Qian
Yang‡
a,
Yaao
Li‡
a,
Yaoxin
Wu
a,
Yuxiang
Li
a,
Chenxia
Yang
a,
Lili
Ban
a,
Yunxia
Zhao
a,
Bin
Dai
a,
Gang
Wang
a,
Yongsheng
Li
a,
Jinli
Zhang
b,
Zongyuan
Wang
*a,
Huan
Pang
*c and
Feng
Yu
*a
aKey Laboratory for Green Processing of Chemical Engineering of Xinjiang Bingtuan, School of Chemistry and Chemical Engineering, Shihezi University, Shihezi 832003, China. E-mail: yufeng05@mail.ipc.ac.cn; zywang@shzu.edu.cn
bSchool of Chemical Engineering & Technology, Tianjin University, Tianjin 300384, China
cSchool of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou 225009, Jiangsu, P. R. China. E-mail: panghuan@yzu.edu.cn
First published on 27th January 2025
The rational design and synthesis of oxygen evolution reaction (OER) electrocatalysts remain critical challenges for water electrolysis in hydrogen production. This study used a strategy to activate the lattice oxygen mechanism (LOM) pathway in Co(OH)2 through uniform co-doping with metallic Fe and nonmetallic N, thereby forming N–Co–O–Fe moieties at the Fe,N–Co(OH)x interface. The synergistic effects of Fe and N accelerated electron redistribution from Co to Fe atoms, promoting the formation of active high-valent Co(IV) and stimulating lattice oxygen activation. The intrinsic activity of Co(OH)2 was enhanced. The as-synthesized Fe,N–Co(OH)x exhibited exceptional performance, with high mass activity (1705 A gmetal−1) and turnover frequency (2.521 s−1), surpassing those of W,N–Co(OH)x by 80.4 and 57 times (21.2 A gmetal−1 and 0.044
s−1), respectively. In situ spectroscopy and 18O isotope-labeled differential electrochemical mass spectrometry confirmed that Fe,N–Co(OH)x achieved direct intramolecular lattice oxygen coupling via the LOM pathway during the OER process. Density functional theory calculations revealed that Fe and N co-doping synergistically modulated the d-band center of Co in Fe,N–Co(OH)x, reducing the energy barrier for OO* desorption to form oxygen vacancies. The proposed method facilitated the preparation of heteroatom-doped hydroxide catalysts to activate the LOM pathway in the OER by co-regulating multiple defects.
Broader contextThe generation of hydrogen through electrochemical water splitting is considered a highly promising approach for harvesting energy and alleviating intermittent availability issues associated with renewable energy sources. However, the overall efficiency of water splitting is strikingly hampered by the sluggish kinetics involved in the anodic oxygen evolution reaction (OER). The lattice oxygen oxidation mechanism (LOM) could enable direct *O–O* coupling, thus providing more efficient OER processing. Construction of “M–O–M” moieties at nanoscale interfaces could accelerate electron redistribution and stimulate direct coupling of intramolecular lattice oxygen. In this study metallic Fe and nonmetallic N co-doped Co(OH)2 (Fe,N–Co(OH)x) was successfully synthesized through plasma discharge in water. The co-doping with metallic Fe and nonmetallic N facilitated the formation of N–Co–O–Fe moiety molecules at the Fe,N–Co(OH)x interface, accelerated electron redistribution from Co to Fe atoms, facilitated the formation of active high-valent Co(IV) and triggered the LOM pathway. This work provides a good reference for the rational design of effective OER catalysts for water electrolysis. |
Heteroatom doping can effectively activate the LOM pathway.9–11 Wang et al. incorporated Zn2+ into CoOOH, while Hou et al. doped Cu into CoFe-layered double hydroxides (LDH).12,13 Incorporating Zn2+ facilitated the formation of accessible non-bonded oxygen states and enhanced Co–O covalency, whereas Cu doping promoted intramolecular electron transfer, activating the LOM pathway. In addition, heteroatom doping can induce interfacial catalysis at the nanoscale, consequently altering the electronic structure of the catalyst, facilitating lattice oxygen activation, and modifying catalytic