Next Article in Journal
Electronic, Optical, Thermoelectric and Elastic Properties of RbxCs1−xPbBr3 Perovskite
Next Article in Special Issue
Improving the Reaction Kinetics by Annealing MoS2/PVP Nanoflowers for Sodium-Ion Storage
Previous Article in Journal
Dioscorea spp.: Bioactive Compounds and Potential for the Treatment of Inflammatory and Metabolic Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rational Fabrication of Defect-Rich and Hierarchically Porous Fe-N-C Nanosheets as Highly Efficient Oxygen Reduction Electrocatalysts for Zinc-Air Battery

State Key Laboratory of Chemistry and Utilization of Carbon Based Energy Resources, College of Chemistry, Xinjiang University, Urumqi 830017, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(7), 2879; https://doi.org/10.3390/molecules28072879
Submission received: 22 February 2023 / Revised: 13 March 2023 / Accepted: 16 March 2023 / Published: 23 March 2023

Abstract

:
The rational design of morphology and structure for oxygen reduction reaction (ORR) catalysts still remains a critical challenge. Herein, we successfully construct defect-rich and hierarchically porous Fe-N-C nanosheets (Fe-N-CNSs), by taking advantage of metal-organic complexation and a mesoporous template. Benefiting from the advantages of high density of active sites, fast mass transfer channels, and sufficient reaction area, the optimal Fe-N-CNSs demonstrate satisfactory ORR activity with an excellent half-wave potential of up to 0.87 V, desirable durability, and robust methanol tolerance. Noteworthy, the Fe-N-CNSs based zinc–air battery shows significant performance with a peak power density of 128.20 mW cm−2 and open circuit voltage of 1.53 V, which reveals that the Fe-N-CNSs catalysts present promising practical application prospects. Therefore, we believe that this research will provide guidance for the optimization of Fe-N-C materials.

1. Introduction

Increasing demand for energy and the massive consumption of non-renewable energy sources have prompted us to explore new types of energy storage devices [1,2,3]. With excellent theoretical energy density (1086 Wh kg−1), affordability, and environmental friendliness, zinc–air batteries (ZABs) have been considered as among the most potential candidates to cope with the energy requirements. The electrochemical performances of ZABs are primarily dependent on the efficiency of oxygen reduction reaction (ORR) [4,5,6,7]. Nevertheless, the sluggish kinetic of ORR suffering from a multiple electron transfer causes a confined output power and energy storage efficiency of ZABs. It is, therefore, essential to create an effective ORR catalyst to promote its slow kinetics, reduce the overpotential, and improve energy storage efficiency the of ZABs. The most effective ORR catalysts have been traditionally regarded as Pt-based ones due to a low binding energy with oxygen [8,9]. The prohibitive cost and inferior stability, however, prevent its large-scale application in air cathodes [10]. Therefore, designing and exploring cost-effective, efficient, and durable ORR catalysts is highly vital. Metal-nitrogen-carbon (M-N-C, M = Fe, Co, Ni, Cu, etc.) catalysts were already extensively studied as ORR catalysts [11,12,13,14]. The transition metals with unfilled d orbitals can bind better to oxygen intermediates and thus exhibit excellent ORR native positive activity [15,16,17,18]. In particular, Fe-N-C with the encouraging ORR activity and excellent catalytic durability has been deemed as the most prospective replacement for Pt-based electrocatalysts [19,20,21,22,23]. Among all 3d-transition metal elements, Fe shows the closest-to-optimum OH* binding energy. After O2 is adsorbed to the centre of the Fe atom, the electrons in its 2π orbital are transferred to the hole dz2 orbital of Fe and the partially filled dxy or dxz is back-bonded to the 2π* orbital of oxygen [24,25,26,27,28,29,30]. However, the micromorphology and nanostructure of numerous reported Fe-N-C materials are inhomogeneous, which leads the catalytic site to be challenging to expose to reduction of catalytic activity [31,32,33]. To address this issue, the reasonable design of the micromorphology and nanostructure of Fe-N-C materials is fundamental for constructing highly efficient Fe-N-C catalysts [23,34,35,36,37].
Two-dimensional (2D) materials were intensively explored as ORR catalysts because of large specific surface area, unique physicochemical properties as well as thin thickness, which can provide rapid mass transfer channels, an abundance of active sites, and sufficient reaction area [38,39,40,41]. Typically, 2D materials are synthesized by top-down or bottom-up methods, including chemical vapor deposition (CVD), physical/chemical exfoliation, and wet chemical synthesis [42,43,44,45]. The simple and green template strategy has been widely applied in the fabrication of 2D materials because of its ability to construct specific morphologies and hierarchical porous structures [46,47]. In addition, the construction of 2D materials with a defect-rich and hierarchical porous structures is highly desirable and challenging for improving ORR performance [48,49]. For instance, the group of Wang prepared Fe-N-C nanosheets using layered montmorillonite (MMT) as a template with a half-wave potential (E1/2) of 0.87 V [50]. The corresponding ZAB, assembled with the obtained material as a catalyst, exhibited a peak power density of 92.5 mW cm−2. Nevertheless, MMT cannot provide abundant micro/mesopores or defects to further enhance ORR performance.
Inspired by the above analysis, herein, we successfully synthesized Fe-N-C nanosheets (Fe-N-CNSs), by employing metal-organic complexation and a mesoporous magnesium carbonate base (MCB) template. The obtained Fe-N-CNSs catalysts featured a defect-rich, hierarchically porous, and high specific surface area (1303.43 m2 g−1), which can provide an abundance of active sites, mass transfer channels, as well as sufficient reaction area. The Fe-N-CNSs catalysts exhibited excellent ORR performance (E1/2 = 0.87 V), surpassing commercial 20% Pt/C (E1/2 = 0.85 V). In addition, the ZABs were assembled with Fe-N-CNSs as the catalysts for the air cathode, which demonstrated superior performance compared to Pt/C, reflecting the promising potential of Fe-N-CNSs in practical applications. The strategy that was developed in this work provides a universal approach to optimize Fe-N-C catalysts.

