Next Article in Journal
2D Nanomaterial—Based Electrocatalyst for Water Soluble Hydroperoxide Reduction
Next Article in Special Issue
Giving New Life to Waste Cigarette Butts: Transformation into Platinum Group Metal-Free Electrocatalysts for Oxygen Reduction Reaction in Acid, Neutral and Alkaline Environment
Previous Article in Journal
Selective Oxidation of Furfural at Room Temperature on a TiO2-Supported Ag Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

MIL-88-Derived N and S Co-Doped Carbon Materials with Supplemental FeSx to Enhance the Oxygen Reduction Reaction Performance

1
College of Chemistry and Materials Science, Inner Mongolia Minzu University, Tongliao 028000, China
2
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Zhongshan Road 457, Dalian 116023, China
3
Dalian Leicester Institute, Dalian University of Technology, Dalian 116024, China
4
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Zhongshan Road 457, Dalian 116023, China
5
Liaoning Key Laboratory of Plasma Technology, School of Physics and Materials Engineering, Dalian Minzu University, Dalian 116600, China
6
Shanghai Key Laboratory of Green Chemistry and Chemical Processes, School of Chemistry and Molecular Engineering, East China Normal University, Shanghai 200062, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(8), 806; https://doi.org/10.3390/catal12080806
Submission received: 6 June 2022 / Revised: 8 July 2022 / Accepted: 18 July 2022 / Published: 22 July 2022
(This article belongs to the Special Issue Non-noble Metal Electrocatalysts for the Oxygen Reduction Reaction)

Abstract

:
To overcome the drawbacks of the single N-doped carbon materials, the further development of dual-heteroatoms (N and S) co-doped electrocatalysts is highly anticipated. Herein, N, S-doping and Fe-based carbon materials were synthesized by pyrolyzing a metal–organic framework (MIL-88) with the addition of N-/N, and S-containing ligands (chitosan and L-Cysteine) in the case of iron salt. The resulting electrocatalyst heat-treated at 850 °C (FeNSC-850) displays superior oxygen reduction reaction (ORR) performances to MIL-88-850, with an overall electron transfer number of 3.97 and a minor yield of HO2- (<2.6%). In addition to the comparable activity to commercial Pt/C in catalyzing the ORR in alkaline solution, the FeNSC-850 also shows higher stability, with a slight decline in half-wave potential (∆E1/2 = 15 mV) after 5000-cycle scanning of cyclic voltammetry. In view of the multiple Fe-based active sites, the additional S doping within FeNSC-850 creates more FeSx active sites for boosting the ORR performances in alkaline solution.

1. Introduction

As next-generation energy conversion and storage devices, fuel cells (FCs) are becoming appealing owing to their high energy density and environmental friendliness [1,2]. The overall efficiency of FCs is determined by the progress of the oxygen reduction reaction (ORR). To date, Pt was commonly used as the best ORR electrocatalysts, whereas the drawbacks of Pt, like high-cost and low-reserve, are topical for a large scale for FCs. Thus, it is inevitable to exploit the alternatives to Pt for FCs. The non-noble metal electrocatalysts (NNMEs), especially for N-doping and Fe-based carbon materials, hitherto have attracted much attention because of their excellent electrochemical catalytic properties for ORR in both alkaline and acidic solution [3,4], while, the N dopant with strong electronegativity will undesirably perturb the electronic properties of metal-based active sites and increase the free energy for the adsorption and activation of the intermediates towards the ORR, restricting the promotion of the intrinsic nature of NNMEs for ORR [5,6,7]. Thus, a secondary heteroatom dopant, such as S or P, is acceptable to optimize the microenvironment of metal active centers [8]. It has been noted that N, S co-doping is superior to N, P or N, B co-doping in Fe-based NNMEs towards the ORR [9].
Recently, FeNSC electrocatalysts with a second heteroatom of S that exhibit superb ORR performances due to S that can induce charge enrichment in N and Fe for optimal O2 binding and fast electron transfer [10]. A series of N, S co-doping and transition metal-based carbon materials has been investigated on both experiments and theoretics for the ORR. Examples include, but are not limited to N, S co-doping and Fe-based porous carbon nanosheets [11,12,13,14]; N, S co-doping and Co/Fe-based carbon nanotubes [15,16]; N, S co-doping and Fe-based hierarchically porous carbon nanorods [5]; N, S co-doping and Co-based hollow carbon spheres [8]; N, S co-doping and Fe-based hierarchically porous honeycomb [17]; and N, S co-doping and Fe-based blocks [18]. All these electrocatalysts demonstrate considerable ORR performances in both alkaline and acidic solution.
In particular, metal–organic frameworks (MOFs) have been considered as ideal substrates due to facile synthetic pathways, their large porosity, and structural versatility after pyrolysis [19,20,21]. Among them, the MIL-88 framework could be chosen as a sacrificial template for a high graphitized structure [22,23]. As reported, iron carbides possess good ORR performances in alkaline solution [24], while the Fe3C derived from the carbonization of MIL-88 shows limited ORR performances, which can be improved by the exposure of dense Fe-based active sites through modifications on MIL-88 [23]. Herein, to address this conventional drawback of MIL-88, we innovatively utilized the synergistic effect of N and S dopants by annealing MIL-88 and additional Fe salt with the assistance of chitosan and L-Cysteine (L-Cy). Especially, chitosan is a low-cost and abundant biomass-derived ligand with strong coordination with metal, and L-Cy is a N, S-containing amino acid that is commonly used to introduce S in electrocatalysts for the ORR [25]. Consequently, the simultaneous introduction of N and S doping in carbon skeletons improves the stability of Fe by the formation of FeNx, FexC, and FeSx to boost the ORR performances in alkaline solution. Combining the porosity structure; N, S-dopants; and multiple Fe-based active species, the resulting electrocatalyst of FeNSC-850 shows excellent ORR performances via a direct four-electron transfer in 0.1 M KOH solution.