activity.14–16 Ou et al. introduced Ga3+ into Co3O4 to create Co2+–O–Co3+ motifs, which successfully activated lattice oxygen.17 Similarly, Yang et al. used O species in CoFe-PBA, forming Co–O–Fe motifs that enhanced the lattice oxygen activation and improved OER performance.14 Liu et al. introduced S and FeOOH on Co(OH)2 nanoneedle arrays, thereby inducing Co–O–Fe motifs at the heterogeneous interface, accelerating electron transfer, and triggering the LOM pathway.18 Thus, the construction of “M–O–M” motifs at nanoscale interfaces accelerates electron redistribution and enhances metal–O covalency, consequently stimulating direct intramolecular lattice oxygen coupling. However, Chen et al. highlighted that the formation of oxygen vacancies (Ov) is the rate-limiting step of the LOM pathway.19 In many cases, the removal of OO* or OOH* intermediate species occurs in this step. Thus, the key challenge in activating the LOM pathway is facilitating lattice oxygen participation in the OER to form Ov. This necessitates the maintenance of a high O exchange capacity at the Ov site and a robust H exchange capacity following the occupation of the site by OH* or OOH*.
With its LDH groups and hydrotalcite-like structures, α-Co(OH)2 can expose several surface-active catalytic sites and provide sufficient structural flexibility for the LOM.20 However, α-Co(OH)2 has a high Ov formation energy and is not naturally inclined to follow the LOM pathway. Moreover, its limited electrical conductivity further impedes charge injection and extraction during O2 evolution. To enhance its activity, Fe species are commonly incorporated with α-Co(OH)2's to form highly active OER catalysts, such as CoFe-LDH and Fe2O3/Co(OH)2.21–23 Fe can effectively modify the hydrogen exchange capacity of α-Co(OH)2 for OOH and OH groups,24–26 while N-doping can adjust the conductivity and electronic structure of Co(OH)2.27,28 Nevertheless, to the best of our knowledge, the activation of the LOM pathway in Co(OH)2 has not been reported. Furthermore, studies have indicated that co-doping with metals and nonmetals can effectively enhance intrinsic conductivity and modulate electronic interactions, thereby improving the intrinsic catalytic activity.10,29,30 However, simultaneously realizing metallic and nonmetallic doping in a single-step process remains challenging. This study aimed to identify an optimal metallic and nonmetallic co-doped structure that could activate the LOM mechanism and improve electron transfer capabilities.
This study proposed a rapid and environmentally friendly method for synthesizing metallic Fe and nonmetallic N co-doped Co(OH)2 (i.e., Fe,N–Co(OH)x) with abundant Ovvia plasma discharge at the gas–water interface. In an aqueous solution, Co2+ and 2-methyl-1H-imidazole (2-MI) facilitated the formation of N-doped Co(OH)2, while Fe rods, used as discharge electrodes, served as the Fe source. The Fe atoms were uniformly sputtered into the N-doped Co(OH)2. By adjusting the Co/2-MI ratio, Fe,N–Co(OH)x and Fe-doped ZIF67 (Fe-ZIF67) were synthesized. W,N–Co(OH)x was also synthesized as a control. Co-doping with Fe and N promoted the formation of N–Co–O–Fe moieties at the Fe,N–Co(OH)x interface, thereby modulating electron redistribution between Co and Fe and optimizing the adsorption and desorption of oxygen intermediates. This enhanced the intrinsic activity of Co(OH)2. In situ ATR-FTIR, in situ18O isotope-labeled differential electrochemical mass spectrometry (DEMS), and in situ Raman spectroscopy confirmed that Fe and N co-doping activated lattice oxygen. This established the LOM pathway as the primary reaction mechanism and facilitated the formation of active high-valent Co(IV) species. Furthermore, in situ electrochemical impedance analysis and density functional theory (DFT) calculations were performed to investigate further the impact of Fe and N co-doping on OER activity and LOM pathway activation.