2. Results and Discussion

Scheme 1 depicts the construction path of Fe-N-CNSs. Owing to the cheating ability, ethylene diamine tetra acetic acid (EDTA) can form complexes with Fe3+. Basic magnesium carbonate (MCB) is decomposed into MgO, CO2, and H2O in a high temperature pyrolysis process. The generated MgO is used as a mesoporous template that can induce the polymerization of small molecules on its surface, while CO2 and H2O escaping from the carbon substrate can create abundant micropores and produce exfoliation, which promotes the construction of sheet-like structures [51,52]. The scanning electron microscopy (SEM) analysis was performed on the samples that were acquired to examine their micromorphology. The pyrolysis temperature can significantly affect the structure and micromorphology of carbon-based products. Hence, we investigate how the pyrolysis temperatures affect the micromorphology of the materials. When the pyrolysis temperature is 800 °C, Fe-N-CNSs-800 is mainly composed of carbon fragments and carbon blocks, showing an irregular curled sheet morphology, as shown in Figure S1, which might be due to insufficient activation. As a result, the substance becomes less porous and has fewer mass transfer pathways. Noteworthy, as shown in Figure 1a and Figure S2, Fe-N-CNSs-900 shows a cross-linked network consisting of 2D porous nanosheets. The large surface area of materials provided by the 2D nanosheet structure makes it simpler to reveal active sites, while the hierarchically porous structure offers a high density of active sites and a significant number of rapid mass transfer pathways. As a result, we anticipate that the produced 2D hierarchically porous Fe-N-C nanosheets will function satisfactorily in ORR. Nevertheless, for Fe-N-CNSs-1000, as shown in Figure S3, the disruption of the sheet-like structures results in fewer catalytic sites. According to the aforementioned findings, one of the crucial elements in the production of a hierarchically porous nanosheet structure is a proper pyrolysis temperature. With regard to Fe-N-CNSs, a hierarchically porous nanosheet structure is best when the pyrolysis temperature is 900 °C. In addition, we look at the morphology of a sample of Fe-N-C-900 to reveal how MCB affects material morphology. It presents a distinctly larger bulk structure, as shown in Figure S4, without lamellar structures or pores being found. The above results indicate that the introduction of MCB can generate 2D porous nanosheet structures, which reflects the dual efficacy of MCB as a template and an activator.
Subsequently, the microscopic morphology of Fe-N-CNSs was explored by transmission electron microscopy (TEM). The Fe-N-CNSs shows a thin sheet-like morphology and the presence of many mesopores on the surface, as shown in Figure 1b and Figure S7, without Fe-associated nanoparticles or nanoclusters, which provides a high surface area and fast mass transfer channels. Figure 1c illustrates the existence of lattice distortion defects in graphitic carbon in the high-resolution transmission electron microscopy (HRTEM) image of Fe-N-CNS, which proves that the heteroatoms are doped into the carbon substrate [53]. In addition, lattice stripes with a lattice spacing of 0.34 nm, which corresponds to the (002) crystal plane of graphitic carbon, are found in the HRTEM images. Further analysis of Fe-N-CNSs was carried out using HADDF-STEM and EDS mapping in conjunction with TEM. As shown in Figure 1d, the high angle angular dark field-scanning transmission electron microscopy (HADDF-STEM) picture clearly shows the existence of pores and hollow structures, which serves as further proof that porous structures were successfully prepared. The results associated element mapping and confirmed that Fe, N, and C elements are all uniformly dispersed in Fe-N-CNSs, proving that Fe and N have been effectively doped into the carbon skeleton. Additionally, the creation of Fe-Nx sites is supported by the substantial overlap between the dispersion of the Fe and N elements.
X-ray diffraction (XRD) was used to evaluate the catalysts in order to better identify their crystalline structure. As shown in Figure 2a, several catalysts exhibit characteristic peaks of the (002) and (101) planes of carbon at 23° and 44°, which corresponds to the classical carbon structure [54]. Meanwhile, the absence of characteristic peaks of Fe and its compounds further confirms the inexistence of highly crystalline Fe-containing species in Fe-N-CNSs catalysts. As a result, we can assume that Fe-Nx may make up the majority of the Fe element in Fe-N-CNSs. By using X-ray photoelectron spectroscopy (XPS), the chemical makeup of the catalyst surface was examined. As shown in Figure 2b, XPS surveys show obvious signals of C, N, and O elements; nevertheless, the signal of the Fe element is undetected, which might be attributed to the low content of elemental Fe, and the findings are aggregated in Table 1. As shown in Figures S8–S10, the C 1s XPS spectra are deconvoluted into five peaks, which are indexed to C-C (284.5 eV, sp2 hybridized carbon), C-C (285.2 eV sp3 disordered carbon), C-N (286.0 eV), C-O (286.5 eV), and C=O (287.1 eV), accordingly [55]. The presence of the C-N peak demonstrates that the N atom has been successfully doped into the carbon skeleton. In addition, the ratio of C-C (sp2) and C-C (sp3) (sp2/sp3) reflects the degree of defects in the carbon-based material. The Fe-N-CNSs shows a sp2/sp3 of 0.99, lower than the N-CNSs (1.11) and Fe-N-C (1.08), indicating that Fe-N-CNSs catalyst is rich in defects. The Fe 2p XPS spectrum of Fe-N-CNSs presents two groups of peaks for Fe2+ (711.3 and 724.5 eV) and Fe3+ (716.0 and 729.1 eV), and no obvious Fe0 appeared, as shown in Figure S11, further proving that no Fe or its compounds are formed. As shown in Figure 2c, to further determine the existing configuration of the N element, the N 1s spectra are deconvoluted into five peaks and the findings are aggregated in Table 2. The N 1s spectra of the catalysts exhibit five peaks at about 398.3, 399.2, 400.3, 401.3 and 402.7 eV that, respectively, correspond to pyridinic N, Fe-N, pyrrolic N, graphitic N, and oxidized N [56]. As there is a Fe-N peak, Fe-Nx sites have probably been formed, and Fe-Nx sites are the most dominant sites for Fe-N-C catalysts [57]. Additionally, the pyridine N content in Fe-N-CNSs reaches up to 41.31%. In addition to the electrons provided to the conjugated π-bond, pyridine N also contains a pair of lone electrons that facilitate the adsorption of O2, thereby increasing the onset potential [58,59]. The inductively coupled plasma optical emission spectrometry (ICP-OES) results show that Fe-N-CNSs presents a higher Fe content of 0.24% compared to Fe-N-C (0.07%).
Subsequently, the extent of defects in the carbon structure of the catalyst was investigated by Raman spectroscopy. As shown in Figure 2d and Figure S12, the D and G bands at 1341 cm−1 and 1590 cm−1, which stand for lattice defects and carbon atom graphitization, accordingly, are clearly visible in the Raman spectra. For the Raman spectrum, the ratio of the D peak and G peak intensity (ID/IG) can reflect some extent the level of disorder or defects in the structure of carbon-based materials [60]. The Fe-N-CNSs exhibits an ID/IG of 1.04, higher than the N-CNSs (0.96) and Fe-N-C (0.98), indicating that the introduction of a templating agent increases the defect degree of the material, and these abundant defect sites may create more catalytic sites, thereby enhancing the catalytic efficiency.