2. Results and Discussion

As depicted in Scheme 1, the N and S co-doping and iron-based carbon material was prepared by carbonizing the mixture of MIL-88, Fe salt, chitosan, and L-Cy at 850 °C. In this synthetic system, chitosan and L-cy were introduced as N-/N, S-containing biochemical precursors. During the pyrolysis step, N and S sources could bind with Fe to create more active sites, such as iron-based nitrides, carbides, and sulfides within carbon materials. Consequently, the porous N, S-co-doping and iron-based carbon materials, denoted as FeNSC-850, were designed and synthesized with excellent catalytic activities towards ORR in alkaline solution.
The morphology of MIL-88-850 in Figure S1a,b (in the Supplementary Materials) and FeNSC-850 (Figure 1a,b) are characterized by the scanning electron microscopy (SEM). After the carbonization of MIL-88 at 850 °C, a spindle-like structure of MIL-88-850 is observed in Figure S1a (in the Supplementary Materials). The enlarged SEM image in Figure S1b (in the Supplementary Materials) shows multiple pores on the surface of spindle-like structures after the pyrolysis process. With the addition of doping materials on MIL-88 in Figure 1a, the morphology of FeNSC-850 changed comparable with the starting templet of MIL-88. FeNSC-850 shows a hierarchical porous structure, including macro-, meso-, and micropores in the enlarged SEM image in Figure 1b. The energy-dispersive X-ray spectroscopy (EDX) pattern in Figure S2 (in the Supplementary Materials) indicates the presence of N, C, S, and Fe in FeNSC-850. As detected in Figure 1c, a high angle annular dark-field scanning TEM (HAADF-STEM) image and the element mapping of FeNSC-850 verify that C, N, and O are uniformly distributed throughout the whole structure, indicating the successful N doping within carbon matrixes. The overlap of S and C in element mapping clearly confirms the doping of S in carbon matrices. Simultaneously, the mapping results of Fe and S are superimposed, which indicates the sulfidation of Fe species encapsulated within carbon matrixes.
The nature of samples was further analyzed by X-ray diffraction (XRD), as shown in Figure 2a and Figure S3 (in the Supplementary Materials). It is turned out that the peaks of as-prepared electrocatalysts appearing around 25° and 44° are indexed to (002) and (101)/(100) planes of graphite carbon [26]. A sharper and narrower peak around 26.4° of FeNSC-850 can be ascribed as Fe3C (PDF #03−0411) [27]. Notably, the apparent diffraction peaks of MIL-88-850 at 42.9, 43.7, 44.6, 44.9, 45.9, 48.6, and 49.1° correspond to (211), (102), (220), (031), (112), (131), and (221) planes of Fe3C (PDF #35−0772) in Figure S3a (in the Supplementary Materials) [28]. In comparison, the well-defined diffraction peaks of FeNSC-850 around 33.1, 37.1, 40.8, 47.5, 50.3, and 56.3° probably arise from the (200), (210), (211), (220), (221), and (311) planes of FeS2 (PDF #42−1340) [29]. Moreover, the XRD patterns at around 20.8° and 50.3° could be linked to the (101) and (103) planes of FeS (PDF #76−0965). Accordingly, combined with the element-mapping results of Fe and S (Figure 1c), it is clear that the addition of L-Cy promotes the conversion of Fe ions to FeSx. The appearance of FeS2 in electrocatalysts not only promotes the smooth transfer of electrons to active sites but also strengthens the conductivity of electrocatalysts [30,31]. The Raman spectra of MIL-88-850 and FeNSC-850 in Figure S4 (in the Supplementary Materials) show two sharp peaks at around 1350 and 1600 cm−1, corresponding to defects (D band) and the planar motion of sp2-carbon atoms in graphite carbon (G band) [32]. The ratio of the D and G band (ID/IG) characterizes the values on MIL-88-850 (ID/IG = 0.899) and FeNSC-850 (ID/IG = 0.854), indicating the higher carbon graphitization of FeNSC-850.
The porosity of as-synthesized electrocatalysts was characterized by N2 adsorption-desorption analysis in Figure 2b,c. Both FeNSC-850 and MIL-88-850 possess a type-IV isotherm in Figure 2b. This suggests the existence of mesopores in the electrocatalysts, which is detected by SEM in Figure 1 and confirmed by pore size distribution shown in Figure 2c. The Brunauer–Emmett–Teller (BET) surface area (SBET = 476 m2 g−1) of FeNSC-850 is similar to that of MIL-88-850 (SBET = 483 m2 g−1). The microporous surface area (SMicro = 291 m2 g−1) of FeNSC-850 with a volume of micropore (VMicro) of 0.15 cm3 g−1 by t-Plot is significantly larger than that of MIL-88-850 (SMicro = 97 m2 g−1, VMicro = 0.05 cm3 g−1 by t-Plot), revealing more accommodation for hosting a high density of Fe-based active species for the enhancement of ORR activities. The release of NH3 and H2S from L-Cy at high pyrolysis temperature leads to a larger SMicro and VMicro, which is believed to be a major benefit for oxygen diffusion and mass transport to boost catalytic activities towards ORR [33]. Additionally, the formation of FeNx moieties has been found to be more favorable in micropores in carbon matrixes. As shown in Figure 2c, the corresponding pore size of MIL-88-850 centers around 2.45 and 16.1 nm. In contrast, the size of the partial pore of FeNSC-850 decreases to around 1.74 nm, and the others are larger pores with a very wide pore size distribution. The observed meso- and macropores are constructed by removing the NPs after the acid-leaching step. In conclusion, the carbonization of N-/N, S-containing ligands can enhance the porosity of FeNSC-850 and produce a larger proportion of micropores for better ORR enhancement.
X-ray photoelectron spectroscopy (XPS) analyses were subsequently employed to characterize the surface electronic states of as-prepared electrocatalysts. The survey spectra of MIL-88-850 and FeNSC-850 were shown in Figure S5 (in the Supplementary Materials), indicting the presence of N, C, and O for MIL-88-850 and N, C, O, and S for FeNSC-850. There is not an obvious signal of Fe in both electrocatalysts, possibly due to the limited detect of XPS for Fe NPs encapsulated within thicker carbon layers. Specifically, the high-resolution C 1s spectrum of FeNSC-850 in Figure 3a could be deconvoluted into four peaks at 284.8, 285.8, 287.0, and 290.4 eV, respectively, corresponding to C=C-C (sp2 graphitic carbon), C=N/C-O/C-S, C-N/C-O-C, and O-C=O, indicating that the heteroatoms (O, N, and S in our case) have been doped in the carbon skeleton [34]. The N 1s XPS spectra in Figure 3b could be categorized into five peaks after fitting at 398.4, 399.3, 401.0, 402.3, and 406.0 eV, respectively, correlated to pyridinic N, Fe-N, pyrrolic N, graphitic N, and oxidized N. It is known that pyridinic N, Fe-N, and graphitic N could contribute to promote the ORR in an alkaline solution. Particularly, graphitic N is beneficial to raise the conductivity of materials by influencing the electronic and geometric structures of carbon [35]. Since the content of the above-mentioned five types of N varies with pyrolysis temperature, the five nitrogen contents (relative to C) in different FeNSCs synthesized at four temperatures and MIL-88-850 are illustrated in Figure 3c and Figure S6 (in the Supplementary Materials). Notably, the graphitic N in FeNSC-850 is higher than samples pyrolyzed at other temperatures. In particular, the N content of FeNSC-850 (relative to C) is about 3.6 at%, which is higher than that of MIL-88-850 (1.8 at%). It confirms that the additional N-containing ligands indeed increase the N-dopant content of electrocatalysts. The high-resolution S 2p spectra of FeNSC-850 in Figure 3d shows that the strong peaks at 164.1 and 165.3 eV correspond to S 2p3/2 and S 2p1/2 of thiophene-like structures (C-S-C), respectively. The two small peaks at the binding energy of 168.5 and 169.7 eV arise from the oxidized S species, namely, -C-SOx-C bonds [1]. Introducing S plays a positive role in the Fe-based microenvironment that promotes the interactions between Fe active sites and oxygenated reactive intermediates in the ORR process [14]. More precisely, the observed C-S-C species are reported to contribute to more effective facilitation towards the ORR than -SOx. [36]. The oxidation state of the surface iron is interpreted from the spectra characteristic peaks in Figure 3e. As a result, the peaks centered at around 711.3 and 714.1 eV are from Fe2+ and Fe3+, respectively. It is found that the positive shift of the high binding energy from 710 to 711 eV is due to heteroatoms (N and S in our case) that decrease the charge density on Fe [33]. Obviously, the binding energy at around 707.3 eV is attributed to FeS2 [37], which is in line with the result of XRD patterns in Figure 2a. Moreover, the Fe content of MIL-88-850 is determined by inductively coupled plasma optical emission spectroscopy (ICP-OES) as 9.34 wt%. Additionally, the Fe content of FeNSC-850 is determined by ICP as 3.37 wt%. Obviously, FeNSC-850 lost some Fe content during the high-temperature pyrolysis and acid-leaching. As shown in the SEM images in Figure 1 and Figure S1 (in the Supplementary Materials), the structure of MIL-88-850 is maintained, while some of the structure of FeNSC-850 collapsed; thus, some Fe-based NPs leached out and Fe content decreased.
