Oxygen reduction reaction performance of Fe-N-C catalyst with dual nitrogen source

Yuan Zhao, Quan Wang, Rongrong Hu, Wenqiang Liu, Xiaojuan Zhang, Wei Wang, Nicolas Alonso-Vante, Dongdong Zhu

Front. Energy ›› 2024, Vol. 18 ›› Issue (6) : 841-849.

PDF(3395 KB)
Front. Energy All Journals
PDF(3395 KB)
Front. Energy ›› 2024, Vol. 18 ›› Issue (6) : 841-849. DOI: 10.1007/s11708-024-0956-2
RESEARCH ARTICLE

Oxygen reduction reaction performance of Fe-N-C catalyst with dual nitrogen source

Author information +
History +

Abstract

Fe-N-C catalysts are potential substitutes to displace electrocatalysts containing noble chemical elements in the oxygen reduction reaction (ORR). However, their application is hampered by unsatisfactory activity and stability issues. The structures and morphologies of Fe-N-C catalysts have been found to be crucial for the number of active sites and local bonding structures. In this work, dicyandiamide (DCDA) and polyaniline (PANI) are shown to act as dual nitrogen sources to tune the morphology and structure of the catalyst and facilitate the ORR process. The dual nitrogen sources not only increase the amount of nitrogen doping atoms in the electrocatalytic Fe-C-N material, but also maintain a high nitrogen-pyrrole/nitrogen-graphitic: (N-P)/(N-G) value, improving the distribution density of catalytic active sites in the material. With a high surface area and amount of N-doping, the Fe-N-C catalyst developed can achieve an improved half-wave potential of 0.886 V (vs. RHE) in alkaline medium, and a better stability and methanol resistance than commercial Pt/C catalyst.

Graphical abstract

Keywords

Fe-N-C / oxygen reduction reaction / nitrogen-doped / dual nitrogen source

Cite this article

Download citation ▾
Yuan Zhao, Quan Wang, Rongrong Hu, Wenqiang Liu, Xiaojuan Zhang, Wei Wang, Nicolas Alonso-Vante, Dongdong Zhu. Oxygen reduction reaction performance of Fe-N-C catalyst with dual nitrogen source. Front. Energy, 2024, 18(6): 841‒849 https://doi.org/10.1007/s11708-024-0956-2

1 Introduction

The design and development of high-activity and low-cost catalysts can effectively improve the slow kinetics of the oxygen reduction reaction (ORR), which is a fundamental reaction for some related technologies, such as fuel cells and metal-air batteries [1,2]. Platinum (Pt) and its alloys are considered to be the most potential oxygen reduction catalysts due to their high activity. However, the high cost and poor stability of Pt seriously hinder its large-scale application in commercial fuel cells. In addition, Pt has other drawbacks, such as the poor durability and the ease with which it can be poisoned [35]. Currently, transition metal supported on nitrogen-doped carbon (M-N-C) materials have been extensively investigated [68], and are considered more promising candidates that can replace Pt or other precious metal-based ORR catalysts [9,10]. Compared to noble metal-based ORR electrocatalysts, M-N-C catalysts generally have some advantages, such as low raw materials cost, simple synthesis route, and excellent ORR performance [1113]. The conventional method for preparation of these materials consists of pyrolyzing the mixed precursor containing metallic elements (M = Fe, Co, etc.), carbon elements, and doping with nitrogen atoms [1417].
It is generally accepted that not only the amount and type of doped-N atoms but also the porosity, surface area and oxygen transport resistance of M-N-C can also affect the ORR performance [18,19]. The morphology and structure of M-N-C catalysts can be tuned by choosing the appropriate carbon source and nitrogen source to achieve an excellent oxygen reduction performance [2022]. In previous studies, polyaniline (PANI) [23,24], dicyandiamide (DCDA) [25,26], and melamine (MLMN) [2729] have been the most commonly used nitrogen sources. DCDA is an aminonitrile dimer, PANI is an aniline polymer, and MLMN contains three amino groups, while some C atoms in the benzene ring are also replaced by N atoms. There exist conjugated structures in PANI, DCDA, and MLMN molecules, and the delocalized π electrons they provided can facilitate the N atoms doped carbon skeleton during pyrolysis [30]. In addition, DCDA and MLMN have high ratio of nitrogen to carbon, PANI is a polymer with good mechanical properties, and the three-dimensional structure is conducive to improve the distribution density of catalytically active sites [31]. Although PANI, DCDA, or MLMN alone have previously been used as nitrogen precursors for M-N-C synthesis, in-depth investigation of M-N-C catalysts prepared by choosing dual nitrogen sources has been rarely reported.
In this work, Fe-PD/PGO (Fe-PANI/DCDA/graphene oxide modified by p-phenylenediamine) nanosheets were synthesized in ice-water bath using PGO as carbon carrier, PANI, and DCDA as dual nitrogen sources, FeCl3·6H2O as single metal source, and Fe-PD/PGO catalytic materials were prepared by pyrolysis. In addition, the effects of different PANI/DCDA addition ratios, metal content, and pyrolysis temperature on the ORR activity of Fe-PD/PGO catalysts were analyzed. The result shows that the structural defects in the Fe-PD/PGO catalytic increase, and the degree of graphitization decreases, which further increase the content of N atoms, thus improving the distribution density of catalytic active sites in the material.