Via a series of SEM and TEM characterizations, the macro- and mesoporous structure of materials can be directly viewed; nonetheless, the smaller pore size micro-pores cannot be seen. Therefore, the pore structure of the obtained materials was further characterized by N2 adsorption and desorption measurements with the outcomes noted in Table 3. All catalysts present an IV isotherm with a distinct hysteresis loop, which confirm the existence of abundant mesopores, as shown in Figure 2e and Figure S13. The Fe-N-CNSs-900 presents the highest surface area (1303.43 m2 g−1) among all the constructed materials, and this value significantly outperforms Fe-N-C (421.15 m2 g−1). The Fe-N-CNSs displays such a large surface area due to its rich porosity and distinctive 2D nanosheet structure. As shown in Figure 2f and Figure S14, from the pore size distribution curve, the Fe-N-CNSs demonstrates more abundant mesopores and micropores than Fe-N-C. These results indicate that MCB as a template agent and activation agent can produce a unique 2D nanosheet morphology, abundance of micro/mesopores, and ultrahigh specific surface area. The numerous micropores deliver more abundant active sites, while mesopores facilitate the transport of reaction intermediates. The high specific surface area provides sufficient reaction area and exposes more active surface sites.
The synthesized Fe-N-CNSs materials are expected to be excellent ORR catalysts due to a defect-rich, hierarchical porous, and ultrahigh surface area. Using cyclic voltammetry (CV), the Fe-N-CNSs is first investigated. As shown in Figure 3a, an obvious oxygen reduction peak is observed at approximately 0.8 V (vs. RHE) in the O2 saturated electrolyte, while it is not observed in the saturated N2 electrolyte, demonstrating that Fe-N-CNSs exhibits positive activity toward ORR in O2 saturated electrolytes. The ORR activity of these materials was further characterized by linear scanning voltammetry (LSV). As shown in Figure 3b,c, the target catalyst, Fe-N-CNSs, exhibits the highest ORR activity, including an excellent E1/2 (0.87 V) and kinetic current density at 0.8 V (jk@0.8 V = 27.66 mA cm−2). The ORR activity of Fe-N-CNSs is not only better than that of the Fe-N-C catalyst but better than that of the Pt/C catalyst (E1/2 = 0.85 V, jk@0.8 V = 20.61 mA cm−2) and many reported Fe-N-C catalysts (Table S1). The carbonation temperature can affect the catalytic performance of carbon-based materials to some extent, so we perform a comparative analysis for the ORR performance of the samples at different temperatures. The E1/2 of the three catalysts reach 0.83, 0.87 and 0.84 V, respectively, and the Fe-N-CNSs-900 catalysts present the highest E1/2, as shown in Figure S15. Additionally, the amount of Fe (NO3)3·9H2O added is investigated to obtain better experimental conditions, as shown in Figure S16, and the highest ORR performance is observed for Fe-N-CNSs-30. The LSV curves are recorded at different rotational speeds, as shown in Figure 3d. The Koutecky–Levich (K–L) plots present several fitted lines with almost identical slopes. The half-wave potential steadily declines while the limiting current density gradually rises as the speed grows, representing the kinetics of vortex diffusion during ORR. Depending on the slope, in the range of 0.3, 0.4, 0.5 and 0.6 V, the average number of electrons transferred (n) for Fe-N-CNSs are calculated to be about 3.94, which reveals that Fe-N-CNSs obeys the first-order kinetics and approximately a 4e- pathway. Electrochemical impedance spectroscopy (EIS) was performed to further analyze the kinetics of the ORR process in the material. As shown in Figure S17, Fe-N-CNSs exhibits smaller semicircle diameters than Pt/C in the low frequency range, which indicates that Fe-N-CNSs has faster transfer kinetics at the electrode/electrolyte interface due to its high density of active sites. Furthermore, in the high frequency range, the Fe-N-CNSs has steeper tilt lines, demonstrating that Fe-N-CNSs has a better diffusion rate at the reaction interface due to its hierarchically porous nanosheet structure.
An excellent ORR catalyst requires not only excellent ORR activity but robust methanol tolerance and remarkable stability. The durability and methanol resistance of Fe-N-CNSs catalysts are examined by chronoamperometric i-t measurements. As a comparison, we also test commercial 20% Pt/C using the same method. As shown in Figure 3e, after 400 s of 3 mL methanol addition to the electrolyte, the relative current of the commercial 20% Pt/C catalysts show a significant drop, while the Fe-N-CNSs catalysts remain almost unchanged. The result indicates the Fe-N-CNSs exhibits a strong methanol tolerance. Moreover, the Fe-N-CNSs catalysts still show 90% current retention after 12,000 s of cycling, as shown in Figure 3f, indicating that the Fe-N-CNSs catalysts are highly stable. The catalysts display exceptional ORR performance, owing to the advantages of abundant catalytic sites, sufficient reaction area and fast mass transfer channels. To investigate the catalytic sites of the Fe-N-CNSs, KSCN is added to poison the Fe-Nx sites during ORR as SCN- has a strong affinity for Fe3+. As shown in Figure S18, the ORR performance drops obviously after adding of KSCN, indicating that Fe-Nx is the most critical catalytic site for Fe-N-CNSs. Notably, the half-wave potential still remains at a high level after the addition of KSCN, indicating that the C-N sites also contribute to increased catalytic activity. Combined with the analysis of the XPS data, the following conclusions are drawn: Fe-Nx and pyridinic N contribute more to ORR, and the higher their content, the better the ORR activity.
Considering the splendid ORR performance, we assemble a Fe-N-CNSs-based zinc–air battery (ZAB), as shown in Figure 4a, to further explore its practical application prospects. As shown in Figure 4b, the open circuit voltage (OCV) of the assembled ZAB is tested separately by the open circuit potential time measurement (OCPT) and an electronic multimeter. The Fe-N-CNSs-based ZAB can sustain a high OCV of 1.53 V, which outperforms commercial 20% Pt/C-based ZABs (1.44 V). Using the assembled Fe-N-CNSs-based ZAB as a power source, an LED bulb (1.2 V) with the word “XJU” can be lit, as shown in Figure 4c, which further confirms that the Fe-N-CNSs materials present an excellent prospect of practical application. The Fe-N-CNSs-based ZAB exhibits a peak power density of 128.20 mW cm−2, as shown in Figure 4d, which exceeds the Pt/C-based ZAB (101.80 mW cm−2) and outperforms most of the reported ZABs assembled from Fe-N-C materials (Table S2). The discharge voltage of Fe-N-CNSs-based ZAB remains stable at all current densities (5–50 mA cm−2) and still maintains 1.10 V, as shown in Figure 4e, which outperforms the Pt/C-based ZAB. The above results reveal that Fe-N-CNSs-based ZAB exhibits a better capacity and multiplicative performance. Furthermore, the specific capacity of the Fe-N-CNSs-based ZAB reaches a high energy density of 746.37 mA h gZn−1, which compares favorably with the Pt/C-based ZAB (713.35 mA h gZn−1), as shown in Figure 4f.