The 57Fe Mössbauer spectroscopy is further used to identify the chemical environment of iron in the electrocatalysts (Figure 4a,b and Table 1). Distinctly, the spectrum of MIL-88-850 in Figure 4a is composed of two typical components, including a doublet ascribed to Fe2+N4 (D1, 23.34%) in the hexagonal symmetry site and a sextet (S1) assigned to Fe3C (S1, 76.66%). The spectrum of FeNSC-850 in Figure 4b is composed of six sextets and two doublets. The fitted Mössbauer parameters in Table 1 indicate that a sextet is assigned to FexC (S2, 11.22%). Three sextets, including S3, S4, and S5, correspond to the various iron sulfides denoted as FeSx (61.73%). Two doublets of D2 and D3 are assigned to FeS2 (22.05%) and Fe2+N4 (5.01%) in the tetrahedral symmetry site, respectively. In contrast to FeN4 (hexagonal symmetry site) in MIL-88-850, the addition of N-containing ligands plays an important role in controlling the insertion of N into the lattice of Fe matrixes such as FeN4 (tetrahedral symmetry site) for different folds as the coordination formation in FeNSC-850. Moreover, the S-containing ligand, namely, L-Cy in FeNSC-850 controls the insertion of S into the lattice of Fe matrixes for supplemental FeSx, which may alter the surface electron configuration for the adsorption/activation of O2 and further provide networks for efficient electron transfer [1]. As a result, FeSx is regarded to be the main Fe-based active sites in FeNSC-850.
Generally, highly efficient NNMEs retained at moderating pyrolysis temperature are due to the balanced properties as for the graphitic structure, the surface area, and the obtained active species, such as C-S, C-N, Fe-N, and Fe-S. [14] The electrochemical study of FeNSCs obtained from various pyrolysis temperatures (FeNSC-T) in Figure 5a was conducted in 0.1 M KOH solution at a rotation rate of 1600 rpm. For clarity, the potentials of electrocatalysts in Figure 5a are summarized and contrasted in Figure 5b at a representative current density of −2.5 mA cm−2. The ORR performance of FeNSC-850 is enhanced relative to those of FeNSC-750 and FeNSC-800. Remarkably, a plummet of activity is observed from FeNSC-900, which can be explained by the decrease of N at% (N relative to C, 1.7 at%) in Figure 3c. Further treatment with a higher temperature probably wouldresult in lower yield of electrocatalyst. Thus, we did not try any other higher temperature over 900 °C for heating the samples. Consequently, FeNSC-850 displays the best ORR performances in terms of the onset potential (Eonset = 0.89 V vs. RHE) and the half-wave potential (E1/2 = 0.82 V vs. RHE) among FeNSC-T electrocatalysts, which is comparable to those of the commercial Pt/C with different Pt loadings (10 µgPt cm−2: Eonset = 0.92 V and E1/2 = 0.84 V vs. RHE; 20 µgPt cm−2: Eonset = 0.94 V and E1/2 = 0.85 V vs. RHE).
The addition of chitosan and L-Cy introduces more N and S elements within carbon matrices during high-temperature pyrolysis process. To investigate the aid of Fe salt, chitosan, and L-Cy in electrocatalysis for ORR, three groups of electrocatalysts without the addition of either Fe salt, chitosan, or L-Cy were synthesized in parallel with FeNSC-850. As shown in Figure 5c, all three as-prepared materials (where chitosan:L-Cy = 0:1, chitosan:L-Cy = 1:0, and the electrocatalyst without additional Fe salt with mass ratio of chitosan and L-Cy as 1:1) apparently present reduced activity compared to FeNSC-850 (where chitosan:L-Cy = 1:1), which confirms the importance of the synergistic effect of N and S elements within carbon matrixes at 850 °C and the necessity of Fe by coordination with N or S as the high-density of active sites for boosting the ORR performances. This confirms that chitosan and L-Cy have a synergistic effect for preparing highly efficient electrocatalysts by increasing the active sites for ORR. According to the reported studies, doping electrocatalysts solely with sulfur actually would dampen their activity towards ORR. S-doping can, however, induce defects in the carbon skeleton and enrich the charge on adjacent Fe and N, thereby accumulating more electrons on adsorbed O2 and facilitating the reductive release of *OH to enhance the ORR performance [42,43]. Thus, chitosan plays an important role in increasing the content of N dopants in electrocatalysts. Afterwards, supplemental FeSx active sites provided by the L-Cy could act as a booster for ORR. Generally, the superior ORR activity of FeNSC-850 is attributed to the co-function of Fe-based active sites and N, S co-doped carbon matrixes [44].
The LSV curves of FeNSC-850 in Figure 5d demonstrate that the increasing limited diffusion current densities increase along with the increasing rotation rates ranging from 400 to 2500 rpm. Accordingly, the Koutecky–Levich (K-L) plots of FeNSC-850 in Figure 5e derived from the data in Figure 5d suggest that the electron transfer number (n) of FeNSC-850 is calculated to be 3.92 at potentials from 0.6 to 0.8 V (vs. RHE). To further investigate the electron transfer number and HO2 intermediate, rotating ring-disk electrode (RRDE) measurements were carried out with FeNSC-850, MIL-88-850, and 20 wt% commercial Pt/C (Figure 5f) in O2-saturated 0.1 M KOH. The results show that the average n of FeNSC-850 from 0 to 0.8 V (vs. RHE) is calculated to be 3.97, which is higher than that of MIL-88-850 (average n = 3.26) and very close to the value (3.92) calculated from K-L plots. The yield of HO2- intermediate is about 41% for MIL-88-850 at 0.47 V (vs. RHE), suggesting that there exists a two-electron ORR pathway from O2 to HO2-, mainly due to the low overall content of N and less Fe-based active sites. The HO2- yield of FeNSC-850 (<2.6%) is comparable to that of 20 wt% Pt/C (HO2% < 1.1%) at 0.47 V (vs. RHE). Together with the electron transfer number, which is comparable to the commercial Pt/C electrocatalyst (average n = 3.96), it indicates that FeNSC-850 displays a four-electron ORR pathway in 0.1 M KOH solution.
To further evaluate the durability of FeNSC-850 and 20 wt% Pt/C, cyclic voltammetry (CV) analysis was employed by scanning 5000 CV cycles from 0.5 to 1.1 V (vs. RHE) at 100 mV s−1. As shown in Figure 6a, the E1/2 of FeNSC-850 decreases by only 15 mV, which is less than that of 20 wt% Pt/C (E1/2 loss = 63 mV). In comparison with others’ work summarized in Table 2, our as-prepared NNMEs exhibit comparable or even superior ORR performances considering their catalytic ORR performances and stability. Furthermore, beneficial for practical fuel cell applications, chronoamperometric (I-t) measurements were conducted to assess the methanol tolerance of FeNSC-850 and the commercial Pt/C by the addition of 10 vol% methanol in O2-saturated 0.1 M KOH solution at around 400 s. The normalized current density in Figure 6b suggests that unlike Pt/C, FeNSC-850 displays outstanding methanol tolerance, whereas Pt/C presents a substantial decline at the point of injection of methanol in O2-saturated 0.1 M KOH media at around 400s.
To the best of our knowledge, the possible reasons for the efficient ORR catalytic activity of FeNSC-850 can be described as follows: (1) The large surface area with abundant microporous architecture guarantees the efficient mass transportation during the ORR process. (2) The higher content of graphitic N is more favorable in the N, S-doped carbon materials [45]. (3) The S-doping may change the microenvironment of iron-based active species during the pyrolysis step. (4) The S-S bonds in the main active sites of FeSx could provide electrons to the iron nitrides to promote the adsorption and reduction of O2 through four electron ORR pathway [30].

