harvesting, Storage, and Utilization[J]. Adv. Mater., 2022, 34(41):  2202457. doi: 10.1002/adma.202202457

    42. [42]

      Zheng X B, Li B B, Wang Q S, Wang D S, Li Y D. Emerging low-nuclearity supported metal catalysts with atomic level precision for efficient heterogeneous catalysis[J]. Nano Res., 2022, 15(9):  7806-7839. doi: 10.1007/s12274-022-4429-9

    43. [43]

      Sun Y Y, Ji H Q, Sun Y J, Zhang G X, Zhou H J, Cao S, Liu S X, Zhang L, Li W T, Zhu X W, Pang H. Synergistic effect of oxygen vacancy and high porosity of nano MIL-125(Ti) for enhanced photocatalytic nitrogen fixation[J]. Angew. Chem. Int. Ed., 2023, :  e202316973.

    44. [44]

      Zhu R M, Ding J W, Yang J P, Pang H, Xu Q, Zhang D L, Braunstein P. Quasi-ZIF-67 for boosted oxygen evolution reaction catalytic activity via a low temperature calcination[J]. ACS Appl. Mater. Interfaces, 2020, 12(22):  25037-25041. doi: 10.1021/acsami.0c05450

    45. [45]

      Ding Y R, Wang Z Y, Liang Z J, Sun X P, Sun Z H, Zhao Y X, Liu J L, Wang C Y, Zeng Z Y, Fu L, Zeng M Q, Tang L. A monolayer high-entropy layered hydroxide frame for efficient oxygen evolution reaction[J]. Adv. Mater., 2023, :  2302860. doi: 10.1002/adma.202302860

    46. [46]

      Tabti H A, Adjdir M, Ammam A, Mdjahed B, Guezzen B, Ramdani A, Benddedouche C K, Bouchikhi N, Chami N. Facile synthesis of Cu-LDH with different Cu/Al molar ratios: Application as antibacterial inhibitors[J]. Res. Chem. Intermed., 2020, 46(12):  5377-5390. doi: 10.1007/s11164-020-04268-8

    47. [47]

      Park S H, Kim H J. Unidirectionally aligned copper hydroxide crystalline nanorods from two-dimensional copper hydroxy nitrate[J]. J. Am. Chem. Soc., 2004, 126(44):  14368-14369. doi: 10.1021/ja047425w

    48. [48]

      Saghir S, Fu E F, Xiao Z G. Synthesis of CoCu-LDH nanosheets derived from zeolitic imidazole framework-67 (ZIF-67) as an efficient adsorbent for azo dye from waste water[J]. Microporous Mesoporous Mat., 2020, 297:  110010. doi: 10.1016/j.micromeso.2020.110010

    49. [49]

      Gholinejad M, Naghshbandi Z, Sansano J M. Co/Cu bimetallic ZIF as new heterogeneous catalyst for reduction of nitroarenes and dyes[J]. Appl. Organomet. Chem., 2020, 34(4):  e5522. doi: 10.1002/aoc.5522

    50. [50]

      Wu C H, Xie D G, Mei Y J, Xiu Z F, Poduska K M, Li D C, Xu B, Sun D F. Unveiling the thermolysis natures of ZIF-8 and ZIF-67 by employing in situ structural characterization studies[J]. Phys. Chem. Chem. Phys., 2019, 21(32):  17571-17577. doi: 10.1039/C9CP02582K

    51. [51]

      Cheng W R, Wu Z P, Luan D Y, Zang S Q, Lou X W. Synergetic cobalt-copper-based bimetal-organic framework nanoboxes toward efficient electrochemical oxygen evolution[J]. Angew. Chem. Int. Ed., 2021, 60(50):  26397-26402. doi: 10.1002/anie.202112775

    52. [52]

      Li X R, Wang C L, Liu Y Y, Xue H G, Pang H, Xu Q. Cu-alanine complex-derived CuO electrocatalysts with hierarchical nanostructures for efficient oxygen evolution[J]. Chin. Chem. Lett., 2021, 32(7):  2239-2242. doi: 10.1016/j.cclet.2020.12.037

    53. [53]

      Wu X Y, Jing Q L, Sun F C, Pang H. The synthesis of zeolitic imidazolate framework/prussian blue analogue heterostructure composites and their application in supercapacitors[J]. Inorg. Chem. Front., 2023, 10(1):  78-84. doi: 10.1039/D2QI01966C

    54. [54]

      Gao G, Zhu Z, Zheng J, Liu Z, Wang Q, Yan Y S. Ultrathin magnetic Mg-Al LDH photocatalyst for enhanced CO2 reduction: Fabrication and mechanism[J]. J. Colloid Interface. Sci., 2019, 555:  1-10. doi: 10.1016/j.jcis.2019.07.025

    55. [55]

      Ren H, Zhang L Y, An J P, Wang T T, Li L, Si X Y, He L, Wu X T, Wang C G, Su Z M. Polyacrylic acid@zeolitic imidazolate framework-8 nanoparticles with ultrahigh drug loading capability for pH-sensitive drug release[J]. Chem. Commun., 2014, 50(8):  1000-1002. doi: 10.1039/C3CC47666A

    56. [56]

      Yang J, Sun Y X, Chang X J, Zhang Z L, Wang X, Zhou G M, Peng J D. MOF-derived N-doped porous carbons activate peroxomonosulfate to efficiently degrade organic pollutants via non-radical pathways[J]. J. Water Process. Eng., 2023, 56:  104295. doi: 10.1016/j.jwpe.2023.104295

  • Scheme 1  Synthetic approach of the Cu-NF composites and their derivatives

    Figure 1  SEM images of (a) CuCo-NS, (b) Cu-NF, and (c) Cu-NP composites; HRTEM images and SAED pattern (Inset) of (d-f) the Cu-NF composites; (g) HAADF-STEM image of the Cu-NF composites and their corresponding elemental mapping images

    Figure 2  (a) XRD patterns of the samples from the reaction mixture taken at different times during the reaction; (b) XRD patterns, (c) FTIR spectrum, (d) full XPS spectrum and XPS high-resolution spectra of (e) C1s, (f) Cu2p, (g) O1s, and (h) N1s of the Cu-NF composites

    Figure 3  (a) TGA curve of the Cu-NF composites; (b) XRD patterns, (c) FTIR spectra of the Cu-NF composites and their derivatives and (d) Raman spectra

    Figure 4  SEM images of (a1) Cu-NF-200, (b1) Cu-NF-250, (c1) Cu-NF-300, (d1) Cu-NF-350, (e1) Cu-NF-400; TEM images of (a2) Cu-NF-200, (b2) Cu-NF-250, (c2) Cu-NF-300, (d2) Cu-NF-350, (e2) Cu-NF-400

    Figure 5  (a) LSV curves, (b) overpotentials at the specific current densities, (c) Tafel plots, (d) ECSAs based on CV curves (Fig.S13) at different scan rates, and (e) Nyquist plots of impedance spectroscopy analysis of the as-prepared electrocatalysts in 1.0 mol·L-1 KOH; (f) LSV curves of Cu-NF-300

    Inset in (e): fitted equivalent circuit model for the electrochemical impedance tests.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  125
  • HTML全文浏览量:  12
文章相关
  • 发布日期:  2024-05-10
  • 收稿日期:  2024-01-05
  • 修回日期:  2024-03-06
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章