3. Materials and Methods

3.1. Materials

Magnesium carbonate basic (MCB), ethanol, Fe (NO3)3·9H2O, ethylene diamine tetra acetic acid (EDTA), commercial 20% Pt/C, concentrated hydrochloric acid (HCl), 0.5 wt% Nafion, concentrated hydrochloric acid (HCl) and methanol were all purchased from Alfa Aesar and Aladdin chemical reagent company. All of the compounds were utilized directly after delivery without additional purification and are of analytical grade.

3.2. Catalyst Preparation

First, 0.5 g of MCB was added to 40 mL of absolute ethanol and ultrasonically dispersed for 0.5 h. Then, 0.8 g EDTA and 30 mg Fe (NO3)3·9H2O were added and stirring was continued for 5 h. Subsequently, the above mixture was dried in an oven at 60 °C to remove ethanol, and the resulting pale–yellow solid was ground into a homogeneous powder. Finally, the precursor was pyrolyzed in a tube furnace under N2 atmosphere at 900 °C for 2 h at a heating rate of 5 °C min−1. The resulting black solid was added to 60 mL of 2 M HCl solution and stirred for 12 h. The prepared material was washed three times by centrifugation with deionized water and absolute ethanol, respectively, and then dried in a vacuum oven overnight. The prepared samples were labeled as Fe-N-CNSs (also called Fe-N-CNSs-900 and Fe-N-CNSs-30). For comparison, the precursors were also pyrolyzed at 800 °C and 1000 °C, and the prepared carbon materials were labeled as Fe-N-CNSs-x (x = 800 and 1000). The addition of the Fe source was adjusted, and the resulting materials were labelled Fe-N-CNSs-m (m = 15, 30, 60). In addition, the precursors without Fe (NO3)3·9H2O or MCB were pyrolyzed at 900 °C, and the prepared materials were labeled as N-CNSs or Fe-N-C.

3.3. Physical Characterizations

The micromorphology of the materials was observed by SEM (SU-4800, Hitachi, Japan) and TEM (JEM-2100F, JEOL, Japan). The crystal structure of the materials was analyzed by XRD (Bruker D8, using filtered Cu Kα radiation). The degree of defects in the materials was analyzed by a Raman spectrometer (Vertrex 70, Bruker, Germany) with a laser excitation wavelength of 532 nm. The exact content of elemental Fe in the material was obtained by ICP-OES (Optima 8000, PerkinElmer, USA). N2 adsorption/desorption tests were conducted at 77 K by a gas adsorption analyzer (ASAP 2020, Micromeritics, USA). The specific surface area and pore size of the samples were calculated by Brunauere–Emmette–Teller (BET) and Barrett–Joyner–Halenda (BJH) models, respectively. The chemical composition of the catalyst was analyzed by XPS (ESCALAB 250, Thermo Scientific, USA).

3.4. Electrochemical Measurements

In a standard three-electrode setup, electrochemical experiments were carried out at room temperature. The electrochemical workstation is CHI 760E. The 0.1 M KOH solution saturated with O2/N2 serves as the electrolyte. The working electrode, counter electrode, and reference electrode that we employed were a spinning disk ring electrode, a platinum wire electrode, and a saturated calomel electrode, respectively. The working electrode was polished with Al2O3 polishing powders with diameters of 1 μm and 50 nm, respectively. To create a well-dispersed ink, 2.5 mg of catalyst and 20 μL of 0.5 wt% Nafion were dispersed in 480 μL of ethanol. Sonication was then applied for 30 min. A loading of 254.78 μg cm−2 was achieved by adding 10 μL of catalyst ink dropwise to the working electrode and drying it at room temperature. Comparatively, the working electrode was loaded with commercial 20% Pt/C, which was applied dropwise. The potentials mentioned in this work are relative to the reversible hydrogen electrode (RHE). Using a scan rate of 50 mV s−1, the CV curves of catalysts were examined in N2 or O2 saturated 0.1 M KOH solution. In an O2 saturated 0.1 M KOH solution, the LSV curves of the catalysts were examined at different rotational speeds of 400, 625, 900, 1225, 1600 and 2025 rpm with a scan rate of 10 mV s−1. The number of transferred electrons (n) can be calculated by the Koutecky–Levich (K–L) formula:
1 j = 1 j k + 1 j l = 1 j k + 1 B ω 1 / 2
B = 0.2 n F C 0 D 0 2 / 3 υ 1 / 6
In the formula, j is the measured current density, jl is the limiting current density, jk is the kinetic current density, and the calculation formula is jk = (jl × j)/(jlj). ω is the rotation rate of the rotating disk electrode (RDE). B is determined by the slope of the K–L equation, n is the number of transferred electrons, F is the Faraday constant (96,485 C mol−1), C0 is the volume concentration of O2 (1.2 × 10−6 mol cm−3), D0 is the O2 at 0.1 M of the diffusion coefficient in KOH (1.9 × 10−5 cm2 s−1), υ is the dynamic viscosity (0.01 cm2 s−1). A constant of 0.2 is adopted when the rotation rate is expressed in rpm.
Electrochemical impedance spectroscopy (EIS) was performed in 0.1 M KOH by applying an AC voltage with 5 mV amplitude in frequencies ranging from 0.01 Hz to 100 kHz. The specific data can be fitted by Zview software.

3.5. Zn-Air Battery Test

The performance characterization of all ZABs were performed on an electrochemical workstation (CHI 760E). The zinc–air batteries used 6 M KOH as the electrolyte, zinc foil as the negative electrode and carbon fiber paper (3 mg cm−2) coated with catalyst ink as the positive electrode. Following this, 3 mg of Fe-N-CNSs and 0.3 mg of acetylene black were dispersed in 480 μL of ethanol and 20 μL of 5% Nafion solution, followed by sonication for 30 min to obtain a well-dispersed ink. For comparison, a Pt/C-based air cathode was prepared using the same method.

4. Conclusions

In summary, we successfully fabricated Fe-N-C nanosheets (Fe-N-CNSs) by a simple, versatile, and green template synthesis strategy. The optimal Fe-N-CNSs catalysts present the large surface area, hierarchical porous and defect-rich structure, which provides sufficient reaction area, mass transfer channels, as well as abundance of active sites, accelerating kinetic ORR activity on Fe-N-CNSs, along with a significant E1/2 of up to 0.87 V and strong stability (superior to Pt/C). Meanwhile, the Fe-N-CNSs-based ZAB also exhibits superior performance with a peak power density of 128.20 mW cm−2 and open circuit voltage of 1.53 V, indicating the Fe-N-CNSs catalysts present a prospective application in practice. In addition, our work provides a novel strategy for further optimizing and improving the catalytic performance of Fe-N-C materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28072879/s1. Figure S1: SEM images of Fe-N-CNSs-800; Figure S2. SEM image of Fe-N-CNSs-900; Figure S3. SEM images of Fe-N-CNSs-1000; Figure S4. SEM images of Fe-N-C-900; Figure S5. SEM images of Fe-N-CNSs-15; Figure S6. SEM images of Fe-N-CNSs-60; Figure S7. TEM images of Fe-N-CNSs; Figure S8. High-resolution C 1s spectrum of N-CNSs; Figure S9. High-resolution C 1s spectrum of Fe-N-CNSs; Figure S10. High-resolution C 1s spectrum of Fe-N-C; Figure S11. High-resolution Fe 2p spectrum of Fe-N-CNSs; Figure S12. Raman spectra of the synthesized materials; Figure S13. N2 adsorption-desorption isotherms of Fe-N-CNSs-800, Fe-N-CNSs-900 and Fe-N-CNSs-1000; Figure S14. The pore size distribution of Fe-N-CNSs-800, Fe-N-CNSs-900 and Fe-N-CNSs-1000; Figure S15. LSV curves of Fe-N-CNSs-800, Fe-N-CNSs-900, Fe-N-CNSs-1000; Figure S16. LSV curves of Fe-N-CNSs-15, Fe-N-CNSs-30, Fe-N-CNSs-60; Figure S17. Nyquist plots of electrochemical impedance spectroscopy (EIS) for Fe-N-CNSs and Pt/C; Figure S18. Poison experiment by KSCN; Table S1. Comparison of ORR activity of various Fe-N-C catalysts; Table S2. Comparison of the performance of the primary zinc-air batteries with the present work.