3. Materials and Methods

3.1. Chemical Materials

Fe(NO3)3·9H2O and NH4Fe(SO4)2·12H2O were purchased from Nanjing Chemical Reagent Co., Ltd. (Nanjing, China) and Shanghai Hushi Laboratorial Equipment Co., Ltd. (Shanghai, China). 2-Amino terephthalic acid was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China) 2-Amino-3-mercaptopropanoic acid (L-Cysteine) was obtained from Bide Pharmatech (Shanghai, China) Ltd. N, N-dimethyl-formamid (DMF) was purchased from Tianjin Dongli Tianda Chemical Reagent Plant (Tianjin, China). Methanol and ethanol were purchased from Shenyang East China Reagent Plant (Shenyang, China) and China National Pharmaceutical Industry Co., Ltd. (Beijing, China) All chemical reagents in this work were used without any further purification. Ultrapure water with the resistance of 18.2 MΩ cm (at 25 °C) was used for all experiments and electrochemical properties.

3.2. Synthesis of MIL-88-850

MIL-88 with some modifications was prepared in light of the work of a report [52]. Typically, 0.50 g of 2-aminoterephthalate was firstly dissolved in 10 mL N, N-dimethylformamide with a vigorous stirring. Secondly, 1.12 g Fe(NO3)3·9H2O was added in the above solution with further stirring for 2 h. Afterwards, the mixture was transferred into a Teflon-lined stainless steel autoclave and heat-treated at 160 °C for 6 h. Subsequently, the mixture was centrifuged and washed with DMF and methanol. The resulting sample was obtained after drying in a vacuum at 60 °C for 24 h. The obtained MIL-88 was directly pyrolyzed at 850 °C for 2 h with a ramp rate of 10 °C min−1 in an Ar atmosphere. Afterwards, the obtained sample was acid-leached with 0.5 M H2SO4 overnight and washed with water until the filtrate became neutral. Finally, the sample was dried and denoted as MIL-88-850.

3.3. Synthesis of FeNSC-850

The obtained 1 g MIL-88 was dispersed into a 50 mL solution containing 25 mL water and 25 mL ethanol. Then, the certain mass ratio (1:1) of chitosan (500 mg) and L-Cy (500 mg), and 48 mg NH4Fe(SO4)2·12H2O, was further added into the above solution with stirring at 60 °C for 3 h, followed by drying in the oven at 60 °C overnight. The obtained mixture was then placed closely at the middle of the furnace and heated at 850 °C for 2 h with a ramp rate of 10 °C min−1 in an Ar atmosphere. Finally, the samples were leached with 0.5 M H2SO4 overnight and washed with water until the filtrate reached neutrality. Finally, the sample was dried and marked as FeNSC-850. For comparison, reference samples with a certain mass ratio of chitosan and L-Cy (1:0 and 0:1) and the sample without the addition of Fe salt were synthesized under the same conditions.

3.4. Characterization

The specific surface area and the corresponding pore size distribution of as-synthesis materials were calculated by means of the Brunauer–Emmett–Teller theory and were determined on the adsorption isotherm by Barrett–Joyner–Halenda (BJH) models. TEM and SEM were carried out on TECNAI G2F30 (FEI Company, USA) and S-4800 (Hitachi, Tokyo, Japan), respectively. XRD was conducted by X’pert Pro-1 (PANAlytical, Almelo, Netherlands) from 10 to 90°. The XPS of samples was conducted on Escalab Xi+ (Thermo Fisher Scientific, Winsford, UK). The Mössbauer spectra were collected at the ambient environment using the 57Co radioactive source and Topologic 500 A spectrometer. The Fe content of the electrocatalysts was measured by inductively coupled plasma optical emission spectroscopy (ICP-OES) (ICPS-8100, Shimadzu, Japan). Raman spectra were recorded on a Raman spectroscopy (Bruker Optics Senterra, Leipzig, Germany) with 532 nm excitation laser.

3.5. Electrochemical Measurements

The electrochemical properties, including the cyclic voltammetry (CV), the rotating disk electrode (RDE), the rotating ring-disk electrode (RRDE), and the I-t tests, were evaluated with a three-electrode system using CHI 760E (Chenhua, Shanghai, China) and VSP-300 (BioLogic, Paris, France). In this three-electrode system, the Hg/HgO and a graphite rod were used as the reference and counter electrode, respectively. In this paper, all the potential values were calibrated to the reversible hydrogen electrode (RHE) as ERHE = EHg/HgO + 0.884 V. The calibration method for RHE follows the reported work [53]. The as-prepared electrocatalyst film-coated RDE and RRDE were utilized as the working electrodes. The electrocatalyst ink was prepared by dispersing 7.7 mg electrocatalysts in a solvent mixed with ethanol (900 μL), nano-pure water (100 μL), and Nafion (4.62 µL, 5 wt%). Afterwards, the electrocatalyst ink (20.4 μL) was dropped onto the polished glassy carbon (GC) on RDE. By air-drying at room temperature, the obtained electrocatalyst loading on GC was calculated to be 0.8 mg cm−2. In contrast, commercial Pt/C was prepared as 1 mg mL−1, and 10 and 20 μL were further coated on the GCs to obtain 10 and 20 μgPt cm−2.
First of all, CV measurements were assessed at the potential of 0 to 1.1 V (vs. RHE) at a scanning rate of 100 mV s−1, after which the surface of electrocatalyst could be further activated on GC. The ORR performances were conducted at 1600 rpm using a scanning rate of 5 mV s−1. Both the RDE and RRDE tests were conducted in O2 and N2-saturated alkaline solution. Moreover, the RRDE technique was further determined at a consistent potential of 1.1 V (vs. RHE). To further discuss the electron transfer number, the Equations (1)–(4) were used as follows:
1 J = 1 J K + 1 B ω 1 2
B = 0.62 n F D o 2 / 3 υ 1 / 6 C o
In the Equation (1), J and JK are the current density and kinetic current density. In the Equation (2), n is the electron transfer number. Co (concentration of dissolved O2) is 1.2 × 10−6 mol cm−3. Do (diffusion coefficient of O2) is 1.9 × 10−5 cm2 s−1. υ is 0.01 cm2 s−1. F (Faraday constant) is 96,485 C mol−1.
n =   4 × I D I R N + I D
H O 2 % = 200 × I R N I R N + I D
ID and IR are the disk current and ring current, respectively. N is the current collection efficiency of 37%.
Normally, the half-wave potential (E1/2) and onset potential (Eonset) are used to evaluate the ORR performances of the obtained electrocatalysts. Here in this work, all E1/2 and Eonset were detected by the methods reported in our previous work [54].
Moreover, durability tests were conducted for 5000 CV cycles in the potential of 0.5 to 1.1 V vs. RHE at 100 mV s−1 in O2-saturated 0.1 M KOH media. The LSV curves were further conducted on the FeNSC-850 and Pt/C before and after 5000 cycles. Moreover, the I-t test of FeNSC-850 and commercial Pt/C were further assessed in the O2-saturated 0.1 M KOH solution with the addition of methanol at around 400 s.

4. Conclusions

To summarize, we reported a synthetic route based on an MOF (MIL-88) template with property modification by N-/N, S-containing ligands and iron salt. After pyrolysis at 850 °C and acid leaching, the resulting FeSNC-850 owns a tremendous improvement in ORR performance. We also discovered that this improvement profits from the synergic effect between the supplemental FeSx sites generated by this method and the Fe-based active sites. The catalytic activity of FeNSC-850 towards ORR in alkaline solution is comparable to that of Pt/C, but the former also presents more durable properties as well as functional stability in a methanol environment. This work investigates FeSx as the main active site within multiple Fe-based active sites in ORR electrocatalysts.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal12080806/s1, Figure S1: the SEM images of (a) and (b) MIL-88-850; Figure S2: the EDX of FeNSC-850m; Figure S3: the XRD patterns of (a) MIL-88-850, (b) FeNSC-750, (c) FeNSC-800, and (d) FeNSC-900; Figure S4: the Raman spectra of (a) FeNSC-850 and (b) MIL-88-850; Figure S5: the survey spectra of (a) MIL-88-850 and (b) FeNSC-850; and Figure S6: the N 1s XPS spectra of (a) MIL-88-850, (b) FeNSC-750, (c) FeNSC-800, and (d) FeNSC-900.