Transmission electron microscopy (TEM) images of Fe,N–Co(OH)x (1:
1) revealed a hexagonal plate-like morphology (Fig. 1f), thereby corroborating the SEM findings. The staggered lattice fringes indicated abundant defects41–43 (Fig. 1i), with measured lattice spacings of 2.60 and 2.67 nm, corresponding to the (102) and (100) planes of α-Co(OH)2, respectively.44 Fe,N–Co(OH)x (1
:
1) exhibited polycrystalline characteristics (Fig. 1j). Energy-dispersive spectroscopy (EDS) mapping indicated Co and O as dominant elements, with Fe and N homogeneously distributed throughout the hexagonal plate region. This confirmed uniform co-doping (Fig. 1k). Inductively coupled plasma (ICP) measurements revealed that the Co and Fe contents in Fe,N–Co(OH)x (1
:
1) were 43.37 wt% and 8.98 wt% (molar ratio of Co/Fe = 4.83), respectively (Table S2, ESI†). Furthermore, W,N–Co(OH)x (1
:
1) exhibited similar 2D morphology and uniform W,N-doping (Fig. S7 and S8, ESI†).
X-ray diffraction (XRD) analysis confirmed the primary crystal structure of α-Co(OH)2 (JCPDS No. 46-0605) and Co(OH)2 (JCPDS No. 45-0031) (Fig. 1l).45 When decreasing the Co2+/2-MI ratio from 4:
1 to 1
:
4, the shape of peaks at 18.8°, 32.2°, 37.7°, and 51.1° became sharper, and the peak intensity enhanced, indicating an increased crystallinity. Fourier transform infrared (FTIR) spectroscopy of Fe,N–Co(OH)x (4
:
1), (1
:
1), and (1
:
4) exhibited vibration bands at 480, 656, 3447, and 3632 cm−1, corresponding to Co–OH, Co–O, –OH groups of typical α-phase hydroxides, and O–H stretching vibrations of interlayer water, respectively (Fig. S9, ESI†).45,46 With increasing 2-MI content, the peaks at 480 and 3632 cm−1 intensified, indicating increased interlayer OH groups. This was consistent with the XRD results, thereby indicating improved crystallinity as the Co2+/2-MI ratio shifts from 4
:
1 to 1
:
4. The NO3− vibration at 1383 cm−1 was attributed to interlayer NO3−,47 which disappeared in Fe-ZIF67 and ZIF-67. Increasing the 2-MI ratio to 58 eq., typical for ZIF-67 synthesis,40 caused the Fe-ZIF67 to exhibit an XRD pattern similar to ZIF-67 (Fig. 1l). The FTIR spectra of Fe-ZIF67 also matched that of ZIF-67 (Fig. S9, ESI†), with peaks between 600 and 1600 cm−1 attributed to imidazole ring stretching and bending.48 Peaks at 425 and 1580 cm−1 corresponded to Co–N and C
N stretching of Co/2-MI hybrids. These findings confirmed the synthesis mechanism illustrated in Fig. 1a. Plasma induced Co2+/2-MI reconstitution into N-doped α-Co(OH)2 at low 2-MI concentrations, incorporating NO3− into the 2D α-Co(OH)2 interlayer. At higher 2-MI concentrations, the coordination of 2-MI with Co2+ facilitated the formation of ZIF-67. The competition between 2-MI and OH− coordination with Co2+ yielded various morphologies. This highlighted the importance of the Co2+/2-MI ratio in synthesizing Fe,N–Co(OH)x.