Author Contributions

Conceptualization, S.L. and X.W.; methodology, S.L.; software, S.L.; validation, S.L., Y.L. and X.Z.; formal analysis, S.L.; investigation, S.L.; resources, J.G.; data curation, S.L. and S.E.; writing—original draft preparation, S.L.; writing—review and editing, X.W. and J.G.; visualization, Z.Y.; supervision, X.W. and J.G.; project administration, X.W. and J.G.; funding acquisition, X.W. and J.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Natural Science Foundation of Xinjiang Uygur Autonomous Region of China (2020D01C062, 2020Q72, 2021D01D09 and 2021D01C097), the National Natural Science Foundation of China (22105163, U2003307), the Open Project of Key Laboratory in Xinjiang Uygur Autonomous Region of China (2022D04032), Scientific Research Program of the Higher Education Institution of Xinjiang (XJEDU2021Y005), Open Project of PCOSS (Xiamen University) (2021X20), and the University Student Innovation Program (S202110755132).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Dare are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds Fe-N-CNSs, N-CNSs, Fe-N-C, Fe-N-CNSs-800, Fe-N-CNSs-1000, Fe-N-CNSs-15 and Fe-N-CNSs-60 are available from the authors.

References

  1. Xu, M.; Dou, H.; Zhang, Z.; Zheng, Y.; Ren, B.; Ma, Q.; Wen, G.; Luo, D.; Yu, A.; Zhang, L. Hierarchically nanostructured solid-state electrolyte for flexible rechargeable zinc–air batteries. Angew. Chem. Int. Ed. 2022, 134, e202117703. [Google Scholar] [CrossRef]
  2. Xu, H.F.; Zhai, Q.G.; Jiang, Y.C.; Huang, H.; Li, S.N. Carbon foam-supported CoN nanoparticles and carbon nanotubes hybrids as bifunctional reduction electrocatalyst. Catal. Commun. 2022, 163, 106408. [Google Scholar] [CrossRef]
  3. Shao, Z.; Zhu, Q.; Sun, Y.; Zhang, Y.; Jiang, Y.; Deng, S.; Zhang, W.; Huang, K.; Feng, S. Phase-reconfiguration-induced NiS/NiFe2O4 composite for performance-enhanced zinc−air batteries. Adv. Mater. 2022, 34, 2110172. [Google Scholar] [CrossRef] [PubMed]
  4. Chang, Y.; Chen, J.; Jia, J.; Hu, X.; Yang, H.; Jia, M.; Wen, Z. The fluorine-doped and defects engineered carbon nanosheets as advanced electrocatalysts for oxygen electroreduction. Appl. Catal. B Environ. 2021, 284, 119721. [Google Scholar] [CrossRef]
  5. Liu, C.; Li, Y.; Cui, J.; Qian, Z.; Liu, D. Fabrication of ORR/OER Electrocatalysts with simple one-step strategy from sustainable cornstalks. Catal. Commun. 2022, 171, 106525. [Google Scholar] [CrossRef]
  6. Zheng, Q.; Xiong, Y.; Tang, K.; Wu, M.; Hu, H.; Zhou, T.; Wu, Y.; Cao, Z.; Sun, J.; Yu, X. Modulation of pore-size in N, S-codoped carbon/Co9S8 hybrid for a stronger O2 affinity toward rechargable zinc-air battery. Nano Energy 2022, 92, 106750. [Google Scholar] [CrossRef]
  7. Li, W.; Liu, B.; Liu, D.; Guo, P.; Liu, J.; Wang, R.; Guo, Y.; Tu, X.; Pan, H.; Sun, D. Alloying Co species into ordered and interconnected macroporous carbon polyhedra for efficient oxygen reduction reaction in rechargeable zinc–air batteries. Adv. Mater. 2022, 34, 2109605. [Google Scholar] [CrossRef]
  8. Liu, J.; He, T.; Wang, Q.; Zhou, Z.; Zhang, Y.; Wu, H.; Li, Q.; Zheng, J.; Sun, Z.; Lei, Y. Confining ultrasmall bimetallic alloys in porous N–carbon for use as scalable and sustainable electrocatalysts for rechargeable Zn–air batteries. J. Mater. Chem. A 2019, 7, 12451–12456. [Google Scholar] [CrossRef]
  9. Mei, Z.Y.; Cai, S.; Zhao, G.; Zou, X.; Fu, Y.; Jiang, J.; An, Q.; Li, M.; Liu, T.; Guo, H. Boosting the ORR active and Zn-air battery performance through ameliorating the coordination environment of iron phthalocyanine. Chem. Eng. J. 2022, 430, 132691. [Google Scholar] [CrossRef]
  10. Li, J.; Chen, M.; Cullen, D.A.; Hwang, S.; Wang, M.; Li, B.; Liu, K.; Karakalos, S.; Lucero, M.; Zhang, H. Atomically dispersed manganese catalysts for oxygen reduction in proton-exchange membrane fuel cells. Nat. Catal. 2018, 1, 935–945. [Google Scholar] [CrossRef]
  11. Das, I.; Noori, M.T.; Shaikh, M.; Ghangrekar, M.M.; Ananthakrishnan, R. Synthesis and application of zirconium metal–organic framework in microbial fuel cells as a cost-effective oxygen reduction catalyst with competitive performance. ACS Appl. Energy Mater. 2020, 3, 3512–3520. [Google Scholar] [CrossRef]
  12. Kong, F.; Cui, X.; Huang, Y.; Yao, H.; Chen, Y.; Tian, H.; Meng, G.; Chen, C.; Chang, Z.; Shi, J. N-Doped carbon electrocatalyst: Marked ORR activity in acidic media without the contribution from metal sites? Angew. Chem. Int. Ed. 2022, 61, e202116290. [Google Scholar] [CrossRef] [PubMed]
  13. Zhao, Z.; Duan, L.; Zhao, Y.; Wang, L.; Zhang, J.; Bu, F.; Sun, Z.; Zhang, T.; Liu, M.; Chen, H. Constructing unique mesoporous carbon superstructures via monomicelle interface confined assembly. J. Am. Chem. Soc. 2022, 144, 11767–11777. [Google Scholar] [CrossRef]
  14. Singh, H.; Zhuang, S.; Ingis, B.; Nunna, B.B.; Lee, E.S. Carbon-based catalysts for oxygen reduction reaction: A review on degradation mechanisms. Carbon 2019, 151, 160–174. [Google Scholar] [CrossRef]
  15. Liang, Q.; Meng, F.; Song, K.; Wang, D.; Zhang, W. Exploiting asymmetric Co states in a Co-NC catalyst for an efficient oxygen reduction reaction. Symmetry 2022, 14, 2496. [Google Scholar] [CrossRef]