Author Contributions

Investigation, methodology, and formal analysis, Y.L. and Y.X. (Yinghao Xu); methodology, H.Z.; characterization, Y.X. (Yinghao Xu) and J.Z.; characterization, H.W.; morphology, L.C.; conceptualization, writing—original draft preparation, and funding acquisition, Y.X. (Yan Xie); supervision, writing—review and editing, and funding acquisition, L.X.; and supervision, writing—review, and editing, J.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, No. 21606219, and Natural Science Foundation of Inner Mongolia, No. 2021LHMS02006.

Data Availability Statement

All data that support the findings of this study are included within the article (and any Supplementary Files).

Acknowledgments

We gratefully acknowledge support from Junhu Wang, the Mössbauer Effect Data Centre, and the Dalian Institute of Chemical Physics for the discussion on the Mössbauer spectra.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fang, H.; Huang, T.; Mao, J.; Yao, S.; Dinesh, M.M.; Sun, Y.; Liang, D.; Qi, L.; Yu, J.; Jiang, Z. Investigation on the Catalytic Performance of Reduced-Graphene-Oxide-Interpolated FeS2 and FeS for Oxygen Reduction Reaction. ChemistrySelect 2018, 3, 10418–10427. [Google Scholar] [CrossRef] [Green Version]
  2. Liu, Y.-L.; Liu, J.; Wang, M.; Shen, Y.; Yang, S.; Li, D. Mesoporous Fe-Nx-C Sub-Microspheres for Highly Efficient Electrocatalytic Oxygen Reduction Reaction. ChemCatChem 2021, 13, 4047–4054. [Google Scholar]
  3. Zhu, C.; Shi, Q.; Xu, B.Z.; Fu, S.; Wan, G.; Yang, C.; Yao, S.; Song, J.; Zhou, H.; Du, D.; et al. Hierarchically Porous M-N-C (M = Co and Fe) Single-Atom Electrocatalysts with Robust MNx Active Moieties Enable Enhanced ORR Performance. Adv. Energy Mater. 2018, 8, 1801956. [Google Scholar] [CrossRef]
  4. Li, Y.; Hu, H.; Song, J.; Wu, Y.; Lv, X.; Xiao, Z.; Wang, F.; Chen, Y. An Excellent Fe, N Co-Doped Porous Biomass Carbon Oxygen Reduction Reaction Electrocatalyst: Effect of Zinc-Based Activators on Catalytic Activity. Energy Technol. 2020, 8, 2000625. [Google Scholar] [CrossRef]
  5. Liu, F.; Shi, L.; Song, S.; Ge, K.; Zhang, X.; Guo, Y.; Liu, D. Simultaneously Engineering the Coordination Environment and Pore Architecture of Metal-Organic Framework-Derived Single-Atomic Iron Catalysts for Ultra Efficient Oxygen Reduction. Small 2021, 17, 2102425. [Google Scholar] [CrossRef] [PubMed]
  6. Zhou, P.; Li, N.; Chao, Y.; Zhang, W.; Lv, F.; Wang, K.; Yang, W.; Gao, P.; Guo, S. Thermolysis of Noble Metal Nanoparticles into Electron-Rich Phosphorus-Coordinated Noble Metal Single Atoms at Low Temperature. Angew. Chem. Int. Ed. 2019, 58, 14184–14188. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, K.; Liu, K.; An, P.; Li, H.; Lin, Y.; Hu, J.; Jia, C.; Fu, J.; Li, H.; Liu, H.; et al. Iron Phthalocyanine with Coordination Induced Electronic Localization to Boost Oxygen Reduction Reaction. Nat. Commun. 2020, 11, 4173. [Google Scholar] [CrossRef]
  8. Zhang, Z.; Zhao, X.; Xi, S.; Zhang, L.; Chen, Z.; Zeng, Z.; Huang, M.; Yang, H.; Liu, B.; Pennycook, S.J.; et al. Atomically Dispersed Cobalt Trifunctional Electrocatalysts with Tailored Coordination Environment for Flexible Rechargeable Zn–Air Battery and Self-Driven Water Splitting. Adv. Energy Mater. 2020, 10, 2002896. [Google Scholar] [CrossRef]
  9. Sahraie, N.R.; Paraknowitsch, J.P.; Göbel, C.; Thomas, A.; Strasser, P. Noble-Metal-Free Electrocatalysts with Enhanced ORR Performance by Task-Specific Functionalization of Carbon using Ionic Liquid Precursor Systems. J. Am. Chem. Soc. 2014, 136, 14486–14497. [Google Scholar] [CrossRef]
  10. Liu, X.; Zhai, X.; Sheng, W.; Tu, J.; Zhao, Z.; Shi, Y.; Xu, C.; Ge, G.; Jia, X. Isolated Single Iron Atoms Anchored on a N, S-Codoped Hierarchically Ordered Porous Carbon Framework for Highly Efficient Oxygen Reduction. J. Mater. Chem. A 2021, 9, 10110–10119. [Google Scholar] [CrossRef]
  11. Jiang, P.-Y.; Xiao, Z.-H.; Li, S.-H.; Luo, Z.-N.; Qiu, R.; Wu, H.; Li, N.; Liu, Z.-Q. Sulfuration of Fe–N/C Porous Nanosheets as Bifunctional Catalyst with Remarkable Biocompatibility for High-Efficient Microbial Fuel Cells. J. Power Sources 2021, 512, 230491. [Google Scholar] [CrossRef]
  12. Hu, K.; Tao, L.; Liu, D.; Huo, J.; Wang, S. Sulfur-Doped Fe/N/C Nanosheets as Highly Efficient Electrocatalysts for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2016, 8, 19379–19385. [Google Scholar] [CrossRef] [PubMed]
  13. Ni, B.; Chen, R.; Wu, L.; Xu, X.; Shi, C.; Sun, P.; Chen, T. Optimized Enhancement Effect of Sulfur in Fe–N–S Codoped Carbon Nanosheets for Efficient Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2020, 12, 23995–24006. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, M.; Yang, W.; Li, X.; Xu, Y.; Zheng, L.; Su, C.; Liu, B. Atomically Dispersed Fe–Heteroatom (N, S) Bridge Sites Anchored on Carbon Nanosheets for Promoting Oxygen Reduction Reaction. ACS Energy Lett. 2021, 6, 379–386. [Google Scholar] [CrossRef]
  15. Rao, P.; Liu, Y.; Su, Y.-Q.; Zhong, M.; Zhang, K.; Luo, J.; Li, J.; Jia, C.; Shen, Y.; Shen, C.; et al. S, N Co-Doped Carbon Nanotube Encased Co NPs as Efficient Bifunctional Oxygen Electrocatalysts for Zinc-Air Batteries. Chem. Eng. J. 2021, 422, 130135. [Google Scholar] [CrossRef]
  16. Bhange, S.N.; Soni, R.; Singla, G.; Ajithkumar, T.G.; Kurungot, S. FeNx/FeSx-Anchored Carbon Sheet-Carbon Nanotube Composite Electrocatalysts for Oxygen Reduction. ACS Appl. Nano Mater. 2020, 3, 2234–2245. [Google Scholar] [CrossRef]