XPS analysis was conducted to identify the electronic structure and elucidate the chemical states (Fig. S10 and S11, ESI†). The Co 2p spectra of Fe,N–Co(OH)x are shown in Fig. S11a (ESI†). Two dominant Co 2p1/2 and Co 2p3/2 peaks with two shakeup satellites were observed.46 The peaks at 780.7 and 782.6 eV corresponded to Co3+ (t62ge0g) and Co2+ (t62ge1g), respectively. Fe,N–Co(OH)x (1:
1) contained the highest percentage of Co3+ (56.4%) (Table S4, ESI†), which may benefit OER performance.49 Notably, there was a slightly negative shift (0.3 eV) of the Co 2p3/2 peak between Fe,N–Co(OH)x (1
:
1), and Fe-ZIF67 (Fig. S11a, ESI†). This indicated increased electron cloud density around the Co atoms after Fe,N co-doping.24 The O 1s spectra of Fe,N–Co(OH)x (1
:
1) exhibited four peaks at 529.9, 530.8, 531.4, and 532.3 eV, attributed to lattice oxygen (metal–O), metal–OH, Ov, and surface-absorption oxygen, respectively (Fig. S11d, ESI†).13 Fe,N–Co(OH)x (1
:
1) exhibited the highest Ov content (51.9%). Electron paramagnetic resonance (EPR) confirmed abundant Ov formation (Fig. S12, ESI†),50 thereby demonstrating that rational Fe,N co-doping facilitated Ov formation. The Fe 2p XPS spectra of Fe,N–Co(OH)x (1
:
1) was deconvoluted into four main peaks representing Fe2+ (711.9 and 722.0 eV) and Fe3+ (715.0 and 725.5 eV) with two satellite peaks (718.7 and 733.14 eV, respectively) (Fig. S11b, ESI†).25 The N 1s spectra of Fe,N–Co(OH)x (1
:
1) revealed five N species: pyridinic N (398.5 eV), Co–N or Co/Fe–N (399.7 eV),28 pyrrolic-N (400.9 eV), graphitic N (402.8 eV), and NO3−(406.7 eV) (Fig. S11c, ESI†). Raman spectra (Fig. S13 and Note 3, ESI†) have been verified and revealed the origin of the metal–N bond. The N–Co–N deformation vibration was clearly observed in the region of 120–195 cm−1. And the C–CH3 stretching peaks of the 2-MI linker were visible at approximately 674 cm−1, indicating that the Co–N bond in Fe,N–Co(OH)x originates from 2-MI.51 Pyridinic and graphitic N, part of the N-doped C, exhibited improved wettability and a strong affinity for water, thereby aiding electron transfer during the OER.52
X-ray absorption spectroscopy (XAS) was conducted to analyze the coordination number and multi-defect structures in Fe,N–Co(OH)x (1:
1) (Fig. 2a–c and Fig. S14, Table S6, ESI†). A detailed analysis of the pre-peak and main peak in the Co K-edge X-ray absorption near-edge structure (XANES) spectra of Fe,N–Co(OH)x (1
:
1) was carried out to elucidate the oxidation states, coordination environment and electronic interactions induced by the Fe and N co-doping. The Co K-edge X-ray XANES spectra (Fig. 2a) exhibit the energy absorption edge between CoO and Co3O4, confirming a mixed valence state of Co2+ and Co3+. The pre-edge peak around 7710 eV corresponds to the Co 1s-3d transition,53 and a higher pre-edge peak represents a lower centrosymmetry of the octahedron in Co(OH)2. The pre-edge peak intensity of Fe,N–Co(OH)x (1
:
1) is higher than Co(OH)2, which suggests a lower centrosymmetry of Fe,N–Co(OH)x (1
:
1) after Fe and N co-doping.54 Compared with Co(OH)2, the white line (7725 to 7730 eV) of Fe,N–Co(OH)x (1
:
1) sample is broader and lower intensity, indicating strong electronic interaction after Fe and N co-doping.55 The Fourier transformed (FT) k3χ(k) spectra in R space (Fig. 2b) revealed three scattering paths around absorbing Co ions. The first peak at 1.53 Å corresponded to the Co–O and Co–N first coordination shell.27 Owing to limited resolution and low N content, differentiating between Co–N and Co–O was challenging.27 The second peak at 2.5 Å indicated the Co–O–Co(Fe) second coordination shell.56 Moreover, the coordination bond length and peak intensity of Fe,N–Co(OH)x (1
:
1) are significantly lower than those of Co(OH)2, which indicates that Fe and N co-doping resulted in the formation of multiple defects.14 According to the (FT) k3χ(k) fitting results and parameters (Fig. 2c and Table S6, ESI†), the ex situ Raman experimental (Fig. S13 and Note 3, ESI†) and fitting data showed high consistency, revealing that the Co–N and Co–O coordination numbers (CN) are 2 and 1.5, respectively. Compared to the 6-coordinated metal–O structure in saturated-coordinated Co(OH)2, the first-shell coordination numbers of Fe,N–Co(OH)x (1
:
1) sum up to 3.5, which is lower than 6. This indicated the presence of abundant Ov,41,57 which was consistent with XPS and EPR findings. The EXAFS wavelet transform analysis revealed an intensity maximum at 3.5 Å−1 owing to Co–N/O scattering (Fig. S14, ESI†). These results indicated that Fe,N co-doping altered the coordination environment in Fe,N–Co(OH)x (1
:
1), leading to multiple defects (Fe,N co-doping and Ov).
XAS and XPS data revealed that introducing metallic Fe and nonmetallic N substantially modified the valence states, coordination environments, and electronic structure of the Co atoms. We proposed the Co, CoN, and FeCoN site structures, as illustrated in Fig. 2d. The lattice oxygen between Co and Fe also affected adsorption–desorption properties (Fig. 2d and Fig. S16–S18, ESI†). Typically, Oh-symmetric metal atoms interacted with bridging oxygen (O2−) through π-donation in hydroxide structures, thereby forming M–O–M moieties.58 In this study, we constructed an N–Co–O–Fe moiety that bridged Co and Fe via O2− to examine partial electron transfer (PET) at the interface and electronic coupling between adjacent Co and Fe (Fig. 2e). The d-orbital configurations of Co2+ (t62ge1g), Co3+ (t62ge0g), and Fe3+ (t32ge2g) are depicted in Fig. S15 (ESI†). The t2g-orbitals of Co2+ (t62ge1g) and Co3+ (t62ge0g) had no unpaired electrons, and their p-orbitals were fully occupied by O2−. This resulted in electron repulsion that hindered the charge transfer (Fig. 2f, top).56,59 Projected density of states (PDOS) calculations indicated substantial overlap between Co t2g-orbitals and O 2p-orbitals near the Fermi level (−0.5–2 eV), thereby confirming electron transfer from the t2g-orbitals (dxy, dyz, dxz) to bridging O2− (Fig. S16a, top, ESI†).12 The varying electronic occupancy of Co and Fe d-orbitals can trigger PET through bridging O2− p-orbitals (Fig. 2f, middle),56,57,60 resulting in double interaction (DEI) between Co and Fe.61,62 N facilitated electron transport by altering the electronic density at adjacent Co sites.63 Introducing N at the CoN site substantially reduced the electron density in Co t2g-orbitals and O 2p orbitals near the Fermi level (−0.5–2 eV) (Fig. S16a, middle, ESI†), while co-doping with Fe and N markedly increased it (Fig. S16a, bottom, ESI†). This highlighted the electron-withdrawing properties of N, which decreased the electron density at the adjacent Co sites. With Fe and N coordinated to Co2+/Co3+, the orbitals contained more electrons, facilitating electron transfer between Co t2g and O 2p-orbitals. This enabled easier electron transfer from Co2+/Co3+ to Fe3+via bridging O2− (Fig. 2f, bottom). This process facilitated electron redistribution.59 Benefiting from the electronic modulation of the N–Co–O–Fe moiety, the Fe sites optimize *OH adsorption, while the Co sites are more conducive to O2 desorption, thus Fe,N–Co(OH)x is more favorable for starting an OER cycle and oxygen release (Fig. S17, S18 and Note 4, ESI†). The mechanism of the effects of Co and Fe on the absorption of oxygen-containing species will be analyzed in the DFT calculation section. Hence, the intrinsic OER activity of Fe,N–Co(OH)x could be enhanced through PET within the N–Co–O–Fe moiety.
Moreover, PDOS analysis validated the variation in the electronic structure of the lattice oxygen based on comparisons of the Co, CoN, and FeCoN sites within the N–Co–O–Fe model (Fig. 2d and g). As shown in Fig. 2g, the moderate d-band center of Co (−4.17 eV) at the FeCoN site, compared to the Co (−3.82 eV) and CoN (−4.63 eV) sites, indicated that Fe,N–Co(OH)x (1:
1) possessed optimal adsorption and desorption energies for OER intermediates. Fe,N co-doping shifted the metal d-band center downward from −3.82 to −4.17 eV, reducing the charge transfer (Δ) between εd and the O 2p-band center (εO
2P) from −0.07 to −0.03 eV.64 This shift facilitated the penetration of the O 2p-band and the formation of ligand holes, which favored the release of lattice oxygen (Fig. 2h). This confirmed the occurrence of lattice oxygen oxidation and enabled the lattice oxygen to function as active sites.20,64–66 For the FeCoN site, the overlap of the Co 3d, Fe 3d, and lattice O 2p orbitals near the Fermi level (shaded in purple) was substantially increased compared to the Co and CoN sites (Fig. 2g, lower panel). This confirmed a strong interaction between Co and Fe atoms and covalent hybridization with oxygen ligands. In addition, the continuous band structure of Fe,N–Co(OH)x (1
:
1) around the Fermi level supported high electrical conductivity and enhanced electron transfer. This synergy between Fe and N atoms strengthened the covalency of the metal–O bond, thereby promoting lattice oxygen activation during the OER.8,12,67 The hybridization of metal 3d and O 2p orbitals improved electron transfer efficiency during OER processes,68 which facilitated the LOM pathway.
To further understand the mechanism governing the high OER activity of Fe,N–Co(OH)x (1:
1), impedance spectroscopy (EIS) and intrinsic activity assessments were conducted. The EIS results (Nyquist plots) indicated that Fe,N–Co(OH)x (1
:
1) exhibited the lowest resistance (2.26 Ω) and efficient proton-coupled electron transfer capability (Fig. S22 and Table S7, ESI†). To elucidate the charge transfer processes at the catalytic interface, the operando EIS test was conducted at various applied biases to track dynamic evolution. The corresponding Bode phase plots illustrate the variation of phase angles with frequency (Fig. 3d and Fig. S23, ESI†). When the voltage was increased from 1.3 V to 1.6 V, the phase angle of Fe,N–Co(OH)x (1
:
1) decreased significantly in the low and high frequency regions of Bode phase plots (Fig. 3d), indicating rapid electron transport at the electrolyte/electrode interfaces and implying an accelerated interfacial reaction during the electrocatalytic OER process.69 However, the frequency peak decreases more slowly with increasing bias for W,N–Co(OH)x (1
:
1) (Fig. S23, ESI†). From the above results, it is confirmed that Fe-doping can promote the electron transfer ability and accelerate surface remodeling at the electrode interface, thus promoting the OER process. The intrinsic OER activity was further evaluated using ECSAs and loading mass to identify active sites. Fe,N–Co(OH)x (1
:
1) exhibited an ECSA of 110 cm2, which was larger than that of other samples (Fig. S24, ESI†). The mass activities and turnover frequencies (TOF) at an overpotential of 350 mV were also estimated. Fe,N–Co(OH)x (1
:
1) exhibited the highest mass activity (1705 A gmetal−1) and TOF (2.521
s−1), which were 80.4 and 57 times greater than those of W,N–Co(OH)x (1
:
1) (21.2 A gmetal−1 and 0.044
s−1), respectively (Fig. S25 and Table S7, ESI†). Radar plots visually compare the performance of these catalysts. They clearly showed the superior intrinsic activity of Fe,N–Co(OH)x (1
:
1) (Fig. 3e), confirming that the incorporation of Fe and N enhanced the intrinsic activity of Co(OH)2. A summary of the electrochemical test results is provided in Table S7 (ESI†).