  16. Zhao, C.-X.; Liu, J.-N.; Wang, J.; Wang, C.; Guo, X.; Li, X.-Y.; Chen, X.; Song, L.; Li, B.-Q.; Zhang, Q. A clicking confinement strategy to fabricate transition metal single-atom sites for bifunctional oxygen electrocatalysis. Sci. Adv. 2022, 8, eabn5091. [Google Scholar] [CrossRef]
  17. Zhou, X.; Song, K.; Feng, Y.; Jiang, C.; Chen, Z.; Wang, Z.; Yue, N.; Ge, X.; Zhang, W.; Zheng, W. Individually-atomic governing d—π* orbital interactions via Cu-promoted optimization of Fe-d band centers for high-efficiency zinc-air battery. Nano Res. 2022, 1–9. [Google Scholar] [CrossRef]
  18. Li, J.; Xia, W.; Tang, J.; Gao, Y.; Jiang, C.; Jia, Y.; Chen, T.; Hou, Z.; Qi, R.; Jiang, D. Metal–organic framework-derived graphene mesh: A robust scaffold for highly exposed Fe–N4 active sites toward an excellent oxygen reduction catalyst in acid media. J. Am. Chem. Soc. 2022, 144, 9280–9291. [Google Scholar] [CrossRef]
  19. Xiao, F.; Xu, G.L.; Sun, C.J.; Xu, M.; Wen, W.; Wang, Q.; Gu, M.; Zhu, S.; Li, Y.; Wei, Z. Nitrogen-coordinated single iron atom catalysts derived from metal organic frameworks for oxygen reduction reaction. Nano Energy 2019, 61, 60–68. [Google Scholar] [CrossRef]
  20. Liu, D.; Wang, B.; Srinivas, K.; Yu, B.; Chen, X.; Ma, F.; Wang, X.; Zhang, X.; Yang, D.; Chen, Y. Rich and uncovered FeNx atom clusters anchored on nitrogen-doped graphene nanosheets for highly efficient and stable oxygen reduction reaction. J. Alloys Compd. 2022, 901, 163763. [Google Scholar] [CrossRef]
  21. Song, A.; Cao, L.; Yang, W.; Yang, W.; Wang, L.; Ma, Z.; Shao, G. In situ construction of nitrogen-doped graphene with surface-grown carbon nanotubes as a multifactorial synergistic catalyst for oxygen reduction. Carbon 2019, 142, 40–50. [Google Scholar] [CrossRef]
  22. Cheng, Y.; He, S.; Lu, S.; Veder, J.P.; Johannessen, B.; Thomsen, L.; Saunders, M.; Becker, T.; De Marco, R.; Li, Q. Iron single atoms on graphene as nonprecious metal catalysts for high-temperature polymer electrolyte membrane fuel cells. Adv. Sci. 2019, 6, 1802066. [Google Scholar] [CrossRef] [Green Version]
  23. Zhang, H.; Chung, H.T.; Cullen, D.A.; Wagner, S.; Kramm, U.I.; More, K.L.; Zelenay, P.; Wu, G. High-performance fuel cell cathodes exclusively containing atomically dispersed iron active sites. Energy Environ. Sci. 2019, 12, 2548–2558. [Google Scholar] [CrossRef]
  24. Bai, X.; Guan, J. MXenes for electrocatalysis applications: Modification and hybridization. Chin. J. Catal. 2022, 43, 2057–2090. [Google Scholar] [CrossRef]
  25. Yang, X.; Feng, Z.; Guo, Z. Theoretical investigation on the hydrogen evolution, oxygen evolution, and oxygen reduction reactions performances of two-dimensional metal-organic frameworks Fe3(C2X)12 (X = NH, O, S). Molecules 2022, 27, 1528. [Google Scholar] [CrossRef]
  26. He, S.; Wu, M.; Li, S.; Jiang, Z.; Hong, H.; Cloutier, S.G.; Yang, H.; Omanovic, S.; Sun, S.; Zhang, G. Research progress on graphite-derived materials for electrocatalysis in energy conversion and storage. Molecules 2022, 27, 8644. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, Q.; Guan, J. Applications of single-atom catalysts. Nano Res. 2022, 15, 38–70. [Google Scholar] [CrossRef]
  28. Bai, X.; Guan, J. Applications of MXene-based single-atom catalysts. Small Struct. 2022, 2022, 2200354. [Google Scholar] [CrossRef]
  29. Mohideen, M.M.; Radhamani, A.V.; Ramakrishna, S.; Wei, Y.; Liu, Y. Recent insights on iron based nanostructured electrocatalyst and current status of proton exchange membrane fuel cell for sustainable transport. J. Energy Chem. 2022, 69, 466–489. [Google Scholar] [CrossRef]
  30. Zhang, X.; Truong-Phuoc, L.; Asset, T.; Pronkin, S.; Pham-Huu, C. Are Fe–N–C electrocatalysts an alternative to Pt-Based electrocatalysts for the next generation of proton exchange membrane fuel cells? ACS Catal. 2022, 12, 13853–13875. [Google Scholar] [CrossRef]
  31. Wang, X.X.; Swihart, M.T.; Wu, G. Achievements, challenges and perspectives on cathode catalysts in proton exchange membrane fuel cells for transportation. Nat Catal. 2019, 2, 578–589. [Google Scholar] [CrossRef]
  32. Song, Z.; Zhang, L.; Doyle-Davis, K.; Fu, X.; Luo, J.L.; Sun, X. Recent advances in MOF-derived single atom catalysts for electrochemical applications. Adv. Energy Mater. 2020, 10, 2001561. [Google Scholar] [CrossRef]
  33. Wang, B.; Wang, X.; Zou, J.; Yan, Y.; Xie, S.; Hu, G.; Li, Y.; Dong, A. Simple-cubic carbon frameworks with atomically dispersed iron dopants toward high-efficiency oxygen reduction. Nano Lett. 2017, 17, 2003–2009. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, J.; Li, Z.; Wu, Y.; Li, Y. Fabrication of single-atom catalysts with precise structure and high metal loading. Adv. Mater. 2018, 30, 1801649. [Google Scholar] [CrossRef]
  35. Zhan, Y.; Zeng, H.; Xie, F.; Zhang, H.; Zhang, W.; Jin, Y.; Zhang, Y.; Chen, J.; Meng, H. Templated growth of Fe/N/C catalyst on hierarchically porous carbon for oxygen reduction reaction in proton exchange membrane fuel cells. J. Power Sources 2019, 431, 31–39. [Google Scholar] [CrossRef]
  36. Cheng, R.; Jiang, M.; Li, K.; Guo, M.; Zhang, J.; Ren, J.; Meng, P.; Li, R.; Fu, C. Dimensional engineering of carbon dots derived sulfur and nitrogen co-doped carbon as efficient oxygen reduction reaction electrocatalysts for aluminum-air batteries. Chem. Eng. J. 2021, 425, 130603. [Google Scholar] [CrossRef]