2 Experimental

2.1 Catalysts preparation

2.1.1 Synthesis of graphene oxide modified by p-phenylenediamine (PGO)

Graphite oxide (GO) was prepared by utilizing the modified Hummers method, and through an amide reaction to synthesis the p-phenylenediamine (PPD) modified GO, denoted as PGO. First, 1.0 g of the GO prepared was transferred to 1.0 L ultrapure water and dispersed for 0.5 h under ultrasonic conditions. Then 16 mmol of PPD and 16 mmol of H2SO4 were added to the GO dispersion subsequently. Meanwhile, 16 mmol of NaNO2 was slowly dissolved in 10 mL of ultrapure water, and the reaction was carried out at 60 °C for 4.0 h. NaNO2 was used to stabilize acidic conditions and under GO an azo reaction with aniline, which favors the reaction of aniline with graphene oxide to form PGO. After centrifuged the suspension, it was washed with ethanol, and then freeze-dried to obtain PGO for storage. Before use, the PGO was immersed in 0.1 mol/L of hydrochloric acid for two hours to remove any metal impurities that may be present in the PGO.

2.1.2 Synthesis of Fe-PD/PGO

0.2 g of acid-soaked PGO was added to 0.2 L of hydrochloric acid solution (0.5 mol/L) and dispersed in an ultrasonic stirrer for 1 h to make it homogeneous. Then, 1 mL aniline and a certain amount of DCDA were added to the FeCl3·6H2O mixture which was put into an ice water bath to bring the temperature of the reaction system below 5 °C. Thereafter, 0.5 mol/L of HCl was used as the solvent and ammonium persulfate as the solute to prepare a solution (0.2 g/mL) which was slowly dripped into the mixture system using a constant-speed dropping funnel. The mixture was stirred continuously for some time to allow the aniline to polymerize completely. Afterwards, solid-liquid separation was conducted by rotary evaporation, and then dried again. The sample dried was heated at 5 °C/min to 900 °C for 1 h under nitrogen conditions with a flow rate of 0.2 L/min. Finally, the product was immersed in 0.5 mol/L H2SO4 and heated at 80 °C for 12 h to remove unstable materials. The precursors were then washed and dried several times by filtration and deionized water, and heated for the second time at 5 °C/min to 900 °C for 3 h, with a nitrogen flow rate of 0.2 L/min. After grinding, the final product was Fe-PD/PGO. Fe-DCDA/PGO and Fe-PANI/PGO were synthesized by adding only DCDA or aniline.

3 Results and discussion

The structure and degree of crystallization of Fe-DCDA/PGO, Fe-PANI/PGO, and Fe-PD/PGO samples were first tested by X-ray diffraction (XRD). In Fig.1(a), the Fe-DCDA/PGO, Fe-PANI/PGO, and Fe-PD/PGO samples show characteristic peaks located around 26.1°, and 42.9°, well assigned to the (002) and (101) planes of carbon, respectively [32]. The height of the peak can represent the perfection of the crystal type, the more perfect the grain arrangement, the higher the peak. Among them, the (002) peak of the Fe-DCDA/PGO sample is the sharpest, indicating that the carbon materials have a higher degree of graphitization, while the broader (002) peak of Fe-PD/PGO indicates a decrease in its degree of graphitization [33]. Furthermore, the diffraction peak of the Fe-DCDA/PGO sample centered at 29.5° belongs to the (111) plane of Fe3C (JCPDS: 035-0772) [21,34], while the peaks of Fe-PANI/PGO and Fe-PD/PGO samples at 35.2° are attributed to the (200) planes of Fe3C, indicating that there is some Fe element in the form of Fe3C in the samples obtained. The Fe/Fe3C nanocrystals boost the activity of Fe-Nx, which is essential for the high ORR performance [35,36]. In addition, Raman spectroscopy analysis was used to characterize the carbon structure of the samples (Fig.1(b)). The ID/IG value of Fe-PD/PGO is 1.02, which is significantly higher than those of Fe-PANI/PGO (0.91), Fe-DCDA/PGO (0.84), and PGO (0.85), (cf., Fig. S1 in Electronic Supplementary Material), indicating that the Fe-PD/PGO sample has the highest amount of carbon structural defects, the highest disorder, and the lowest degree of graphitization among the three samples, which is in agreement with previous XRD results [37]. Therefore, it can be concluded that by choosing the right nitrogen sources, the product obtained can have a low degree of graphitization, along with a large number of carbon structural defects, which are highly desirable to achieve an excellent ORR performance [38].
Fig.1 (a) XRD patterns and (b) Raman spectrums of Fe-DCDA/PGO, Fe-PANI/PGO, and Fe-PD/PGO samples.