The long-time operational stability of Fe,N–Co(OH)x (1:
1) was assessed through chronopotentiometry measurements. The operating potential remained stable, with real-time potentials demonstrating minimal change following 50 h of oxygen evolution (Fig. S26, ESI†). The catalyst durability was further demonstrated at higher current densities of 10, 50, and 100 mA cm−2 (Fig. 3f). Stability results showed that Fe,N–Co(OH)x (1
:
1) requires only ∼1.5 V to drive a current density of 10 mA cm−2, whereas W,N–Co(OH)x (1
:
1) requires ∼1.65 V (Fig. S27, ESI†), indicating that Fe,N-codoping Co(OH)x can achieve long-term OER reactions at lower voltages. These results confirmed that Fe,N–Co(OH)x (1
:
1) exhibited excellent stability in alkaline solutions. Fig. 3j presents a comparison of the OER performance of Fe,N–Co(OH)x (1
:
1) with previously reported catalysts, highlighting its superior performance.
Post-characterization was performed following 50 h of OER using chronoamperometry. The XRD patterns before and after the stability test revealed no significant changes in the structure of the post-OER Fe,N–Co(OH)x (1:
1) (Fig. S30, ESI†). ICP-MS measurements proved the Co and Fe contents in the post-OER Fe,N–Co(OH)x (1
:
1) were 40.1 wt% and 10.2 wt%, respectively (Table S2, ESI†). Furthermore, we tested the metal content in the electrolyte after the OER. The concentrations of Fe and Co in the electrolyte after the OER were 0.00024 mg L−1 and 0.00137 mg L−1, respectively, indicating that only <0.03% of the electrode metals were dissolved in the electrolyte. Thus, XRD and ICP analysis showed that Fe,N–Co(OH)x (1
:
1) maintained good structural stability and nearly no dissolution during the electrocatalytic tests. XPS analysis of post-OER Fe,N–Co(OH)x (1
:
1) is shown in Fig. S31 (ESI†). The Co 2p spectrum following the 50-hour durability test exhibited a negative shift of 0.4 eV, indicating an increased valence state (Fig. S31a, ESI†).70 The Co 2p spectrum was deconvoluted into four prominent peaks corresponding to Co3+ (780.1 and 795.0 eV) and Co2+ (781.6 and 796.6 eV), along with four satellite peaks. The relative percentages of Fe2+ (60.8%) and Fe3+ (39.2%) in post-OER Fe,N–Co(OH)x (1
:
1) were almost unchanged compared to the original Fe,N–Co(OH)x (1
:
1) (Table S4, ESI†). Minimal change in the Fe 2p XPS spectra signified that Fe remained intact without dissolution following stability testing (Fig. S31b and Table S5, ESI†). The C 1s spectra showed peaks at 284.8, 288.5, 291.7, 292.8, and 295.4 eV, which were attributed to C
C, C
O, π
π, and C–Fx from Nafion,71 respectively (Fig. S31e, ESI†). The proportion of the Co–N bond decreased from 20.78% to 11.65%, indicating a partial transformation from Co–N to Co–OH during surface reconstruction (Fig. S31c, d and Table S5, ESI†). Furthermore, the O 1s spectra revealed that the material retained a high proportion of Ov (27.4%) following 50 h of OER, indicating low Ov formation energy (Fig. S31d, ESI†). This indicated that lattice oxygen readily desorbed oxygen species to produce Ov and release O2. Post-characterization confirmed that Fe,N co-doping effectively reduced Ov formation energy and synergistically activated the LOM pathway.
The origin of the O–O bond configuration in the OO* intermediate for Fe,N–Co(OH)x (1:
1) during the OER process was demonstrated through in situ18O isotope-labeled DEMS experiments, as shown in Fig. 4c. After Fe,N–Co(OH)x (1
:
1) was labeled with the isotope 18O in 0.1 M K18OH (Step 2: lattice oxygen exchange, shown in Fig. 5a), three types of O2 molecules were monitored by DEMS when tested in the 0.1 M K16OH electrolyte: 16O16O, 16O18O, and 18O18O. The pronounced, high-intensity periodic signal of 16O18O (m/z = 34) provided strong evidence that the O–O bond configuration in Fe,N–Co(OH)x (1
:
1) originated from lattice oxygen during the OER process, thereby confirming that Fe,N–Co(OH)x (1
:
1) followed the LOM pathway.74 Therefore, the in situ ATR-FTIR measurements and in situ18O isotope-labeled DEMS collectively provided compelling evidence that Fe,N co-doping activated lattice oxygen, with the LOM reaction mechanism being the primary pathway.
