Xudun
Shen
a,
Liping
Huang
a,
Shuaishuai
Li
a,
Longnian
Tang
a,
Qiumei
Lei
a,
Bowang
Zhao
a,
Huilian
Hao
a,
Wenyao
Li
*ac,
Min
Zeng
*b and
Guanjie
He
*c
aSchool of Materials Science and Engineering, Shanghai University of Engineering Science, Shanghai 201620, China. E-mail: liwenyao314@gmail.com
bDepartment of Micro/Nano Electronics, School of Electronic Information and Electrical Engineering, Shanghai Jiao Tong University, 800 Dong Chuan Road, Shanghai, 200240, PR China. E-mail: minzeng@sjtu.edu.cn
cElectrochemical Innovation Lab, Department of Chemical Engineering, University College London, London, WC1E 7JE, UK. E-mail: g.he@ucl.ac.uk
First published on 20th October 2023
We used sodium hydroxide-mediated approach and tannic acid etching to prepare hollow structured trimetallic MOF-derived CoFeNi/Z-P NC nanocomposites. Remarkably, the resulting CoFeNi/Z-P NC nanocomposites have large specific surface area and mesoporous structure, making their active sites more accessible and mass transfer more effective. More complex trimetallic components provide greater possibilities for further improving electrocatalytic performance. The CoFeNi/Z-P NC nanocomposites demonstrate notable enhancements for the OER, and 10 mA cm−2 current density is achieved at a low overpotential of 244 mV, with a low Tafel slope of 66.2 mV dec−1 and have good stability in alkaline solutions. In addition, as a cathode material for overall alkaline water splitting, CoFeNi/Z-P NC is better than RuO2 with longer cycling stability.
With the aim of resolving this issue, nanostructured transition metal phosphides (TMPs) were recognized as promising candidates for catalysts in the oxygen evolution reaction (OER), and important developments have been achieved in this area.11–13 However, these catalysts based on transition metal phosphides still need to perform better in the OER. According to studies, controlling the structure and chemical composition of transition metal-based electrocatalysts is necessary to increase their intrinsic OER activity.14,15 The first control includes changing the structure to promote mass transportation and increase the exposure of active sites.16 Given this, hollow-structured nanoporous materials have gained a lot of attention including those that have a large specific surface area, lots of active sites, and short charge transfer distances.17,18 The latter control entails purposeful modification of stoichiometric compositions to regulate the electrical characteristics, thus affecting interactions with reaction intermediates and reaction dynamics.19 Consequently, internal or phase interface engineering has emerged as a viable strategy for controlling the electrical properties.20 In their bimetallic phosphides made from MOF precursors, Chen et al. showed that a Ni2P/CoP contact was essential for producing a synergistic effect and boosting the electrocatalytic activity.21 However, due to the highly intricate chemical composition of transition metal phosphides, even slight variations in the metal-to-phosphorus stoichiometric ratio can result in substantial alterations in their structures. As a result, studies concentrating on combining hollow structures with various transition metal phosphides are scarce.
Zhang et al.22 synthesized NiCoP/NC PHCs by using bimetal as precursor and combining tannic acid-induced chemical corrosion with subsequent roasting and phosphating. The obtained composite material has a good hollow structure. Hong et al.23 developed a very simple sodium hydroxide-mediated method to prepare Fe–Co bimetallic MOFs in high yields. However, there are currently few methods for doping two other metals into ZIF-67 to form trimetallic ZIFs, and there are even fewer studies on hollow trimetallic MOFs and their derivatives.
In this study, we synthesized trimetallic CoFeNi-ZIF specimens by incorporating iron and nickel dopants into ZIF-67. Then, we used tannic acid as an etchant to perform in situ chemical etching on the cobalt-based trimetallic zeolite imidazole framework polyhedron (CoFeNi-ZIF), forming hollow-structured CoFeNi-Z. Subsequently, two steps of carbonization and phosphating were carried out. The as-synthesized CoFeNi/Z-P NC nanocomposite exhibits a distinctive hollow structure and benefits from the coordinating impact arising from the uniform integration of metallic Fe and Ni. These nanocomposites display a low overpotential of 244 mV at a current density of 10 mA cm−2 due to increased electrocatalytic activity for the OER. Additionally, as a cathode material for overall alkaline water electrolysis, CoFeNi/Z-P NC is better than RuO2 with longer cycling stability.
The as-prepared CoFeNi/Z NC and NaH2PO2·H2O at a mass ratio of 1
:
10 were placed downstream and upstream in the tube furnace. The sample was then heated to 400 °C at an average rate of 5 °C min−1 and maintained at this temperature for two hours under Ar atmosphere. After cooling to room temperature, a phosphide product called CoFeNi/Z-P NC was obtained and further utilized.
Tannic acid is used as the etching agent in in situ chemical etching to convert CoFeNi-ZIF into CoFeNi-Z (step 2 in Fig. 1). It takes 5 minutes to complete this process. The etching reaction induced by tannic acid is a surface functionalization-assisted etching process. Tannic acid provides protons to destroy the coordination bonds of CoFeNi-ZIF, causing the disintegration of the MOF. At the same time, tannic acid can be adsorbed on the surface of CoFeNi-ZIF to protect the MOF from being etched.24 SEM and TEM were employed to illustrate the unique structure and morphology of the sample (Fig. S4 and S5†). The color change from CoFeNi-ZIF to CoFeNi-Z also confirms the evolution of the structure (Fig. S6†), which is consistent with our previous findings.22 CoFeNi/ZIF-P can be obtained by the direct carbonization of CoFeNi-ZIF and then phosphating. CoFeNi-Z is first carbonized and then phosphorized to obtain CoFeNi/Z-P NC. CoFeNi/ZIF NC and CoFeNi/Z NC are obtained by the direct carbonization of CoFeNi-ZIF and CoFeNi-Z under an argon atmosphere.
X-ray diffraction, Raman spectroscopy, and BET tests were also carried out to determine the chemical state of the sample and to disclose the synergistic impact of various metal phosphides. The CoFeNi/Z-P NC alloy single-particle X-ray diffraction (XRD) pattern is shown in Fig. 2a. The diffraction peaks of CoP (PDF#29-0497), FeP2 (PDF#06-0561), and NiP2 (PDF#13-0213) appeared in CoFeNi/Z-P NC, indicating that the crystalline phases of CoP, FeP2, and NiP2 were formed.23,25,26 The D band and G band are represented by two unique peaks in the Raman spectrum at 1351 and 1596 cm−1, respectively (Fig. 2b). A 2D band peak was also observed at 2703 cm−1.27 While the G bands come from in-plane sp2 carbon vibrations, the D bands come from sp2 carbon imperfections. Additionally we noticed weaker 2D bands indicating the presence of graphene-like carbon sheets in the material.28 The calculated intensity ratio ID/IG is 0.99, 1.08, and 0.95 for CoFeNi/ZIF NC, CoFeNi/Z NC, and CoFeNi/Z-P NC electrocatalysts, respectively. The high ID/IG ratio indicates the generation of numerous defects, implying significant doping of the graphene layer with heterogeneous atoms.29 Additionally, the larger and weaker 2D bands suggest that the alloy has thin graphene layers on it. BET analysis confirms some evolution in the material structure (Fig. 2c and d), demonstrating that the CoFeNi-Z material has a specific surface area of 270.138 m2 g−1, a typical pore size of 8.885 nm, and a total volume of pores of 0.429 cm3 g−1. Comparatively, CoFeNi-ZIF has a surface area of 361.033 m2 g−1, a pore size of 6.022 nm on average, and a total volume of pores of 0.392 cm3 g−1. This further confirms the impact of etching on the material structure.
Fig. 3a shows the SEM image of CoFeNi-ZIF. It can be seen that it is a dodecahedron between 50–100 nm in size. Fig. 3b shows the scanning electron microscope image of CoFeNi-Z; the morphology of the sample after in situ etching has changed; the surface is rough and more irregular than before. Fig. 3c shows the morphology of CoFeNi/Z-P NC. The transmission electron microscope (TEM) image in Fig. 3d confirms the sample's hollow structure, and it also clearly shows the N-doped graphene shell and that the number of layers of graphene is approximately 5 to 10. The ESI† contains higher resolution HRTEM images of the hollow structure (Fig. S7†). The (200), (200), and (310) planes of FeP2, CoP, and NiP2 are depicted in the high-resolution TEM (HRTEM) image as lattice fringes with d-spacing values of approximately 0.247 nm, 0.252 nm, and 0.162 nm,30 respectively, which closely match the expected spacing values (Fig. 3e). A polycrystalline structure can be seen in the CoFeNi/Z-P NC selected area electron diffraction (SAED) pattern (Fig. 3f), as is evident from the diffraction rings. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) (Fig. 3g), which shows differential properties in the inner and outer portions, further confirms a core–shell structure for the nanocrystals, and analysis using energy-dispersive X-rays (EDS) confirms the elemental species of CoFeNi/Z-P (Fig. S8†).
Analysis of CoFeNi/Z-P NC samples using X-ray photoelectron spectroscopy (XPS) reveals important information about their chemical composition and bonding states. On the surface of CoFeNi/Z-P NC, the whole XPS spectrum in Fig. 4a verifies the existence of C, N, P, and transition metal ions Co, Fe, and Ni.31 According to the Co 2p spectra (Fig. 4b), the first pair of peaks identified at 780.18 eV (Co 2p3/2) and 792.58 eV (Co 2p1/2) corresponds to Co in the Co2+ oxidation state. The second set of peaks, located at 782.38 eV (Co 2p3/2) and 796.68 eV (Co 2p1/2), indicates the existence of Co3+ species. The peaks at 785.88 eV and 802.48 eV correspond to shakeup satellites. Additionally, the emergence of a unique spin–orbit peak at 777.78 eV provides further evidence that Co–P bonds were produced in the CoFeNi/Z-P NC sample.32,33 The peaks at 706.18 eV (Fe 2p2/3) and 718.38 eV (Fe 2p1/2), in the Fe 2p spectra (Fig. 4c) are indicative of Fe–P bonds, and peaks at 711.88 eV (Fe 2p2/3) and 723.68 eV (Fe 2p1/2) represent Fe–O surface bonds.34 The Ni 2p XPS spectrum (Fig. 4d) reveals the detailed electronic states of the Ni2P compound, exhibiting four spin–orbit peaks along with two satellite peaks labeled as “Sat.”. The existence of both Ni2+ and Ni3+ species is indicated by peaks at 852.38 eV and 855.38 eV, which are followed by peaks at 869.18 eV and 873.18 eV, respectively.35,36 A pair of peaks at 128.78 eV and 129.28 eV, attributed to P 2p3/2 and P 2p1/2, respectively, in the high-resolution P 2p spectrum (Fig. 4e), suggests the formation of metal phosphides. The peak at 133.08 eV suggests that sample oxidation in the air caused the formation of P–O bonds.37,38 Pyridinic nitrogen (397.98 eV), pyrrolic nitrogen (399.28 eV), graphitized nitrogen (400.68 eV), and nitrogen oxide (404.68 eV) correspond to the four peaks seen in the N 1s spectrum (Fig. 4f).39,40 The electrocatalytic activity is significantly impacted by both pyridinic-N and graphitic-N species; this is important to highlight.41,42 Table S1† shows the chemical compositions of CoFeNi/Z-P NC by XPS measurement.
![]() | ||
| Fig. 4 XPS spectra of CoFeNi/Z-P NC. (a) Survey scan, (b) Co 2p, (c) Fe 2p, (d) Ni 2p, (e) P 2p, and (f) N 1s peaks of CoFeNi/Z-P NC. | ||
The catalytic performance of the catalyst for the oxygen evolution reaction (OER) was assessed using a typical three-electrode electrochemical cell device in 1 M potassium hydroxide solution. The counter electrode was a graphite rod, while the reference electrode was a saturated Ag/AgCl electrode. LSV was used to assess the OER activity of samples and this was compared to that of a commercial RuO2 electrocatalyst used as a benchmark. The LSV curves in Fig. 5a demonstrate the excellent performance of the CoFeNi/Z-P NC catalyst, which has an overpotential of 244 mV at a current density of 10 mA cm−2. Compared to the values presented in Fig. 5b for RuO2 (344 mV), CoFeNi/ZIF NC (319 mV), CoFeNi/Z NC (299 mV), and CoFeNi/ZIF-P NC (269 mV), this value is much lower. The Tafel slope is then obtained by fitting the linear component of the Tafel diagram with the Tafel equation (η = blog
j + a), where b is the Tafel slope and j is the current density. RuO2 has a value of 97.4 mV dec−1 on the Tafel slopes in Fig. 5c, which is consistent with the theoretical value. Compared to CoFeNi/ZIF NC, CoFeNi/Z, and CoFeNi/ZIF-P NC, which had Tafel slopes of 59.7 mV dec−1, 65 mV dec−1, and 68.4 mV dec−1, respectively, CoFeNi/Z-P NC had a measured Tafel slope of 66.2 mV dec−1. We compared the OER electrocatalytic performance of the obtained catalyst with that of different catalysts (Table S2†). At the same time, we compared the RuO2 parameter obtained from the experiment with the literature reports (Table S3†). The Nyquist plots in Fig. 5d indicate that CoFeNi/Z-P NC exhibited smaller semicircle diameters compared to CoFeNi/ZIF NC, CoFeNi/Z NC, and CoFeNi/ZIF-P NC, suggesting a decreased charge transfer resistance at the catalyst/electrolyte contact. A decreased charge transfer resistance denotes greater electrode material conductivity, which is generally recognised to be connected to electrocatalysis kinetics.
Furthermore, cyclic voltammetry was used to quantify the electrochemical double-layer capacitance (Cdl), which is thought to be directly proportional to the electrochemically active surface area (ECSA). The CV curves for CoFeNi/ZIF NC, CoFeNi/Z NC, CoFeNi/ZIF-P NC, and CoFeNi/Z-P NC in the non-faradaic area are shown in Fig. S8† at various scan rates (20, 40, 60, 80, and 100 mV s−1). As depicted in Fig. 5e, the calculated Cdl value for CoFeNi/Z-P NC is 5.63 mF cm−2, surpassing those of CoFeNi/ZIF NC (1.06 mF cm−2), CoFeNi/Z NC (2.73 mF cm−2), and CoFeNi/ZIF-P NC (3.64 mF cm−2). With only a slight decline in current density over a 10-hour period, Fig. 5f and the inset show that CoFeNi/Z-P NC has strong durability. After this step, we collected the sample and performed TEM analysis (Fig. S10†). The graphene shell is clearly visible and the lattice parameters correspond to those before testing (Fig. 3e).
As shown in Fig. 6a, a current density of 10 mA cm−2 could be achieved by applying a potential of 1.588 V between two electrodes, which is even smaller than that of its Pt-RuO2 counterpart (1.627 V). The very stable current density at a high potential of 1.59 V as shown in i–t curves (Fig. 6b) further demonstrated the strong stability of CoFeNi/Z-P NC. The negligible activity of carbon paper further proved that the activity came from the catalysts. The excellent activity of the alloy catalyst, accompanied by high stability, made it a potential alternative to precious metal catalysts in practical water splitting.
Footnote |
| † Electronic supplementary information (ESI) available: The characterization instrument model of the sample, the method of electrochemical measurement, and the synthesis process diagram of the sample and the SEM, TEM, XRD, EDS and other information of the comparative samples. See DOI: https://doi.org/10.1039/d3dt02818f |
| This journal is © The Royal Society of Chemistry 2023 |