Full size|PPT slide

A comparison of Figs. S3 and S2 indicates that the pyrolyzed catalyst has a stacked lamellar structure. The three-dimensional porous structure in Fe-PD/PGO materials is more pronounced, and the layers are interconnected through porous carbon structures to form a network structure. As shown in Fig.2, the Fe-DCDA/PGO material exhibits a silk-like microstructure, and the Fe-PANI/PGO catalytic material contains some fibrous structures. The morphology of Fe-PD/PGO differs significantly from that of Fe-DCDA/PGO and Fe-PANI/PGO materials. The silk-like structure is no longer obvious, and many fibrous structures appear, which is consistent with the three-dimensional network structure observed in SEM.
Fig.2 Transmission electron microscope (TEM) of catalysts.

Full size|PPT slide

As illustrated in Fig.3(a) and Fig.3(b), a typical N2 adsorption–desorption isotherms type-IV with a hysteresis loop can be observed in the section of relatively high pressures, suggesting that all catalysts contain the mesoporous structure. In the case of Fe-PD/PGO, the volume adsorbed demonstrates a rapid increase at a relatively low pressure (P/P0 = 0–0.01), indicating the existence of abundant micropores. The pore size distribution of Fe-PD/PGO catalytic materials is concentrated below 2.0 nm. The surface areas of Fe-DCDA/PGO, Fe-PANI/PGO, and Fe-PD/PGO samples were calculated to be 305.4, 359.8, and 639.1 m2/g, respectively. Moreover, the pore volumes were 0.33, 0.48, and 0.86 cm3/g, respectively, indicating that the use of DCDA and PANI as dual nitrogen sources can effectively enlarge the specific surface area and cumulative pore volume. In fact, the large specific surface area and the high proportion of micropores in the catalyst favor the exposure of more catalytic active sites, and improve the probability of their contact with O2, thus improving the ORR performance of the materials [39,40]. In addition, voltametric tests were performed on Fe-DCDA/PGO, Fe-PANI/PGO, and Fe-PD/PGO samples in 0.1 mol/L of KOH to measure the electrochemical active surface area [41]. Compared with the specific surface area of the catalytic material itself, its corresponding electrochemically active surface area (EASA) is lower, indicating that the specific surface area possessed by the catalytic material cannot be fully utilized in the electrochemical reactions. Therefore, using EASA (converted from cm2 to m2/g by dividing the EASA area by the catalysts mass loading) to study the ORR activity of catalytic materials is straightforward. The EASA of Fe-DCDA/PGO, Fe-PANI/PGO, and Fe-PD/PGO catalytic materials are 191.4, 283.4, and 317.5 m2/g, respectively (Fig.3(c)). Combined with the ORR performance tests, it is found that the EASA size and the ORR activity level have a similar trend, i.e., the ORR activity of the catalyst increases with the increase of EASA [42].
Fig.3 BET and EASA of catalysts.

Full size|PPT slide

The high-resolution XPS spectra of N 1s for Fe-PANI/PGO, Fe-DCDA/PGO, and Fe-PD/PGO were displayed in Fig.4. The N 1s can be roughly divided into four peaks with binding energies centered at 398.7 ± 0.2, 400.1 ± 0.2, 401.5 ± 0.3, and ≈ 403.5 eV, corresponding to pyridinic N (N-P), pyrrolic N (N-Pr), graphitic N (N-G), and oxidized pyridinic N (N-OP) [28]. The content of different types of N is shown in Tab.1. Compared with Fe-DCDA/PGO and Fe-PANI/PGO, the N atom content in Fe-PD/PGO catalytic materials is further increased, making them have a higher N-P/N-G value (0.84). This may be due to the synergistic promotion of doping with the N atom by the presence of DCDA and PANI. Meanwhile, PANI can form a structurally stable six-membered ring structure containing two N atoms under the action of oxidants during the polymerization process, effectively promoting the formation of N-P and N-G, among which N-P is believed to favor the performance of the catalysts [12,30,43]. From Fig. S4, it can be seen, that Fe-DCDA/PGO and Fe-PANI/PGO materials have emerging emission peaks at 724.3 and 710.9 eV, with the Fe 2p3/2 peak, in the high-energy region (724.3 eV), caused by Fe (III); and the Fe 2p1/2 peak, in the low-energy region (710.9 eV), caused by Fe2+ or Fe3+ bound to N atoms [44]. The Fe-Nx structure is widely recognized as an active site with a catalytic ability [45,46]. The synergy interaction at the Fe-Nx nanoparticles/N-C interface, certainly redistribute the electron density of the active sites [47,48].
Tab.1 Nitrogen content of different types in different catalysts
Catalysts Nitrogen type composition/% Total N/% N-P/N-G
N-OP N-G N-Pr N-P
Fe-DCDA/PGO 12.23 64.92 5.53 17.32 1.55 0.27
Fe-PANI/PGO 6.81 44.98 8.30 39.91 2.99 0.89
Fe-PD/PGO 9.34 44.23 9.24 37.19 3.29 0.84
Fig.4 XPS high-resolution N 1s spectra.

Full size|PPT slide

As shown in Fig.5(a), there is no obvious redox peak in the cyclic voltammetry (CV) curves of the different catalysts in the N2-saturated system. However, when tested in the O2-saturated system, the Fe-DCDA/PGO catalyst produces a reduction peak at 0.795 V vs. reversible hydrogen electrode (RHE). Combined with CV tests under N2 saturated conditions, this peak can be determined to be the O2 reduction peak. Compared with Fe-DCDA/PGO materials, the reduction peak potentials of Fe-PANI/PGO, Fe-PD/PGO, and Pt/C were shifted positively by 20, 41, and 39 mV, respectively, indicating that Fe-PD/PGO has a comparable ORR performance. The onset potentials (Eonset) were determined at the current density of 0.1 mA/cm2. The Eonset, E1/2, and Jk@0.9V of Fe-PD/PGO materials reaches 0.951 V (vs. RHE), 0.886 V (vs. RHE), and 3.178 mA/cm2, respectively, much higher than those of Fe-DCDA/PGO and Fe-PANI/PGO materials, and even exceeded the Eonset (0.932 V vs. RHE) and E1/2 (0.849 V vs. RHE) of Pt/C catalysts (Fig.5(b) and Fig.5(d)). The corresponding Tafel slope was calculated based on the LSV curve mentioned above, and the specific slope is shown in Table S1. The Tafel slope calculated for Fe-PD/PGO is 76.8 mV/dec1, which is lower than that of other catalysts, indicating relatively faster ORR kinetics with Fe-PD/PGO [24].
Fig.5 Electrochemical performance of catalysts.

Full size|PPT slide

At the same time, to study the changes in the number of transferred electrons (n) of the catalytic materials mentioned above during the ORR process, LSV tests were performed on Fe-PANI/PGO, Fe-PD/PGO, and Pt/C materials at different rotation rates under the same testing conditions, and the corresponding Koutecky–Levich (K–L) equations were calculated (Fig. S5). The number of transferred electrons of the Fe-PD/PGO catalytic material is 3.98, which is higher than that of Fe-PANI/PGO (3.62), and is equivalent to the number of transferred electrons of Pt/C catalyst (3.98), indicating that the Fe-PD/PGO catalytic material has a high electron transfer efficiency in the ORR process, which agrees with the conclusion obtained by Tafel analysis.
The effect of different PD molar ratios on the ORR performance of Fe-PD/PGO catalytic materials were also analyzed. Figure S6 shows the CVs, linear sweep voltammetries (LSVs), Tafel, and performance parameter plots of Fe-PD (1:1/2:3/1:2/2:5/1:3)/PGO catalysts, respectively. According to the CV test in Fig. S6(a), there is a significant oxygen reduction peak at 0.786 V (vs. RHE) for Fe-PD (1:1)/PGO in an O2-saturated electrolyte.
Compared with the Fe-PD (1:1)/PGO catalyst, the peak potentials of Fe-PD (2:3/1:2/2:5/1:3)/PGO were positively shifted by 30, −17, 37, and 34 mV, respectively, indicating that the Fe-PD (2:5)/PGO catalytic material has the best ORR activity under alkaline conditions. To accurately evaluate the ORR performance of the different catalytic materials, further LSV tests were performed. As shown in Figs. S6(b) and S6(d), the Fe-PD (2:5)/PGO catalytic material shows a significant improvement in Eonset, E1/2, and Jk@0.9 V, with Eonset and E1/2 reaching 0.951 V (vs. RHE) and 0.886 V (vs. RHE), respectively. Therefore, when the PANI/DCDA molar ratio is 2:5, the Fe-PD/PGO catalytic material prepared has the best ORR activity and relatively faster ORR kinetics (Table S2).
The ORR activity of Fe-PD/PGO catalytic material first increases and then decreases with increasing PD ratio, which is mainly caused by the fact that the addition of DCDA can synergize with PANI, which increases the defects in the catalytic material during pyrolysis, and then allows more N atoms to be incorporated into the carbon structure, improving the distribution density of the catalytic active site. However, an excess of DCDA can significantly reduce the PANI content in the precursor, and PANI can form a stable six-membered ring structure containing two N atoms during the polymerization process, which can effectively promote N-P and N-G formation. Therefore, an excess of DCDA can weaken the ORR performance of the catalytic material.
The durability of the catalysts was evaluated, as shown in Fig.6(a) and Fig.6(b). After 5000 and 10000 CV cycles, the Eonset of the LSV curve of the Fe-PD/PGO catalytic material measured had almost no attenuation, while E1/2 only decreased by 6 and 7 mV, respectively; under the same testing conditions, after 10000 cycles of CV cycles, the Eonset and E1/2 of the LSV curves of the Pt/C catalyst measured decreased by 19 and 25 mV, respectively. Additionally, the durability of Fe-PD/PGO was also tested by chronoamperometry. The current attenuation of the catalyst is very slow, and the current density still maintains 95.6% after 30000 s (Fig.6(c)). Therefore, the Fe-PD/PGO material shows a good electrochemical stability performance.
Fig.6 Durability and methanol resistance of catalysts.

Full size|PPT slide

In addition, the catalyst performance in terms of resistance to methanol tolerance was tested by chronoamperometry, j(t), at 0.7 V (vs. RHE in an O2-saturated electrolyte, without rotation) CH3OH (3.0 mol/L) was added at 100 s to evaluate the methanol tolerance resistance. As illustrated in Fig.6 (d), compared with Fe-PD/PGO, Pt/C catalyst shows a higher current density at the initial stage of the reaction. However, with the addition of CH3OH, its current density sharply decreases and cannot return to its original state. The main reason for this is that CH3OH undergoes an oxidation reaction on the catalyst surface, and the oxidation current generated interacts with the ORR reduction current, resulting in a mixed electrode potential formation leading to a decrease. The current densities of Fe-PD/PGO changes only slightly after experiencing brief fluctuations, indicating that under the same conditions, Fe-PD/PGO presents a better methanol tolerance than Pt/C catalysts.

4 Conclusions

In this work, the Fe-PD/PGO catalyst was prepared by two-step pyrolysis using PGO as carbon carrier, DCDA and PANI as dual nitrogen sources, and FeCl3·6H2O as a single metal source. In material synthesis, PANI preferentially generates a coarser carbon planar structure with a relatively lower degree of graphitization, a higher content of nitrogen and iron, and dominant micropores. On the other hand, the graphite produced by DCDA has a finer structure, a higher degree of graphitization, a lower nitrogen and iron content, and a higher mesoporous/microporous ratio. Therefore, the dual nitrogen sources not only enhance the doping amount of N atoms in the Fe-PD/PGO catalytic material, but also maintain a high N-P/N-G value (0.84), favorable in increasing the distribution density of catalytic active sites in the material. In addition, the high specific surface area of Fe-PD/PGO and EASA can fully expose active sites on the surface of the material, which improves the contact probability with O2 molecules. The Eonset, E1/2 and, Jk@0.9V of Fe-PD/PGO reach 0.951 V (vs. RHE), 0.886 V (vs. RHE), and 3.178 mA/cm, respectively. Compared with Pt/C, Eonset and E1/2 have a positive shift of 19 and 37 mV, respectively, and the electron transfer number (n) of Fe-PD/PGO in the ORR process reaches 3.98. Moreover, the Fe-PD/PGO electrocatalysts show a better electrochemical stability and methanol resistance than Pt/C catalyst.

References

[1]
Kiani M, Tian X Q, Zhang W. Non-precious metal electrocatalysts design for oxygen reduction reaction in polymer electrolyte membrane fuel cells: Recent advances, challenges and future perspectives. Coordination Chemistry Reviews, 2021, 441: 213954
CrossRef Google scholar
[2]
Zhang J, Yang H, Liu B. Coordination engineering of single-atom catalysts for the oxygen reduction reaction: A review. Advanced Energy Materials, 2021, 11(3): 2002473
CrossRef Google scholar
[3]
Shao M, Chang Q, Dodelet J P. . Recent advances in electrocatalysts for oxygen reduction reaction. Chemical Reviews, 2016, 116(6): 3594–3657
CrossRef Google scholar
[4]
Osmieri L, Pezzolato L, Specchia S. Recent trends on the application of PGM-free catalysts at the cathode of anion exchange membrane fuel cells. Current Opinion in Electrochemistry, 2018, 9: 240–256
CrossRef Google scholar
[5]
Wu J, Yang H. Platinum-based oxygen reduction electrocatalysts. Accounts of Chemical Research, 2013, 46(8): 1848–1857
CrossRef Google scholar
[6]
LeiCYangRZhaoJ, . Highly efficient and active Co-N-C catalysts for oxygen reduction and Zn–air batteries. Frontiers in Energy, 2024, early access, https://doi.org/10.1007/s11708-024-0928-6
[7]
Chen Y, Zhang J, Yang L. . Recent advances in non-precious metal–nitrogen–carbon single-site catalysts for CO2 electroreduction reaction to CO. Electrochemical Energy Reviews, 2022, 5(4): 11
CrossRef Google scholar
[8]
Fang C, Tang X, Wang J. . Performance of iron-air battery with iron nanoparticle-encapsulated C–N composite electrode. Frontiers in Energy, 2024, 18(1): 42–53
CrossRef Google scholar
[9]
Shen H, Thomas T, Rasaki S A. . Oxygen reduction reactions of Fe-NC catalysts: Current status and the way forward. Electrochemical Energy Reviews, 2019, 2(2): 252–276
CrossRef Google scholar
[10]
Bates J S, Johnson M R, Khamespanah F. . Heterogeneous MNC catalysts for aerobic oxidation reactions: Lessons from oxygen reduction electrocatalysts. Chemical Reviews, 2023, 123(9): 6233–6256
CrossRef Google scholar
[11]
Luo E, Chu Y, Liu J. . Pyrolyzed M–Nx catalysts for oxygen reduction reaction: Progress and prospects. Energy & Environmental Science, 2021, 14(4): 2158–2185
CrossRef Google scholar
[12]
Wang W, Jia Q, Mukerjee S. . Recent insights into the oxygen-reduction electrocatalysis of Fe/N/C materials. ACS Catalysis, 2019, 9(11): 10126–10141
CrossRef Google scholar
[13]
Tian H, Song A, Zhang P. . High durability of Fe-N-C single-atom catalysts with carbon vacancies toward the oxygen reduction reaction in alkaline media. Advanced Materials, 2023, 35(14): 2210714
CrossRef Google scholar
[14]
Luo Y, Zhang J, Chen J. . Dual-template construction of iron-nitrogen-codoped hierarchically porous carbon electrocatalyst for oxygen reduction reaction. Energy & Fuels, 2020, 34(12): 16720–16728
CrossRef Google scholar
[15]
Wang S, Wang H, Huang C. . Trifunctional electrocatalyst of N-doped graphitic carbon nanosheets encapsulated with CoFe alloy nanocrystals: The key roles of bimetal components and high-content graphitic-N. Applied Catalysis B: Environmental, 2021, 298: 120512
CrossRef Google scholar
[16]
Xu H, Jia H, Li H. . Dual carbon-hosted Co-N3 enabling unusual reaction pathway for efficient oxygen reduction reaction. Applied Catalysis B: Environmental, 2021, 297: 120390
CrossRef Google scholar
[17]
Han A, Sun W, Wan X. . Construction of CO4 atomic clusters to enable Fe-N4 motifs with highly active and durable oxygen reduction performance. Angewandte Chemie, 2023, 62(30): e202303185
CrossRef Google scholar
[18]
Qin J, Yang Z, Xing F. . Two-dimensional mesoporous materials for energy storage and conversion: Current status, chemical synthesis and challenging perspectives. Electrochemical Energy Reviews, 2023, 6(1): 9
CrossRef Google scholar
[19]
Chen S, Xiang S, Tan Z. . Exploration of the oxygen transport behavior in non-precious metal catalyst-based cathode catalyst layer for proton exchange membrane fuel cells. Frontiers in Energy, 2023, 17(1): 123–133
CrossRef Google scholar
[20]
Zhang Y, Qian L, Zhao W. . Highly efficient Fe-NC nanoparticles modified porous graphene composites for oxygen reduction reaction. Journal of the Electrochemical Society, 2018, 165(9): H510–H516
CrossRef Google scholar
[21]
Zhang Z, Sun J, Wang F. . Efficient oxygen reduction reaction (ORR) catalysts based on single iron atoms dispersed on a hierarchically structured porous carbon framework. Angewandte Chemie, 2018, 130(29): 9176–9181
CrossRef Google scholar
[22]
Hu Y, Lu Y, Zhao X. . Highly active N-doped carbon encapsulated Pd–Fe intermetallic nanoparticles for the oxygen reduction reaction. Nano Research, 2020, 13(9): 2365–2370
CrossRef Google scholar
[23]
Wu G, More K L, Johnston C M. . High-performance electrocatalysts for oxygen reduction derived from polyaniline, iron, and cobalt. Science, 2011, 332(6028): 443–447
CrossRef Google scholar
[24]
Sudarsono W, Wong W Y, Loh K S. . Elucidating the roles of the Fe-Nx active sites and pore characteristics on Fe-PANI-biomass-derived RGO as oxygen reduction catalysts in PEMFCs. Materials Research Bulletin, 2022, 145: 111526
CrossRef Google scholar
[25]
Zhang R, Liu L, Zhang J. . Fabrication of the pyrolyzing carbon-supported cobalt–dicyandiamide electrocatalysts and study on the active sites and mechanism for oxygen reduction in alkaline electrolyte. Journal of Solid State Electrochemistry, 2015, 19(6): 1695–1707
CrossRef Google scholar
[26]
Gao Y, Hou M, Qi M. . New insight into effect of potential on degradation of Fe-NC catalyst for ORR. Frontiers in Energy, 2021, 15(2): 421–430
CrossRef Google scholar
[27]
Lee W H, Lee D W, Kim H. Development of nitrogen-doped carbon catalysts using melamine-based polymer as a nitrogen precursor for the oxygen reduction reaction. Journal of the Electrochemical Society, 2015, 162(7): F744–F749
CrossRef Google scholar
[28]
Mahmood A, Xie N, Zhao B. . Optimizing surface N-doping of Fe-N-C catalysts derived from Fe/melamine-decorated polyaniline for oxygen reduction electrocatalysis. Advanced Materials Interfaces, 2021, 8(13): 2100197
CrossRef Google scholar
[29]
Wang D, Hu J, Yang J. . Fe and N co-doped carbon derived from melamine resin capsuled biomass as efficient oxygen reduction catalyst for air-cathode microbial fuel cells. International Journal of Hydrogen Energy, 2020, 45(4): 3163–3175
CrossRef Google scholar
[30]
Gupta S, Zhao S, Ogoke O. . Engineering favorable morphology and structure of Fe-N-C oxygen-reduction catalysts through tuning of nitrogen/carbon precursors. ChemSusChem, 2017, 10(4): 774–785
CrossRef Google scholar
[31]
Qiao Z, Zhang H, Karakalos S. . 3D polymer hydrogel for high-performance atomic iron-rich catalysts for oxygen reduction in acidic media. Applied Catalysis B: Environmental, 2017, 219: 629–639
CrossRef Google scholar
[32]
Wu M, Yang X, Cui X. . Engineering Fe-N4 electronic structure with adjacent Co-N2C2 and Co Nanoclusters on carbon nanotubes for efficient oxygen electrocatalysis. Nano-Micro Letters, 2023, 15(1): 232
CrossRef Google scholar
[33]
Gupta S, Zhao S, Ogoke O. . Engineering favorable morphology and structure of Fe-NC oxygen reduction catalysts via tuning nitrogen/carbon precursors. ChemSusChem, 2017, 10(4): 774–785
CrossRef Google scholar
[34]
Xie H, Du B, Huang X. . High density single fe atoms on mesoporous N-doped carbons: Noble metal-free electrocatalysts for oxygen reduction reaction in acidic and alkaline media. Small, 2023, 19(32): 2303214
CrossRef Google scholar
[35]
Sun X, Wei P, Gu S. . Atomic-level Fe-N-C coupled with Fe3C-Fe nanocomposites in carbon matrixes as high-efficiency bifunctional oxygen catalysts. Small, 2020, 16(6): 1906057
CrossRef Google scholar
[36]
Jiang W, Gu L, Li L. . Understanding the high activity of Fe-N-C electrocatalysts in oxygen reduction: Fe/Fe3C nanoparticles boost the activity of Fe–Nx. Journal of the American Chemical Society, 2016, 138(10): 3570–3578
CrossRef Google scholar
[37]
He J, Zheng T, Wu D. . Insights into the determining effect of carbon support properties on anchoring active sites in Fe-N-C catalysts toward the oxygen reduction reaction. ACS Catalysis, 2022, 12(3): 1601–1613
CrossRef Google scholar
[38]
Zhu G, Yang H, Jiang Y. . Modulating the electronic structure of FeCo nanoparticles in N-doped mesoporous carbon for efficient oxygen reduction reaction. Advanced Science, 2022, 9(15): 2200394
CrossRef Google scholar
[39]
Qiao M, Wang Y, Wang Q. . Hierarchically ordered porous carbon with atomically dispersed FeN4 for ultraefficient oxygen reduction reaction in proton-exchange membrane fuel cells. Angewandte Chemie International Edition, 2020, 59(7): 2688–2694
CrossRef Google scholar
[40]
Luo X, Wei X, Wang H. . Secondary-atom-doping enables robust Fe-N-C single-atom catalysts with enhanced oxygen reduction reaction. Nano-Micro Letters, 2020, 12(1): 163
CrossRef Google scholar
[41]
Wu G, Chen Y S, Xu B Q. Remarkable support effect of SWNTs in Pt catalyst for methanol electrooxidation. Electrochemistry Communications, 2005, 7(12): 1237–1243
CrossRef Google scholar
[42]
Soo L T, Loh K S, Mohamad A B. . Effect of nitrogen precursors on the electrochemical performance of nitrogen-doped reduced graphene oxide towards oxygen reduction reaction. Journal of Alloys and Compounds, 2016, 677: 112–120
CrossRef Google scholar
[43]
Yang H, Liu Y, Liu X. . Large-scale synthesis of N-doped carbon capsules supporting atomically dispersed iron for efficient oxygen reduction reaction electrocatalysis. eScience, 2022, 2(2): 227–234
CrossRef Google scholar
[44]
Chen J, Huang B, Cao R. . Steering local electronic configuration of Fe-N-C-based coupling catalysts via ligand engineering for efficient oxygen electroreduction. Advanced Functional Materials, 2023, 33(4): 2209315
CrossRef Google scholar
[45]
Wei J, Xia D, Wei Y. . Probing the oxygen reduction reaction intermediates and dynamic active site structures of molecular and pyrolyzed Fe-N-C electrocatalysts by in situ raman spectroscopy. ACS Catalysis, 2022, 12(13): 7811–7820
CrossRef Google scholar
[46]
Ni L, Gallenkamp C, Wagner S. . Identification of the catalytically dominant iron environment in iron-and nitrogen-doped carbon catalysts for the oxygen reduction reaction. Journal of the American Chemical Society, 2022, 144(37): 16827–16840
CrossRef Google scholar
[47]
Zhang G, Liu X, Yu P. . Fe3C coupled with Fe-Nx supported on N-doped carbon as oxygen reduction catalyst for assembling Zn-air battery to drive water splitting. Chinese Chemical Letters, 2022, 33(8): 3903–3908
CrossRef Google scholar
[48]
Li J, Xiao D, Wang P. . Highly strong interaction between Fe/Fe3C nanoparticles and N-doped carbon toward enhanced oxygen reduction reaction performance. Particle & Particle Systems Characterization, 2023, 40(1): 2200141
CrossRef Google scholar

Acknowledgements

This work was supported by the Natural Science Foundation of the Jiangsu Higher Education Institutions of China (Grant No. 23KJD430007), Qinglan Project of Jiangsu Province, Innovative Research Group Project of the National Natural Science Foundation of China (Grant No. 51902145), Nanjing Technology Innovation Team of Optometric Materials and Application and Doctoral Start-up Fund Research supported by Jinling Institute of Technology (jit-b-202026), and the European Union (ERDF) and Région Nouvelle Aquitaine.

Competing Interests

The authors declare that they have no competing interests.

Electronic Supplementary Material

Supplementary material is available in the online version of this article at https://doi.org/10.1007/s11708-024-0956-2 and is accessible for authorized users.

RIGHTS & PERMISSIONS

2024 Higher Education Press 2024
AI Summary AI Mindmap
PDF(3395 KB)

Supplementary files

FEP-24028-OF-ZY_suppl_1 (2082 KB)

1762

Accesses

2

Citations

21

Altmetric

Detail

Sections
Recommended

/