Subsequently, electrochemical in situ Raman spectroscopy was conducted to obtain deeper insight into the active structures of Fe,N–Co(OH)x (1:
1) and W,N–Co(OH)x (1
:
1) during the OER (Fig. 4d and e). For Fe,N–Co(OH)x (1
:
1), two spectral features of Co(OH)2 were observed at approximately 480 and 688 cm−1 at open circuit potential, corresponding to the Eg vibration modes of Co(OH)2 (Fig. 4d).75 As the voltage increased to 1.3 V, the A1g vibration mode gradually disappeared, while new broadband emerged at 580 cm−1, which was attributed to the Eg mode of the Co(III)–O band in CoOOH.76 This peak narrowed and red-shifted to approximately 550 cm−1 as the voltage increased from 1.4 to 1.8 V, indicating a new phase associated with the A1g vibration mode of the Co(IV)–O band in CoO2.18,77 This shift indicated that the strong PET from the N–Co–O–Fe moiety at the Fe,N–Co(OH)x (1
:
1) interface promoted the formation of high-valent Co(IV) species and involved active lattice oxygen in the OER. Moreover, no peaks related to Fe species were detected, indicating that Fe site was unlikely to be the active center in the Fe,N–Co(OH)x (1
:
1) catalyst. In contrast, W,N–Co(OH)x (1
:
1) exhibited CoOOH peaks at 580 cm−1 only when the applied potential ranged as 1.5–1.8 V (Fig. 4e). Notably, Fe,N–Co(OH)x (1
:
1) underwent a phase transition at a considerably lower potential (1.30 V), indicating that Fe,N co-doping accelerated the formation of CoOOH and CoO2 species. Thus, in situ Raman revealed that the surface of the Fe,N–Co(OH)x (1
:
1) electrocatalyst reconstructed to CoOOH and CoO2 with increasing voltage, with high-valent Co(IV) acting as the true active center.
To rationally design Fe and N co-doped defects, we analyzed the impact of Fe and N locations on OER activity. Three Fe locations (Fig. S32, ESI†) were examined. The Fe1 site corresponded to the previously discussed FeCoN sites, which exhibited an overpotential of 0.81 eV. The rate-determining steps for Fe2 site and Fe3 site were not shifted from OO* desorption to other reaction steps, with η values of 2.15 and 2.53 eV, respectively (Fig. S33, ESI†). The OO* desorption energies for the Fe1 site (0.43 eV) and the Fe2 site (0.92 eV) were substantially lower than that for the Fe3 site (1.30 eV). The key distinction of the Fe3 site was that N was bonded to Fe rather than Co. Therefore, for optimal activity with Fe and N co-doping, N must form as many Co–N bonds as possible. In our synthesis strategy, Fe was incorporated through Fe electrode sputtering, whereas N was pre-coordinated with Co in the solution. This approach minimized the formation of Fe–N bonds, thereby ensuring that Co–N bonds were the dominant metal–N coordination. The impact of N locations on the OER pathway was also analyzed (Fig. S34–S37, ESI†). The CoN2 configuration exhibited a higher OO* desorption energy (1.63 eV) compared to the CoN1 site (1.25 eV) and the FeCoN site (0.43 eV). This analysis confirmed that Fe and N synergistically reduced OO* desorption energy, promoted Ov formation and activated the LOM pathway. For designing multi-structures, the prioritization of Co–N coordination while minimizing Fe–N coordination is essential to construct lattice oxygen-participating in the N–Co–O–Fe active site.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ey00280f |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |