Jiaqi
Zhou
a,
Yuhong
Zhang
a,
Tianrui
Yu
a,
Mingxin
Feng
a,
Tong
Wang
a,
Chuangyi
Tong
a,
Zewu
Zhang
b,
Jiehua
Bao
*b and
Yuming
Zhou
*a
aSchool of Chemistry and Chemical Engineering, Southeast University, Jiangsu Optoelectronic Functional Materials and Engineering Research Center, Nanjing, 211100, Jiangsu Province, China. E-mail: ymzhou@seu.edu.cn
bSchool of Materials Science and Engineering, Nanjing Institute of Technology, Nanjing, 211167, Jiangsu Province, China. E-mail: baojh@njit.edu.cn
First published on 25th November 2024
Reasonable morphology regulation and electronic structure modulation enhance the oxygen evolution reaction (OER) performance of the catalyst. In this study, amorphous citrate-chelated CoNiFe trimetallic hydroxide nanoparticles were synthesized in one step via coprecipitation at room temperature. The choice of solvents controlled the hydrolysis rate of metal cations, allowing for the regulation of the product morphology. An alcohol–water system as the solvent facilitated the formation of more uniform and well-dispersed nanoparticles. Citrate was employed as a chelating agent, and its strong interaction with metal cations improved the stability of the amorphous materials, regulated the particle size, and increased the electrochemically active surface area. Furthermore, varying amounts of Ni ions were doped to modulate the electronic structure, exerting tri-metallic synergistic effects, which enhanced the OER performance. The results demonstrated that, in 1.0 M KOH, the optimized Co:
Ni
:
Fe molar ratio of 2
:
1
:
1 achieved the highest OER activity, with an overpotential of 287 mV and a Tafel slope of 56.3 mV dec−1, delivering a current density of 10 mA cm−2, and maintaining stable performance over 24 hours with only a minor increase in the overpotential.
In recent years, sulfides,6–8 nitrides,9,10 phosphides11,12 and selenides13,14 based on non-precious metal transition metals (Co, Ni, Fe, and Mn) have been extensively investigated, although the synthesis of these high-purity materials remains both time-consuming and technically challenging. However, these materials are prone to oxidation into their corresponding metal oxides or hydroxides. Under alkaline OER conditions, the hydroxide phase of these metals acts as the actual catalytic site.15 Consequently, transition metal hydroxides hold significant potential as OER electrocatalysts.16,17
To enhance the electrocatalytic activity of cost-effective transition metal hydroxides, numerous constructive strategies have been proposed, including morphology control, metal doping, interlayer distance adjustment, defect engineering, phase engineering, and interface engineering. Zhou et al.18 prepared ultra-thin NiFe-layered double hydroxides (LDHs) enriched with oxygen vacancies (VO) using a coprecipitation method in an alcohol–water system at room temperature. They investigated how different alcohol-to-water ratios affected the product structure and catalytic performance. N. K. Shrestha et al.19 performed a simple immersion chemical etching of foam nickel (NF) in the ethanol FeCl3 solution to generate a microporous nickel (Ni) framework decorated with hierarchically structured metallic Fe doped Ni–Fe-hydroxide nanoparticles. The synergistic interaction between uniformly distributed micropores and metal iron doped iron hydroxide nanocatalysts, as well as the local amorphization in the lattice, promotes the transfer of mass and charge, thereby demonstrating excellent electrocatalytic performance and stability, and can operate for 80 hours at a current density of 400 mA cm−2 without evidence of voltage decay. Citrate is frequently employed as an intercalation agent to aid in the synthesis of NiFe-LDHs,20,21 and it also serves as a chelating agent.22 Citrate enhances the synergistic effect between Ni and Fe.20 Additionally, citrate salts can effectively reduce and control the nanoparticle size.23,24 