  17. Zeng, H.; Wang, W.; Li, J.; Luo, J.; Chen, S. In Situ Generated Dual-Template Method for Fe/N/S Co-Doped Hierarchically Porous Honeycomb Carbon for High-Performance Oxygen Reduction. ACS Appl. Mater. Interfaces 2018, 10, 8721–8729. [Google Scholar] [CrossRef]
  18. Zhao, H.; Chen, L.; Xu, Y.; Wang, H.; Li, J.-Y.; Xie, Y.; Wang, L. A Nitrogen and Sulfur Co-Doped Iron-Based Electrocatalyst Derived from Iron and Biomass Ligand towards the Oxygen Reduction Reaction in Alkaline Media. Dalton Trans. 2021, 50, 13943–13950. [Google Scholar] [CrossRef]
  19. Zhang, T.; Mao, S.; Sun, P.; Gao, X.; Fang, H.; Luo, H.; Zhang, W.; Zhou, B. Nanosized FeS/ZnS Heterojunctions Derived Using Zeolitic Imidazolate Framework-8 (ZIF-8) for pH-Universal Oxygen Reduction and High-Efficiency Zn–Air Battery. J. Colloid Interface Sci. 2022, 608, 446–458. [Google Scholar] [CrossRef]
  20. Ji, S.-M.; Muthurasu, A.; Chhetri, K.; Kim, H.Y. Metal-Organic Framework Assisted Vanadium Oxide Nanorods as Efficient Electrode Materials for Water Oxidation. J. Colloid Interface Sci. 2022, 618, 475–482. [Google Scholar] [CrossRef]
  21. Chhetri, K.; Dahal, B.; Mukhiya, T.; Tiwari, A.P.; Muthurasu, A.; Kim, T.; Kim, H.; Kim, H.Y. Integrated Hybrid of Graphitic Carbon-Encapsulated CuxO on Multilayered Mesoporous Carbon from Copper MOFs and Polyaniline for Asymmetric Supercapacitor and Oxygen Reduction Reactions. Carbon 2021, 179, 89–99. [Google Scholar] [CrossRef]
  22. Jafari, M.; Parnian, M.J.; Gharibi, H.; Liang, Z.; Zou, R. Carbon-Based Nonprecious Metal Electrocatalysts Derived from MOFs for Oxygen Reduction Reaction. Int. J. Energy Res. 2021, 45, 15676–15738. [Google Scholar] [CrossRef]
  23. Qian, Y.; Cavanaugh, J.; Khan, I.A.; Wang, X.; Peng, Y.; Hu, Z.; Wang, Y.; Zhao, D. Fe/Fe3C/N-Doped Carbon Materials from Metal-Organic Framework Composites as Highly Efficient Oxygen Reduction Reaction Electrocatalysts. ChemPlusChem 2016, 81, 718–723. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, X.; Zhang, L.; Liu, C.-P.; Ge, J.-J.; Zhu, J.-B.; Xing, W. Recent Advances in Structural Regulation on Non-Precious Metal Catalysts for Oxygen Reduction Reaction in Alkaline Electrolytes. J. Electrochem. 2022, 28, 2108501. [Google Scholar]
  25. Yang, H.B.; Guo, C.; Zhang, L.; Hu, F.X.; Cai, W.; Gao, J.; Li, C.M.; Liu, B. Nitrogen and Sulfur Co-Doped Graphene Inlaid with Cobalt Clusters for Efficient Oxygen Reduction Reaction. Mater. Today Energy 2018, 10, 184–190. [Google Scholar] [CrossRef]
  26. Li, Q.; Chen, W.; Xiao, H.; Gong, Y.; Li, Z.; Zheng, L.; Zheng, X.; Yan, W.; Cheong, W.; Shen, R.; et al. Fe Isolated Single Atoms on S, N Codoped Carbon by Copolymer Pyrolysis Strategy for Highly Efficient Oxygen Reduction Reaction. Adv. Mater. 2018, 30, e1800588. [Google Scholar] [CrossRef]
  27. Cao, L.; Li, Z.-H.; Gu, Y.; Li, D.-H.; Su, K.-M.; Yang, D.-J.; Cheng, B.-W. Rational Design of N-Doped Carbon Nano-Box-Supported Fe/Fe2N/Fe3C Nanoparticles as Efficient Oxygen Reduction Catalysts for Zn–Air Batteries. J. Mater. Chem. A 2017, 5, 11340–11347. [Google Scholar] [CrossRef]
  28. Lian, Y.; Shi, K.; Yang, H.; Sun, H.; Qi, P.; Ye, J.; Wu, W.; Deng, Z.; Peng, Y. Elucidation of Active Sites on S, N Codoped Carbon Cubes Embedding Co-Fe Carbides toward Reversible Oxygen Conversion in High-Performance Zinc-Air Batteries. Small 2020, 16, 1907368. [Google Scholar] [CrossRef]
  29. Zhang, D.; Mai, Y.J.; Xiang, J.Y.; Xia, X.H.; Qiao, Y.Q.; Tu, J.P. FeS2/C Composite as an Anode for Lithium Ion Batteries with Enhanced Reversible Capacity. J. Power Sources 2012, 217, 229–235. [Google Scholar] [CrossRef]
  30. Liu, M.; Zhang, M.; Zhang, P.; Xing, Z.; Jiang, B.; Yu, Y.; Cai, Z.; Li, J.; Zou, J. ZIF-67-Derived Dodecahedral Co@N-Doped Graphitized Carbon Protected by a Porous FeS2 Thin-Layer as an Efficient Catalyst to Promote the Oxygen Reduction Reaction. ACS Sustain. Chem. Eng. 2020, 8, 4194–4206. [Google Scholar] [CrossRef]
  31. Zang, R.; Li, P.; Guo, X.; Man, Z.; Zhang, S.; Wang, C.; Wang, G. Yolk-Shell N-Doped Carbon Coated FeS2 Nanocages as a High-Performance Anode for Sodium-Ion Batteries. J. Mater. Chem. A 2019, 7, 14051–14059. [Google Scholar] [CrossRef]
  32. Wang, H.; Yin, F.X.; Liu, N.; Kou, R.H.; He, X.B.; Sun, C.J.; Chen, B.H.; Liu, D.J.; Yin, H.Q. Engineering Fe–Fe3C@Fe–N–C Active Sites and Hybrid Structures from Dual Metal–Organic Frameworks for Oxygen Reduction Reaction in H2–O2 Fuel Cell and Li–O2 Battery. Adv. Funct. Mater. 2019, 29, 1901531. [Google Scholar] [CrossRef]
  33. Cui, Z.; Bai, X. Highly Active and Stable Fe/Co/N Co-doped Carbon-Anchored Pd Nanoparticles for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2022, 14, 9024–9035. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, S.; Liu, L.; Chen, X.; Yang, Z.; Li, M.; Wang, Y.; Lv, W.; Zhu, P.; Zhao, X.; Wang, G. On an Easy Way to Prepare Fe, S, N Tri-Doped Mesoporous Carbon Materials as Efficient Electrocatalysts for Oxygen Reduction Reaction. Electrocatalysis 2018, 10, 72–81. [Google Scholar] [CrossRef]
  35. Li, Y.; Yang, J.; Huang, J.; Zhou, Y.; Xu, K.; Zhao, N.; Cheng, X. Soft Template-assisted Method for Synthesis of Nitrogen and Sulfur Co-Doped Three-Dimensional Reduced Graphene Oxide as an Efficient Metal Free Catalyst for Oxygen Reduction Reaction. Carbon 2017, 122, 237–246. [Google Scholar] [CrossRef]
  36. Men, B.; Sun, Y.; Liu, J.; Tang, Y.; Chen, Y.; Wan, P.; Pan, J. Synergistically Enhanced Electrocatalytic Activity of Sandwich-like N-Doped Graphene/Carbon Nanosheets Decorated by Fe and S for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2016, 8, 19533–19541. [Google Scholar] [CrossRef]
  37. Fantauzzi, M.; Licheri, C.; Atzei, D.; Loi, G.; Elsener, B.; Rossi, G.; Rossi, A. Arsenopyrite and Pyrite Bioleaching: Evidence from XPS, XRD and ICP Techniques. Anal. Bioanal. Chem. 2011, 401, 2237–2248. [Google Scholar] [CrossRef]
  38. Kramm, U.I.; Ni, L.; Wagner, S. 57Fe Mössbauer Spectroscopy Characterization of Electrocatalysts. Adv. Mater. 2019, 31, 1805623. [Google Scholar] [CrossRef]
  39. Perez, A.; Treilleux, M.; Capra, T.; Griscom, D.L. Precipitation Phenomena in High-Dose Iron-Implanted Silica and Annealing Behavior. J. Mater. Res. 1987, 2, 910–917. [Google Scholar] [CrossRef]
  40. Hoque, M.; Zhang, S.; Thomas, M.L.; Li, Z.; Suzuki, S.; Ando, A.; Yanagi, M.; Kobayashi, Y.; Dokko, K.; Watanabe, M. Simple Combination of a Protic Salt and an Iron Halide: Precursor for a Fe, N and S Co-Doped Catalyst for the Oxygen Reduction Reaction in Alkaline and Acidic Media. J. Mater. Chem. A 2018, 6, 1138–1149. [Google Scholar] [CrossRef]
  41. Chen, C.; Li, H.-J.; Bai, Y.; Feng, F.-X.; Tian, L.; Yang, Y.; Liu, Y.; Guo, Q. Effect of Sulfidation Temperature on Component Transformation and Catalytic Performance of Direct Coal Liquefaction Catalyst. J. Fuel Chem. Technol. 2022, 50, 54–62. [Google Scholar] [CrossRef]
  42. Wu, Y.; Yang, Y.; Gao, T.; Chen, X.; Huang, Z.; Zhang, J.; Guo, Y.; Xiao, D. Ligand Regulation to Prepare an Fe, N, S Tri-Codoped Hollow Carbon Electrocatalyst for Enhanced ORR Performance and Zn-Air Batteries. Sustain. Energy Fuels 2020, 4, 4117–4125. [Google Scholar] [CrossRef]
  43. 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]
  44. Zhang, J.; Xu, D.; Wang, C.; Guo, J.; Yan, F. Rational Design of Fe1-xS/Fe3O4/Nitrogen and Sulfur-Doped Porous Carbon with Enhanced Oxygen Reduction Reaction Catalytic Activity. Adv. Mater. Interfaces 2018, 5, 1701641. [Google Scholar] [CrossRef]
  45. Xu, H.; Wang, D.; Yang, P.; Liu, A.; Li, R.; Xiao, L.; Zhang, J.; Qu, Z.; An, M. FeS Encapsulated Hierarchical Porous S, N-Dual-Doped Carbon for Oxygen Reduction Reaction Facilitation in Zn-Air Batteries. Sustain. Energy Fuels 2021, 5, 2695–2703. [Google Scholar] [CrossRef]
  46. Chen, D.; Fang, Z.; Ma, X.; Li, Z.; Lin, H.; Ying, W.; Peng, X. Rational Design of Fe/S/N/C Catalyst from ZIF-8 for Efficient Oxygen Reduction Reaction. Nanotechnology 2020, 31, 475404. [Google Scholar] [CrossRef]
  47. She, Y.; Liu, J.; Wang, H.; Li, L.; Zhou, J.; Leung, M.K.H. Bubble-like Fe-Encapsulated N, S-Codoped Carbon Nanofibers as Efficient Bifunctional Oxygen Electrocatalysts for Robust Zn-Air Batteries. Nano Res. 2020, 13, 2175–2182. [Google Scholar] [CrossRef]
  48. Javed, R.; Liu, J.; Kang, Z.; Gong, Y.; Xu, Y.; Ye, D.; Zhao, H.; Zhang, J. Self-Doped Sulfur Enhanced the Activity of Fe-N for Oxygen Reduction Reaction. J. Electrochem. Soc. 2021, 168, 046512. [Google Scholar] [CrossRef]
  49. Li, Z.; Wang, W.; Zhou, M.; He, B.; Ren, W.; Chen, L.; Xu, W.; Hou, Z.; Chen, Y. In-Situ Self-Templated Preparation of Porous Core–Shell Fe1–S@N, S Co-Doped Carbon Architecture for Highly Efficient Oxygen Reduction Reaction. J. Energy Chem. 2020, 54, 310–317. [Google Scholar] [CrossRef]
  50. Zhang, H.-M.; Zhu, C. Co Nanoparticles-Embedded N, S-Codoped Hierarchically Porous Graphene Sheets as Efficient Bifunctional Electrocatalysts for Oxygen Reduction Reaction and Hydrogen Evolution Reaction. J. Mater. Res. Technol. 2020, 9, 16270–16279. [Google Scholar] [CrossRef]
  51. Chae, G.S.; Youn, D.H.; Lee, J.S. Nanostructured Iron Sulfide/N, S Dual-Doped Carbon Nanotube-Graphene Composites as Efficient Electrocatalysts for Oxygen Reduction Reaction. Materials 2021, 14, 2146. [Google Scholar] [CrossRef] [PubMed]
  52. Yang, W.; Zhang, Y.; Liu, X.; Chen, L.; Liu, M.; Jia, J. Polymerization-Dissolution Strategy to Prepare Fe, N, S Tri-Doped Carbon Nanostructures for Zn-Air batteries. Carbon 2019, 147, 83–89. [Google Scholar] [CrossRef]
  53. Li, J.; Song, Y.; Zhang, G.; Liu, H.; Wang, Y.; Sun, S.; Guo, X. Pyrolysis of Self-Assembled Iron Porphyrin on Carbon Black as Core/Shell Structured Electrocatalysts for Highly Efficient Oxygen Reduction in Both Alkaline and Acidic Medium. Adv. Funct. Mater. 2016, 27, 1604356. [Google Scholar] [CrossRef]
  54. Xie, Y.; Li, H.; Tang, C.; Li, S.; Li, J.; Lv, Y.; Wei, X.; Song, Y. A High-Performance Electrocatalyst for Oxygen Reduction Based on Reduced Graphene Oxide Modified with Oxide Nanoparticles, Nitrogen Dopants, and Possible Metal-N-C Sites. J. Mater. Chem. A 2014, 2, 1631–1635. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration for the synthesis of FeNSC-850.
Scheme 1. Schematic illustration for the synthesis of FeNSC-850.
Catalysts 12 00806 sch001
Figure 1. SEM images of (a) and (b) FeNSC-850; (c) the HAADF-STEM image of FeNSC-850 and its corresponding element mapping of C, N, O, S, and Fe.
Figure 1. SEM images of (a) and (b) FeNSC-850; (c) the HAADF-STEM image of FeNSC-850 and its corresponding element mapping of C, N, O, S, and Fe.
Catalysts 12 00806 g001
Figure 2. Physical characterizations of structure and composition of electrocatalysts: (a) X-ray diffraction pattern of FeNSC-850 accompanied by the standard XRD patterns of related substances of FeS (PDF#76−0965), FeS2 (PDF#42−1340), and Fe3C (PDF#03−0411); (b) N2 adsorption–desorption isotherms and (c) corresponding pore size distributions of FeNSC-850 and MIL-88-850.
Figure 2. Physical characterizations of structure and composition of electrocatalysts: (a) X-ray diffraction pattern of FeNSC-850 accompanied by the standard XRD patterns of related substances of FeS (PDF#76−0965), FeS2 (PDF#42−1340), and Fe3C (PDF#03−0411); (b) N2 adsorption–desorption isotherms and (c) corresponding pore size distributions of FeNSC-850 and MIL-88-850.