  37. Dong, Y.; Fang, Z.; Ou, D.; Shi, Q.; Ma, Y.; Yang, W.; Tang, B.; Liu, Q. Rational fabrication of S-modified Fe–N–C nanosheet electrocatalysts for efficient and stable pH-universal oxygen reduction. Chem. Eng. J. 2022, 444, 136433. [Google Scholar] [CrossRef]
  38. Ross, R.D.; Sheng, H.; Ding, Y.; Janes, A.N.; Feng, D.; Schmidt, J.; Segre, C.U.; Jin, S. Operando elucidation of electrocatalytic and redox mechanisms on a 2D metal organic framework catalyst for efficient electrosynthesis of hydrogen peroxide in neutral media. J. Am. Chem. Soc. 2022, 144, 15845–15854. [Google Scholar] [CrossRef]
  39. Xu, X.; Xie, J.; Liu, B.; Wang, R.; Liu, M.; Zhang, J.; Liu, J.; Cai, Z.; Zou, J. PBA-derived FeCo alloy with core-shell structure embedded in 2D N-doped ultrathin carbon sheets as a bifunctional catalyst for rechargeable Zn-air batteries. Appl. Catal. B Environ. 2022, 316, 121687. [Google Scholar] [CrossRef]
  40. Hou, Y.; Lohe, M.R.; Zhang, J.; Liu, S.; Zhuang, X.; Feng, X. Vertically oriented cobalt selenide/NiFe layered-double-hydroxide nanosheets supported on exfoliated graphene foil: An efficient 3D electrode for overall water splitting. Energy Environ. Sci. 2016, 9, 478–483. [Google Scholar] [CrossRef] [Green Version]
  41. Hou, Y.; Qiu, M.; Zhang, T.; Zhuang, X.; Kim, C.S.; Yuan, C.; Feng, X. Ternary porous cobalt phosphoselenide nanosheets: An efficient electrocatalyst for electrocatalytic and photoelectrochemical water splitting. Adv. Mater. 2017, 29, 1701589. [Google Scholar] [CrossRef] [PubMed]
  42. Li, Z.; Chen, Y.; Ma, T.; Jiang, Y.; Chen, J.; Pan, H.; Sun, W. 2D metal-free nanomaterials beyond graphene and its analogues toward electrocatalysis applications. Adv. Energy Mater. 2021, 11, 2101202. [Google Scholar] [CrossRef]
  43. Niu, W.J.; He, J.Z.; Gu, B.N.; Liu, M.C.; Chueh, Y.L. Opportunities and challenges in precise synthesis of transition metal single-atom supported by 2D materials as catalysts toward oxygen reduction reaction. Adv. Funct. Mater. 2021, 31, 2103558. [Google Scholar] [CrossRef]
  44. Xiao, L.; Wang, Z.; Guan, J. 2D MOFs and their derivatives for electrocatalytic applications: Recent advances and new challenges. Coordin. Chem. Rev. 2022, 472, 214777. [Google Scholar] [CrossRef]
  45. Qiao, J.; Kong, L.; Xu, S.; Lin, K.; He, W.; Ni, M.; Ruan, Q.; Zhang, P.; Liu, Y.; Zhang, W. Research progress of MXene-based catalysts for electrochemical water-splitting and metal-air batteries. Energy Storage Mater. 2021, 43, 509–530. [Google Scholar] [CrossRef]
  46. Hu, C.; Zhou, Y.; Ma, R.; Liu, Q.; Wang, J. Reactive template synthesis of nitrogen-doped graphene-like carbon nanosheets derived from hydroxypropyl methylcellulose and dicyandiamide as efficient oxygen reduction electrocatalysts. J. Power Sources 2017, 345, 120–130. [Google Scholar] [CrossRef]
  47. Wang, C.; Liu, Y.; Li, Z.; Wang, L.; Niu, X.; Sun, P. Novel space-confinement synthesis of two-dimensional Fe, N-codoped graphene bifunctional oxygen electrocatalyst for rechargeable air-cathode. Chem. Eng. J. 2021, 411, 128492. [Google Scholar] [CrossRef]
  48. Yang, M.; Zhang, W.; Chen, H.; Li, H.; Liu, P. Medulla stachyuri-derived iron and nitrogen co-doped 2D porous carbon-flakes for highly efficient oxygen reduction electrocatalysis and supercapacitors. Int. J. Hydrogen Energy 2019, 44, 21726–21737. [Google Scholar] [CrossRef]
  49. Cao, X.; Li, J.; Dong, X.; Song, R.; Zhou, X.; Wang, X.; Yuan, N.; Ding, J. Self-template synthesized 2D porous carbon with FeA@ Fe-NC sites as oxygen electrocatalysts for Zn-Air and membraneless methanol fuel cells. J. Alloys Compd. 2022, 928, 166932. [Google Scholar] [CrossRef]
  50. Wang, M.; Gao, P.; Liu, J.; Li, D.; Yang, M.; Shen, Y.; Yang, S.; Hu, X.; Liu, Z.; Liu, Y. Layered-template synthesis of graphene-like Fe-NC nanosheets for highly efficient oxygen reduction reaction. Energy Fuel 2021, 35, 20349–20357. [Google Scholar] [CrossRef]
  51. Gu, B.; Su, H.; Chu, X.; Wang, Q.; Huang, H.; He, J.; Wu, T.; Deng, W.; Zhang, H.; Yang, W. Rationally assembled porous carbon superstructures for advanced supercapacitors. Chem. Eng. J. 2019, 361, 1296–1303. [Google Scholar] [CrossRef]
  52. Yang, W.; Yang, W.; Kong, L.; Song, A.; Qin, X.; Shao, G. Phosphorus-doped 3D hierarchical porous carbon for high-performance supercapacitors: A balanced strategy for pore structure and chemical composition. Carbon 2018, 127, 557–567. [Google Scholar] [CrossRef]
  53. Cai, H.; Zhang, G.; Zhang, X.; Chen, B.; Lu, Z.; Xu, H.; Gao, R.; Shi, C. Engineering the local coordination environment and density of FeN4 sites by Mn cooperation for electrocatalytic oxygen reduction. Small 2022, 18, 2200911. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, L.; Xiong, J.; Qin, Y.-H.; Wang, C.-W. Porous N–C catalyst synthesized by pyrolyzing g-C3N4 embedded in carbon as highly effcient oxygen reduction electrocatalysts for primary Zn-air battery. Carbon 2019, 150, 475–484. [Google Scholar] [CrossRef]
  55. Chen, H.; Liu, T.; Mou, J.; Zhang, W.; Jiang, Z.; Liu, J.; Huang, J.; Liu, M. Free-standing N-self-doped carbon nanofiber aerogels for high-performance all-solid-state supercapacitors. Nano Energy 2019, 63, 103836. [Google Scholar] [CrossRef]
  56. Ao, X.; Ding, Y.; Nam, G.; Soule, L.; Jing, P.; Zhao, B.; Hwang, J.Y.; Jang, J.H.; Wang, C.; Liu, M. A single-atom Fe-N-C catalyst with ultrahigh utilization of active sites for efficient oxygen reduction. Small 2022, 18, 2203326. [Google Scholar] [CrossRef]
  57. Jiao, L.; Wan, G.; Zhang, R.; Zhou, H.; Yu, S.H.; Jiang, H.L. From metal–organic frameworks to single-atom Fe implanted N-doped porous carbons: Efficient oxygen reduction in both alkaline and acidic media. Angew. Chem. Int.l Ed. 2018, 57, 8525–8529. [Google Scholar] [CrossRef]
  58. Subramanian, N.P.; Li, X.; Nallathambi, V.; Kumaraguru, S.P.; Colon-Mercado, H.; Wu, G.; Lee, J.W.; Popov, B.N. Nitrogen-modified carbon-based catalysts for oxygen reduction reaction in polymer electrolyte membrane fuel cells. J. Power Sources 2009, 188, 38–44. [Google Scholar] [CrossRef]
  59. Lai, L.; Potts, J.R.; Zhan, D.; Wang, L.; Poh, C.K.; Tang, C.; Gong, H.; Shen, Z.; Lin, J.; Ruoff, R.S. Exploration of the active center structure of nitrogen-doped graphene-based catalysts for oxygen reduction reaction. Energy Environ. Sci. 2012, 5, 7936–7942. [Google Scholar] [CrossRef]
  60. Tang, J.; Wang, J.; Shrestha, L.K.; Hossain, M.S.A.; Alothman, Z.A.; Yamauchi, Y.; Ariga, K. Activated porous carbon spheres with customized mesopores through assembly of diblock copolymers for electrochemical capacitor. ACS Appl. Mater. Inter. 2017, 9, 18986–18993. [Google Scholar] [CrossRef]
Scheme 1. The synthesis flow for Fe-N-CNSs.
Scheme 1. The synthesis flow for Fe-N-CNSs.
Molecules 28 02879 sch001
Figure 1. (a) SEM, (b) TEM and (c) HRTEM of Fe-N-CNSs. (d) HAADF-STEM and corresponding elemental mapping images of Fe-N-CNSs.
Figure 1. (a) SEM, (b) TEM and (c) HRTEM of Fe-N-CNSs. (d) HAADF-STEM and corresponding elemental mapping images of Fe-N-CNSs.
Molecules 28 02879 g001
Figure 2. (a) XRD patterns, (b) XPS surveys, (c) N 1s spectra, (d) Raman spectra, (e) The N2 adsorption-desorption isotherms and (f) the pore size distribution of N-CNSs, Fe-N-CNSs and Fe-N-C.
Figure 2. (a) XRD patterns, (b) XPS surveys, (c) N 1s spectra, (d) Raman spectra, (e) The N2 adsorption-desorption isotherms and (f) the pore size distribution of N-CNSs, Fe-N-CNSs and Fe-N-C.
Molecules 28 02879 g002
Figure 3. (a) CV curves of Fe-N-CNSs in O2/N2 saturated electrolyte. (b) LSV curves of the synthesized materials and Pt/C. (c) E1/2 together with jk@0.8 V towards these materials. (d) LSV curves of Fe-N-CNSs at different rotational speeds and the K-L plots. (e) Methanol tolerance and (f) i-t curves of Fe-N-CNSs and Pt/C.
Figure 3. (a) CV curves of Fe-N-CNSs in O2/N2 saturated electrolyte. (b) LSV curves of the synthesized materials and Pt/C. (c) E1/2 together with jk@0.8 V towards these materials. (d) LSV curves of Fe-N-CNSs at different rotational speeds and the K-L plots. (e) Methanol tolerance and (f) i-t curves of Fe-N-CNSs and Pt/C.
Molecules 28 02879 g003
Figure 4. (a) Schematic diagram of ZAB. (b) OCV of two ZABs. (c) The photograph of the LED panel powered by the Fe-N-CNSs-based ZAB. (d) Discharging LSV and power density curves, (e) discharge curves of two materials from 5 to 50 mA cm−2, and (f) the specific capacity of two ZABs.
Figure 4. (a) Schematic diagram of ZAB. (b) OCV of two ZABs. (c) The photograph of the LED panel powered by the Fe-N-CNSs-based ZAB. (d) Discharging LSV and power density curves, (e) discharge curves of two materials from 5 to 50 mA cm−2, and (f) the specific capacity of two ZABs.
Molecules 28 02879 g004
Table 1. Elemental composition of the catalysts measured by XPS.
Table 1. Elemental composition of the catalysts measured by XPS.
SampleC Content
(at %)
N Content
(at %)
O Content
(at %)
Fe Content
(at %)
N-CNSs87.533.219.260
Fe-N-CNSs88.83.797.20.21
Fe-N-C90.152.297.510.05
Table 2. The relative contents of various N existing forms were measured by XPS.
Table 2. The relative contents of various N existing forms were measured by XPS.
SamplePyridinic N
(at %)
Fe-N
(at %)
Pyrrolic N
(at %)
Graphitic N
(at %)
Oxidized N
(at %)
N-CNSs46.01034.6512.446.90
Fe-N-CNSs41.349.1728.1211.879.50
Fe-N-C27.366.0611.6435.3611.58
Table 3. BET surface area and BJH pore volume for Fe-N-CNSs-x (x = 800, 900 and 1000), N-CNSs and Fe-N-C.
Table 3. BET surface area and BJH pore volume for Fe-N-CNSs-x (x = 800, 900 and 1000), N-CNSs and Fe-N-C.
SampleBET
Surface Area (m2 g−1)
BJH
Pore Volume (cm3 g−1)
Fe-N-CNSs-800801.780.69
Fe-N-CNSs-9001303.431.76
Fe-N-CNSs-10001178.321.69
N-CNSs985.161.34
Fe-N-C421.150.57
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, S.; Lv, Y.; Elam, S.; Zhang, X.; Yang, Z.; Wu, X.; Guo, J. Rational Fabrication of Defect-Rich and Hierarchically Porous Fe-N-C Nanosheets as Highly Efficient Oxygen Reduction Electrocatalysts for Zinc-Air Battery. Molecules 2023, 28, 2879. https://doi.org/10.3390/molecules28072879

AMA Style

Li S, Lv Y, Elam S, Zhang X, Yang Z, Wu X, Guo J. Rational Fabrication of Defect-Rich and Hierarchically Porous Fe-N-C Nanosheets as Highly Efficient Oxygen Reduction Electrocatalysts for Zinc-Air Battery. Molecules. 2023; 28(7):2879. https://doi.org/10.3390/molecules28072879

Chicago/Turabian Style

Li, Sensen, Yan Lv, Sawida Elam, Xiuli Zhang, Zhuojun Yang, Xueyan Wu, and Jixi Guo. 2023. "Rational Fabrication of Defect-Rich and Hierarchically Porous Fe-N-C Nanosheets as Highly Efficient Oxygen Reduction Electrocatalysts for Zinc-Air Battery" Molecules 28, no. 7: 2879. https://doi.org/10.3390/molecules28072879

Article Metrics

Back to TopTop