Xu et al.20 utilized trisodium citrate as an intercalating agent to produce a series of partially delaminated NiFe-LDHs by varying citrate contents. They examined the impact of the citrate concentration on interlayer spacing, morphology, surface elemental composition, surface metal electronic states, and electrochemical properties. Metal ion doping can modify the electronic structure and optimize adsorption/desorption properties for reactants and intermediates.25,26 Yang et al.27 synthesized a range of CoFeM-LDH nanorods doped with various inexpensive metals using a one-step hydrothermal method, designated as CoFeM-x (M = Cu, K, Al, Zn, and Cr; x = 5, 10, 15, and 20). CoFeK-15 exhibited the best OER performance, attributed to K ion-induced lattice distortion that optimized the electronic structure and increased oxygen vacancies. Chen et al.28 developed hierarchical trimetallic NiFeCe-LDH with intercalated citrate via a coprecipitation method, and the introduction of Ce ions enhanced intrinsic activity by modifying the electronic structure through 3d–4f electron interactions. Currently, most of the studied OER catalysts are crystalline. Amorphous materials, with their abundance of coordination-unsaturated atoms, often demonstrate superior OER electrocatalytic activity compared to crystalline materials. Additionally, the defects and inherent structural disorder in amorphous materials facilitate charge transfer and ion diffusion.29–31 N. K. Shrestha et al.32 induce site-selective disordering of the crystalline structure of MIL88B(Ni) frameworks via Ce-doping, resulting in crystalline/amorphous heterostructures. DFT calculations support the experimentally observed Ce3+ ion doping effect inducing site selective crystal disorder on the MIL-88B(Ni) framework structure, thereby enhancing OER electrocatalytic activity. However, amorphous materials with metastable structures can show instability during measurements in acidic or alkaline solutions.33,34 Poor stability will seriously affect the practical application prospects of catalysts.
Herein, we synthesized CoFe hydroxide particles containing carbonate ions in a single step using the coprecipitation method at room temperature, and selected various solvents to control the product morphology. Ammonium citrate was used as a chelating agent to control the particle size and increase the number of exposed active sites, thereby enhancing the catalyst's performance. Research indicates that the addition of citrate not only improves the OER performance of the catalyst but also significantly enhances its stability during the OER process. Furthermore, Ni ions were doped to modify the electronic structure and examine the effect of doping concentration on the morphology and catalytic performance of the product. Finally, we synthesized amorphous citrate-chelated CoNiFe trimetal hydroxide nanoparticles using an alcohol–water solvent in a single step at room temperature. Under 1.0 M KOH conditions, the catalyst achieved a current density of 10 mA cm−2 with an overpotential of only 287 mV and maintained stable performance over 24 hours with a minimal increase in overpotential.
ERHE = E(Hg/HgO) + 0.098 V + pH × 0.059 |
Prior to each electrochemical test, the working electrode was activated by cycling the potential from 0 to 1.6 V versus Hg/HgO for 20 cycles at a scan rate of 100 mV s−1. Linear sweep voltammetry (LSV) curves were recorded at a scan rate of 5 mV s−1 with 90% iR compensation. Double-layer capacitance (Cdl), which is linearly proportional to the electrochemical surface area (ECSA), was estimated from cyclic voltammetry (CV) curves within the non-faradaic potential window of 0.25 to 0.35 V versus Hg/HgO, using incremental scan rates of 20, 40, 60, 80, and 100 mV s−1. Electrochemical impedance spectroscopy (EIS) data were collected over a frequency range of 100000 Hz to 0.1 Hz with an applied voltage of 0.25 V versus Hg/HgO.
The morphology and nanostructure of the product were studied using SEM and TEM. The SEM images of Co2NiFe-CA are shown in Fig. 2a and b. Co2NiFe-CA consists of nanoparticles. To investigate the effects of different conditions on the product morphology, SEM images of other samples were also taken (Fig. S1–S8, ESI†). Co3Fe-W consists of irregular particles and flakes that aggregate, which can be attributed to the faster hydrolysis rate leading to a more rapid precipitation growth and larger particle sizes. Co3Fe-E particles are smaller but exhibit a low yield across multiple experiments. This is because NH4HCO3 generates fewer hydroxide ions in pure ethanol, leading to insufficient precipitation with metal ions. Co3Fe-W/E particles have an average diameter of approximately 95 nm and are more uniform compared to those in the pure ethanol system. This is attributed to the presence of water, which increases the concentration of hydroxide ions and facilitates particle growth. The addition of ammonium citrate resulted in Co3Fe-CA particles with a smaller average diameter of approximately 72 nm. This indicates that ammonium citrate effectively regulates the particle size, resulting in a higher specific surface area. This is because citrate ions complex with metal cations, competing with hydroxide ion precipitation and slowing down the nucleation rate during the reaction.21,37 Following the addition of a small amount of Ni ions, there was no significant change in the product (Fig. S5 and S6, ESI†). However, when the Co/Ni molar ratio reached 1:
2 (Fig. S7, ESI†), a noticeable increase in the particle size and agglomeration occurred, which may be due to structural collapse from the in situ synthesis of cobalt nickel iron hydroxide. This phenomenon was also observed by Wu et al. during their preparation of amorphous/crystalline CoNiFe-LDH hollow nanocages.33 Crystalline Co2NiFe-CS is composed of aggregated irregular flakes. The particle size of the aggregates is significantly larger than that of Co2NiFe-CA. The particle size distributions of the different catalysts (except for irregular Co3Fe-E and Co2NiFe-CS) were obtained based on SEM maps (Fig. S9, ESI†), which shows that the particle sizes are homogeneous. In addition, EDX was employed to determine the elemental ratio of the sample, revealing that the molar ratio of Co
:
Ni
:
Fe matched that of the raw material for Co2NiFe-CA, as shown in Fig. S1–S8 (ESI†). The TEM images of Co2NiFe-CA are shown in Fig. 2c and d. The absence of lattice fringes in the TEM images confirms that Co2NiFe-CA has an amorphous structure. The crystal composition of all the prepared products was analyzed using XRD (Fig. S10, ESI†). Fig. 3a shows that Co3Fe-W and Co2NiFe-CS are crystalline products. Co3Fe-W exhibits characteristic peaks of hydrotalcite, corresponding to Co5.84Fe2.16-LDH (JCPDS No. 50-0235). The four characteristic peaks are located at 11.5°, 23.3°, 34°, and 59.3°, corresponding to the (003), (006), (012), and (110) planes, respectively, which align with the layered structure observed in the SEM image of Co3Fe-W. Co2NiFe-CS also exhibits hydrotalcite characteristic peaks. When using ethanol or alcohol–water system as solvents, the characteristic hydrotalcite peaks disappeared (Fig. 3b), and the XRD pattern of the product exhibited a completely amorphous state. This is likely due to NH4HCO3 generating fewer hydroxide ions in ethanol-rich solvent systems, preventing the formation of a layered hydrotalcite structure. Fig. 3c presents the XRD pattern of pure carbon paper and Co2NiFe-CA after CP testing at a current density of 10 mA cm−2. Following CP testing, XRD tests were carried out on Co2NiFe-CA stuck on carbon paper and Co2NiFe-CA scraped off from carbon paper, respectively. It can be seen that the XRD patterns of pure carbon paper and Co2NiFe-CA stuck on carbon paper are almost identical because the signal of amorphous products is masked by the signal of carbon paper. And Co2NiFe-CA scraped off from carbon paper also exhibits amorphous characteristics. This indicates that after testing, Co2NiFe-CA still has an amorphous structure and no new crystalline material has been generated.
![]() | ||
Fig. 2 (a) and (b) SEM images. (c) and (d) HRTEM images. (e) Elemental mapping images of Co2NiFe-CA. |
FT-IR characterization was performed to verify the presence of citrate anions in Co3Fe-CA, Co2NiFe-CA, Co1.5Ni1.5Fe-CA, CoNi2Fe-CA and Co2NiFe-CS (Fig. 3d). In the spectrum of Co3Fe-W and Co2NiFe-CS, bands at around 1364 cm−1 and 1490 cm−1 are associated with CO32− vibrations, while the band at around 3450 cm−1 is characteristic of O–H stretching vibrations.38–40 For Co3Fe-E and Co3Fe-W/E, the CO32− vibrational bands remain at 1384 cm−1 and 1490 cm−1. For Co3Fe-CA, Co2NiFe-CA, Co1.5Ni1.5Fe-CA, and CoNi2Fe-CA, new bands at around 1399 cm−1 and 1573 cm−1 are observed, assigned to the RCOO− symmetric and asymmetric stretches.40 Additionally, signals at 1252 cm−1 and 1073 cm−1 correspond to C–O stretching, while the signal at 905 cm−1 is attributed to C–H bending.41 These results confirm the presence of citrate anions in Co3Fe-CA, Co2NiFe-CA, Co1.5Ni1.5Fe-CA, and CoNi2Fe-CA.
The electronic structure and surface elemental composition of metal sites are closely related to the catalytic activity of catalysts.38,39,42 X-ray photoelectron spectroscopy (XPS) was employed to investigate the electronic structure of Co3Fe-W/E, Co3Fe-CA, and Co2NiFe-CA. The XPS survey scan shown in Fig. 4a reveals the presence of Co, Ni, and Fe in Co2NiFe-CA. All XPS data in this study were corrected for C 1s spectral contributions. Co2+ and Co3+ are represented by two peaks in the Co 2p XPS spectrum of Co3Fe-W/E (Fig. 4b), with Co3+ at 781.4 and 797.2 eV, and Co2+ at 783.7 and 799.5 eV. The remaining four peaks correspond to the satellite (Sat.) peaks of Co 2p3/2 and Co 2p1/2. Additionally, shake-up satellite peaks were observed at 786.0 eV and 802.5 eV, characteristic of cobalt oxide.36,43 Upon the addition of ammonium citrate (Co3Fe-CA), the peak position of Co3+ shifted to a lower binding energy (−0.3 eV) compared to Co3Fe-W/E, and the Co3+/Co2+ molar ratio decreased from 2.72 to 1.94. However, with additional nickel doping (Co2NiFe-CA), the Co3+/Co2+ molar ratio increased to 2.10 but did not exceed the initial 2.72. It has been reported in the literature that high valence cobalt enhances electrochemical activity,44 suggesting that changes in cobalt's electronic structure are not the direct cause of performance improvement. In the Fe 2p spectrum of Co3Fe-W/E (Fig. 3b), peaks at 712.5 and 725.3 eV correspond to Fe 2p3/2 and Fe 2p1/2 of Fe3+, while peaks at 710.7 and 723.2 eV correspond to Fe 2p3/2 and Fe 2p1/2 of Fe2+.27 Upon comparison, it is observed that adding ammonium citrate (Co3Fe-CA) shifts the peak position of Fe3+ to higher binding energy by 0.4 eV compared to Co3Fe-W/E, and the molar ratio of Fe3+/Fe2+ significantly increases from 1.09 to 1.49. Xu et al.20 reported a similar phenomenon in their study of citrate-intercalated NiFe-LDH, where the molar ratio of Fe3+/Fe2+ increased further with additional citrate. Research generally indicates that high-valence iron enhances oxygen evolution reaction (OER) activity.45 After further nickel doping (Co2NiFe-CA), the peak position of Fe3+ shifts to a lower binding energy by 0.3 eV compared to Co3Fe-CA, and the Fe3+/Fe2+ molar ratio decreases to 1.12. This shift may be attributed to electron transfer from Ni to Fe, as explained by Yao et al.46 through electron coupling. In the Ni 2p spectrum of Co2NiFe-CA (Fig. 3c), the peaks at 857.4 and 875.3 eV correspond to Ni 2p3/2 and Ni 2p1/2 of Ni3+, respectively, while the peaks at 856.0 and 873.5 eV are assigned to Ni 2p3/2 and Ni 2p1/2 of Ni2+. The remaining two peaks are attributed to the double satellite (Sat.) peaks of Ni 2p3/2 and Ni 2p1/2, located at 861.5 and 879.5 eV, respectively.46 The molar ratio of Ni3+/Ni2+ is 0.48. It is widely accepted that catalysts with a higher proportion of high-valence Ni components can more readily form Ni4+-OO or Ni3+-OO, which serve as active centers for the oxygen evolution reaction (OER).47 Based on the XPS spectral analysis, after adding ammonium citrate, the Co3+/Co2+ molar ratio decreases while the Fe3+/Fe2+ molar ratio increases. This suggests that electrons may have transferred from iron to cobalt, leading to an increase in high-valence iron, which enhances the oxygen evolution reaction (OER). Subsequent doping with Ni ions results in a slight increase in the Co3+/Co2+ molar ratio, while the Fe3+/Fe2+ molar ratio decreases significantly. This indicates that electrons may have transferred from Ni to Fe, resulting in the formation of more Ni3+, which enhances the OER through a synergistic effect of the three metals, optimizing electron transfer and the electronic structure.36,46
![]() | ||
Fig. 4 (a) Overall XPS surveys of Co3Fe-W/E, Co3Fe-CA and Co2NiFe-CA. High resolution XPS spectra of (b) Co 2p, (c) Fe 2p and (d) Ni 2p. |
![]() | ||
Fig. 5 Electrocatalytic OER performance in 1.0 M KOH electrolyte. (a) LSV curves. (b) Tafel slopes. (c) Double-layer capacitances (Cdl). (d) Chronopotentiometric curves of Co2NiFe-CA at 10 mA cm−2. |
The electrochemically active surface areas (ECSAs) of all catalysts were determined by electrochemical double layer capacitance measurements.50 The double layer capacitance was estimated by plotting Δj(ja − jc) against the scan rate, where ja and jc are the anodic and cathodic current densities, respectively, at 0.1 V vs. Hg/HgO, as derived from the CV curves in 1 M KOH (Fig. S13, ESI†). The slope of the fitted linear curve, which is twice the double layer capacitance, was used to represent the ECSA.47Fig. 5c shows that the Cdl values for Co3Fe-W, Co3Fe-E, Co3Fe-W/E, and Co3Fe-CA are estimated to be 1.1, 1.7, 2.4, and 7.9 mF cm−2, respectively. The variation in Cdl corresponds to the activity order, suggesting that the increased ECSA, resulting from the use of an alcohol–water solvent and the addition of ammonium citrate, exposes more edge active sites. However, the Cdl for Co2NiFe-CA, doped with a small amount of Ni ions, decreased to 4.9 mF cm−2. As the doping amount increased, Cdl continued to decline, indicating that Ni ion doping can lead to structural collapse and reduced ECSA. Nevertheless, the doping of nickel ions optimizes the electronic structure, exerts a synergistic effect among the three metals, increases the activity per unit area, compensates for the reduction of ECSA, and enables Co2NiFe-CA to exhibit optimal OER activity. The ECSA of crystalline Co2NiFe-CS is 1.7 mF cm−2, significantly lower than that of amorphous Co2NiFe-CA.
To further investigate interfacial kinetics between the electrolyte and catalysts, electrochemical impedance spectroscopy (EIS) measurements were conducted (Fig. S14, ESI†). The Rct values are provided in Fig. S15 (ESI†). Co2NiFe-CA exhibited the optimized Rct value, indicating reduced charge transfer resistance and faster electron transfer rate due to the combined effects of appropriate Ni ion doping and ammonium citrate addition. The Rct value of Co2NiFe-CA is 1.45 ohms, much lower than that of crystalline Co2NiFe-CS (2.43 ohms). The durability of the Co2NiFe-CA catalyst was assessed using chronopotentiometric measurements at a fixed current density of 10 mA cm−2 and room temperature, as shown in Fig. 5d. The Co2NiFe-CA catalyst exhibited minimal voltage attenuation during the 24-hour test, demonstrating its good durability. At the same time, the OER durability of all other catalysts was tested using chronopotentiometry at a current density of 10 mA cm−2 (Fig. S16, ESI†), and it can be seen that the durability of the catalyst was significantly improved after adding citrate, while the durability was not affected after further doping with nickel ions. In addition, to investigate the durability of Co2NiFe-CA at higher current densities, we conducted CP testing at a current density of 250 mA cm−2 and performed XRD, XPS, and SEM tests on the tested samples. CP testing shows (Fig. S17, ESI†) that Co2NiFe-CA can remain stable for over 5 hours at a current density of 250 mA cm−2. The XRD pattern (Fig. S18, ESI†) shows that after CP testing, Co2NiFe-CA still has an amorphous structure. By comparing the XPS (Fig. S19, ESI†) of Co2NiFe-CA before and after the CP test, it can be found that the proportion of trivalent nickel and cobalt has increased significantly after the test. Meanwhile, the positions of the trivalent peaks of the three metals have all shifted positively, which is conducive to the oxygen evolution reaction. The SEM image of Co2NiFe-CA coated on carbon paper after CP testing was taken (Fig. S20, ESI†), which shows that the morphology of the catalyst particles did not change and remained in a spherical structure.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4nj04240a |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025 |