Catalysts 12 00806 g002
Figure 3. XPS (a) C 1s and (b) N 1s spectra of FeNSC-850; and (c) the N content relative to C in electrocatalysts of FeNSC-T and MIL-88-850, XPS (d) S 2p, and (e) Fe 2p spectra of FeNSC-850.
Figure 3. XPS (a) C 1s and (b) N 1s spectra of FeNSC-850; and (c) the N content relative to C in electrocatalysts of FeNSC-T and MIL-88-850, XPS (d) S 2p, and (e) Fe 2p spectra of FeNSC-850.
Catalysts 12 00806 g003
Figure 4. Mössbauer spectra of (a) MIL-88-850 and (b) FeNSC-850, with the assignment of Mössbauer sites to iron species in the relative absorption area.
Figure 4. Mössbauer spectra of (a) MIL-88-850 and (b) FeNSC-850, with the assignment of Mössbauer sites to iron species in the relative absorption area.
Catalysts 12 00806 g004
Figure 5. (a) Linear sweep voltammogram (LSV) curves of FeNSCs pyrolyzed with various heating-temperature from 750 to 900 °C and MIL-88 carbonized at 850 °C at 1600 rpm; (b) potentials of FeNSC-T (T = 750 − 900 °C) and MIL-88-850 at −2.5 mA cm−2 based on the LSV data from (a); (c) LSV curves of as-prepared electrocatalysts with different mass ratio of chitosan and L-Cy, the electrocatalyst without additional Fe salt with mass ratio of chitosan and L-Cy as 1:1, and commercial Pt/C at loading of 10 and 20 µgPt cm−2 with a scanning rate of 5 mV s−1; (d) LSV curves of FeNSC-850 with various rotation rates from 400 to 2500 rpm; (e) Koutecky-Levich (K-L) plots of FeNSC-850 based on the data from (d); and (f) HO2-% and electron transfer number of FeNSC-850, MIL-88-850, and 20 wt% Pt/C with Pt loading of 20 µgPt cm−2 in 0.1 M KOH electrolyte.
Figure 5. (a) Linear sweep voltammogram (LSV) curves of FeNSCs pyrolyzed with various heating-temperature from 750 to 900 °C and MIL-88 carbonized at 850 °C at 1600 rpm; (b) potentials of FeNSC-T (T = 750 − 900 °C) and MIL-88-850 at −2.5 mA cm−2 based on the LSV data from (a); (c) LSV curves of as-prepared electrocatalysts with different mass ratio of chitosan and L-Cy, the electrocatalyst without additional Fe salt with mass ratio of chitosan and L-Cy as 1:1, and commercial Pt/C at loading of 10 and 20 µgPt cm−2 with a scanning rate of 5 mV s−1; (d) LSV curves of FeNSC-850 with various rotation rates from 400 to 2500 rpm; (e) Koutecky-Levich (K-L) plots of FeNSC-850 based on the data from (d); and (f) HO2-% and electron transfer number of FeNSC-850, MIL-88-850, and 20 wt% Pt/C with Pt loading of 20 µgPt cm−2 in 0.1 M KOH electrolyte.
Catalysts 12 00806 g005
Figure 6. (a) LSV curves of FeNSC-850 at 1600 rpm with a scanning rate of 5 mV s−1 before and after 5000 cycles in O2-saturated 0.1 M KOH solution, (b) I-t response of FeNSC-850 and commercial Pt/C with the injection of methanol at ~400s in O2-saturated 0.1 M KOH electrolyte.
Figure 6. (a) LSV curves of FeNSC-850 at 1600 rpm with a scanning rate of 5 mV s−1 before and after 5000 cycles in O2-saturated 0.1 M KOH solution, (b) I-t response of FeNSC-850 and commercial Pt/C with the injection of methanol at ~400s in O2-saturated 0.1 M KOH electrolyte.
Catalysts 12 00806 g006
Table 1. The Mössbauer spectra parameters of MIL-88-850 and FeNSC-850.
Table 1. The Mössbauer spectra parameters of MIL-88-850 and FeNSC-850.
SampleIS a
(mm s−1)
QS b
(mm s−1)
LW c
(mm s−1)
AssignmentRef.
MIL-88-850Doublet (1)0.260.610.37Fe2+N4 (hexagonal symmetry)[38]
Sextet (1)0.190.00280.45Fe3C[38]
FeNSC-850Doublet (2)0.310.720.65FeS2[38]
Doublet (3)0.611.340.28Fe2+N4 (tetrahedral symmetry)[39]
Sextet (2)0.15−0.190.64FexC[40]
Sextet (3)0.670.0710.60FeSx[38,41]
Sextet (4)0.74−0.170.28
Sextet (5)0.72−0.0280.40
a IS isomer shift. b QS is quadrupole splitting. c LW is line width.
Table 2. Comparison of ORR performances of FeNSC-850 with some representative reports in 0.1 M KOH solution.
Table 2. Comparison of ORR performances of FeNSC-850 with some representative reports in 0.1 M KOH solution.
SampleE1/2
(V vs. RHE)
n a
/RRDE
n b
/K-L
HO2%
/RRDE
StabilityRef.
Fe/S/N/C-20.853.93–3.98~40.92–3.91
(0.1–0.9 V)
∆E1/2 = 21 mV
(10000 CV cycles)
[46]
FeNSC8000.783.77–3.903.56–3.87<11.4No obvious decay
(3000 CV cycles)
[47]
Fe-N/S-C0.87-3.9<5No obvious decay
(2000 CV cycles)
[48]
Fe1−xS@NSC-240.70-3.82–4.00-No obvious decay
(4000 CV cycles)
[49]
Co@NSC-10000.83-3.7-∆E1/2 = 18 mV
(5000 CV cycles)
[50]
Fe,S/NGC-9000.833.963.98−4.00<2No obvious decay
(20000 s I-t)
[36]
Co@NPC@FeS2-0.50.853.86–3.88~45.1−5.6
(0–0.8 V)
∆E1/2 = 13 mV
(10000 CV cycles)
[30]
FeS/N,S:CNT–GR0.83-~4-∆E1/2 = 6 mV
(6000 CV cycles)
[51]
FeNSC-8500.823.973.92<2.6∆E1/2 = 15 mV
(5000 CV cycles)
This work
a n is determined by the data from RRDE. b n is determined by the data from K-L.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Xu, Y.; Wang, H.; Zhang, J.; Zhao, H.; Chen, L.; Xu, L.; Xie, Y.; Huang, J. MIL-88-Derived N and S Co-Doped Carbon Materials with Supplemental FeSx to Enhance the Oxygen Reduction Reaction Performance. Catalysts 2022, 12, 806. https://doi.org/10.3390/catal12080806

AMA Style

Liu Y, Xu Y, Wang H, Zhang J, Zhao H, Chen L, Xu L, Xie Y, Huang J. MIL-88-Derived N and S Co-Doped Carbon Materials with Supplemental FeSx to Enhance the Oxygen Reduction Reaction Performance. Catalysts. 2022; 12(8):806. https://doi.org/10.3390/catal12080806

Chicago/Turabian Style

Liu, Yu, Yinghao Xu, He Wang, Jia Zhang, Haiyan Zhao, Li Chen, Ling Xu, Yan Xie, and Jiahui Huang. 2022. "MIL-88-Derived N and S Co-Doped Carbon Materials with Supplemental FeSx to Enhance the Oxygen Reduction Reaction Performance" Catalysts 12, no. 8: 806. https://doi.org/10.3390/catal12080806

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop