Chinese Chemical Letters  2017, Vol. 28 Issue (4): 748-754   PDF    
Red-blood-cell like nitrogen-doped carbons with highly catalytic activity towards oxygen reduction reaction
Jing-Jing Xu, Chun-Hui Xiao, Shu-Jiang Ding     
Department of Applied Chemistry, School of Science, Xi'an Jiaotong University, Xi'an 710049, China
Abstract: A highly active nitrogen-doped catalyst with a unique red-blood-cell(RBC)like structure is reported for oxygen reduction reaction(ORR).The catalyst Fe, N-C@carbon-900 was prepared by pyrolysis of the polyaniline(PANI)and polystyrene(PS)composites with adsorption of ferric ion on the shell of sphere structure at 900 ℃.Fe, N-C@carbon-900 with a unique RBC-like structure provides plenty of catalytic sites combining the electrical conductivity of the carbon sphere with the catalytic activity of the nitrogen-doped layer.The four-electron reduction pathway is selected for the catalyst Fe, N-C@carbon-900.The catalyst exhibit the ORR Eonset at 0.87 V(potentials is versus to reversible hydrogen electrode (RHE)), E1/2 at 0.78 V and high diffusion-limiting current density(5.20 mA/cm2).Furthermore, this work indicates that both N and Fe accounted for high activity of the catalyst Fe, N-C@carbon-900 toward the oxygen reduction process.It is concluded that Fe and N exhibit synergistically promotion in the ORR activity for the catalyst Fe, N-C@carbon-900.We also provide a rational design of electrocatalysts with high ORR activity to further clarify the essential ORR sites of heteroatom doped carbon materials for fuel cells and metal-air battery applications.
Key words: Nitrogen-doped carbon     Red-blood-cell like structure     ORR     PANI     Electrocatalysts     Four-electronpathway    
1. Introduction

The oxygen reduction reaction plays a vital role in energy conversion and storage applications such as fuel cells and metal– air batteries. The electrocatalysts for ORR are crucial in determining the energy conversion efficiency because the ORR is a kinetically sluggish reaction [1, 2]. Up to now, platinum (Pt) based catalysts are regarded as the best catalysts [3, 4]. However, poor durability and high cost of the available Pt-based catalysts restricts the commercialization and application of fuel cell power systems [57]. Therefore, seeking for suitable alternatives with low cost, high efficiency and durability for ORR has become a major research focus in the past decades.

Highly efficient nonprecious metal and metal-free catalysts have attracted numerous researchers' attention due to their superior activity, comparable stability and enhanced durability against CO or methanol [811]. Many research efforts have been devoted to develop porous carbon based materials as catalyst supports [1216]. High surface area, porous architecture and high conductivity facilitate gas diffusion, electron transfer [1720]. Nitrogen-doped carbon materials represent one of the most promising electrocatalysts for ORR due to their unique electronic and structure properties [2125]. The strong electron-withdraw property of N atom could alter the charge density and spin density of the adjacent C atoms, which facilitate the adsorption of O2 and weaken the O-O bond [26]. Apparently, four-electron pathway is preferential to proceed based on the above mentioned conditions. However, the ORR performance of Ndoped carbons is still inferior to the commercial Pt-based catalysts. An agreement has been reached that Fe is important for the ORR activity of the N-doped carbons [2731]. However, the real active sites for N-dope carbon with Fe catalysts are still under discussion [32].

Herein, we prepare a red-blood-cell (RBC) like Fe, N-C@carbon sphere ORR catalyst (detailed synthetic scheme are shown in Scheme 1). Nitrogen-containing active sites could be incorporated into the adjacent carbon matrix with pyrolysis treatment. Polystyrene (PS) are chemically modified with sulfonate group and used as the template. The polyaniline (PANI)@PS composite is pyrolyzed to prepare the catalyst Fe, N-C@carbon. The high surface area and core–shell interconnection could provide rich active sites and good electronic transfer pathway. This catalyst has the similar morphology with RBC to adsorb and diffuse O2. The catalyst exhibits superior catalytic activity for ORR with the onset potential (Eonset) of 0.87 V and the half-wave potential (E1/2) of 0.78 V in 0.1 mol/L KOH solution, which is close to those of commercial Pt/C. It is found that the carbon sphere support of the catalyst ruptured at high pyrolysis temperature. The ruptured morphology dramatically increases the active sites and the catalytic performance of the Fe, N-C@Carbon catalyst improves significantly. This work provides the solid evidence that this novel morphology and the introduction of Fe boost the catalytic activity of Fe, N-C@carbon catalyst.

Download:
Scheme 1. Schematic illustration of the fabrication procedure of nitrogen-doped RBC-like carbon electrocatalyst.

2. Results and discussion

As shown in Fig. 1, the morphology of the resulting RBC-like PANI@PS composites were characterized by SEM and TEM. A thin hydrophilic interior layer and a permeable transverse channel, comprised of poly(methyl methacrylate)-poly(methacrylic acid) (PMMA-PMA), were embedded in the PS shell of the hollow template spheres with an average diameter of 400 nm (Fig. 1a and b) [38]. As shown in Fig. 1cf, nanosized uniform trichoid PANI was formed on the exterior surface of the PS template after polymerization. Repeated centrifugation and wash with deionized water before polymerization ensured that PANI pillars mostly grew on the PS latex. The resulting RBC-like PANI@PS composites were obtained with the PANI pillars growing radially [34, 38]. Some hollow spheres were broken to confirm that the spheres were hollow (Fig. 1gi). We could easily see that the ruptured morphology dramatically increased the active sites and the catalytic performance of the Fe, N-C@carbon-900 catalyst improved significantly.

Download:
Figure 1. SEM of PS (a), crosslinkedsulfonated PS (b), PANI@PS composites (c–e); (f) TEM of PANI@PS composites; SEM of pyrolysis of the PANI@PS composites at 900 ℃ (g–i).

The emeraldine form of polyaniline was confirmed by the FT-IR spectra shown in Fig. 2a. The two moderate peaks at 1570 cm-1 and 1490 cm-1 corresponded to the C=C stretching deformation of quinoid and benzenoid rings, respectively. The absorption bands at 1297 cm-1 and 1132 cm-1 were assigned to the stretching of secondary aromatic amine C-N and in-plane bending of the aromatic C-H. The typical out-of-plane deformation of C-H of 1, 4-disubstituted benzene ring were observed at 797 cm-1 and 505 cm-1, respectively. These characteristic peaks indicated that PANI possessed the conductive emeraldine structure.

Download:
Figure 2. (a) FT-IR of the PS template and the PANI@PS composites; (b) Raman of the carbon-900 and the Fe, N-C@carbon-900; (c) XRD of the Fe, N-C@carbon-900, Fe, NC@Carbon-800, Fe, N-C@carbon-700and N-C@carbon; (d) The TGA curve of the PANI@PS composites.

The Raman spectra of carbon sphere and the catalyst Fe, NC@carbon-900 displayed two typical peaks about 1320 cm-1 and 1592 cm-1, corresponding to the D bands and G bands, respectively (Fig. 2b) [39, 40]. The D bands were assigned to the defective nature of carbon. The G bands denoted to graphitic carbon which were attributed to the coupling effect between the in-plane-bond stretching of C=C and the hexagonal sheets of graphene [4143]. The intensity ratio of D band and G band (ID/IG) elucidated information about the degree of the defects and disordered structures. The decreased ID/IG ratio from 1.315 to 1.086 for the carbon sphere and the catalyst Fe, N-C@carbon-900 gave the evidence that higher degree of defects existed in the carbonized PS template and the introduction of Fe and N contributed to produce the graphitic structure. As for the catalyst Fe, N-C@carbon-900, relatively stronger peak intensity of the D band than the G band also provided the information that a certain amount of edges/ defects existed in the Fe, N-C@carbon. It might also play a crucial role to enhance the ORR catalytic activity [44, 45]. Furthermore, we carried out XRD analysis (Fig. 2c) to understand structural transformation involved during the pyrolysis process. The NC@carbon-900 showed a characteristic peak at about 27.3°, which corresponded to the stacking peak of π-conjugated layers and (002) phase of graphitic materials [46]. Graphite structure show good electronic conductivity which is essential for the electrocatalysts. When introducing FeCl3, peaks of FeCl3·nH2O, iron hydroxide, ferric oxide, Fe–C and ferriccarbide are clearly found according to the XRD results of Fe, N-C@carbon-900. As the annealing temperature increased, Fe was gradually generated. The higher catalytic activity of the Fe, N-C@carbon heat-treated at 900 ℃ indicated that the forms of Fe in the catalyst at this temperature enhanced the number of active sites of the catalysts. It is speculated that, during the heat treated process, iron might react with the nitrogen contained carbon derived from PANI in situ and catalyzed to form more active sites with the existence of Fe and N at the same time under the proper annealing temperature.

Furthermore, TGA analysis has been performed with Fe, NC@carbon-900 precursor under N2 atmosphere with a scan rate of 10 ℃/min in the range of 25–900 ℃. The possible mechanism involves the following reactions [47]:

(1)
(2)
(3)
(4)

As shown in Fig. 2d, weight loss below 158.6 ℃ is due to the loss of absorbed water molecules. At this stage, FeCl3·nH2O is hydrolyzed into amorphous Fe(OH)3 (as shown in the Eq. (1)). The weight loss is rather obvious at between 158.6 ℃ and 453.9 ℃, which is assigned to the decomposition of low molecular weight polymers or oligomers of PS and PANI, along with the loss of H2O released from the decomposition of Fe(OH)3 (Eq. (2)). The weight loss starts at 453.9 ℃ and ends at around 563.4 ℃, due to the decomposition of FeO(OH) (Eq. (3)). During the range of 158.6– 563.4 ℃, the total weight loss of Fe, N-C@carbon is 61.6%. The weight loss between 563.4 ℃ and 755.6 ℃ indicated that Fe2O3 reacted with carbon (Eq. (4)). At higher temperature from 755.6 ℃, Fe2O3 gradually transformed into Fe.

Cyclic voltammetries (CVs) (Fig. 3a) of the prepared electrocatalysts were carried out in 0.1 mol/L saturated KOH. The catalyst Fe, N-C@Carbon-900 showed a remarkable reduction peak at 0.820 V and the highest peak current -303.92 μA, which were almost as better as those of commercial 20 wt.% Pt/C (0.780 V, -316.63 μA, respectively). In contrast, carbon-900 exhibits ORR peak potential at 0.750 V and a small peak current -141.93 μA, which meant poor ORR catalytic activity. The catalyst N-C@carbon-900 showed more positive peak potential (0.770 V). Fe, NC@carbon-700 (0.758 V) and Fe, N-C@carbon-800 (0.763 V) showed more positive peak potential than the carbon sphere and less positive peak potential than the Fe, N-C@carbon-900. The results indicated that both nitrogen-doped carbon and Fe enhanced ORR catalytic activity and the proper pyrolysis temperature of the catalysis was 900 ℃.

Download:
Figure 3. (a) CVs of Fe, N-C@carbon-900, Pt/C, Fe, N-C@Carbon-800, Fe, N-C@carbon-700, N-C@carbon-900 and carbon sphere-900; (b) RDE test of Fe, N-C@carbon-900, Pt/C, NC@carbon-900 and carbon-900 at 1600 rpm; RDE test of Fe, N-C@carbon-900 (c) and Pt/C (d) at different rotation rate.

Furthermore, to clearly understand the intrinsic electrocatalytic activity, the ORR activities of the catalysts were characterized by linear sweep voltammograms (LSVs). Specifically, a series of RDE studies of carbon sphere-900, N-C@carbon-900, Fe, N-C@carbon-900 and Pt/C were collected on RDE at 1600 rpm (Fig. 3b). The onset potential (Eonset) of the LSVs curves was used to evaluate the activity of catalysts. And the Eonset is defined as the potential is 5% of the ORR current measured at 0.33 V, at which the diffusionlimiting current for all the catalysts was obtained. The Eonset of Fe, N-C@Carbon-900 is 0.87 V, slightly more negative than Pt/C (0.96 V). The catalyst Fe, N-C@carbon-900 showed superior activity in comparison to carbon sphere-900 and N-C@carbon-900, with the onset potential of 0.82 V and 0.85 V, respectively. The half-wave potential (E1/2) of the Fe, N-C@carbon-900 (0.78 V) was very close to the E1/2 of Pt/C (0.82 V). The comparative E1/2 of carbon sphere-900 and N-C@carbon-900 were 0.69 V and 0.75 V. The current density at 0.4 V of the Fe, N-C@carbon-900 (5.20 mA/cm2) was close to Pt/C (5.85 mA/cm2) and much higher than carbon sphere-900 (3.43 mA/cm2) and N-C@carbon-900 (4.34 mA/cm2). Moreover, the unique and the wide current plateau from 0.2 V to 0.6 V of the Fe, NC@carbon-900 indicated a diffusion-controlled process for the four-electron-dominated ORR pathway [48]. The remarkable catalytic performance of the Fe, N-C@carbon-900 was due to its large surface areas, unique porous structure and accessible catalytic active sites. It could be concluded that both Fe and N improved the catalytic activities.

To further understand ORR activity of Fe, N-C@carbon-900, LSVs studies at different rotation rates were performed (Fig. 3c). The catalytic current density increased with the increase of rotation rate. The electron transfer numbers (n) and kinetic limiting current density are also calculated based on Koutecky–Levich (KL) equation. All K–L plots displayed good linearity from 0.2 V to 0.6 V. The electron transfer number for Fe, N-C@carbon-900 was 3.94 on average and suggested good selectivity for the fourelectron dominated pathway for ORR, close to that of the commercial Pt/C with an average n value of 4.00 (Fig. 3d). The large n value and high limiting current density (JK) with a wide current plateau from 0.2 to 0.6 V confirmed the superior and stable electrochemical performance for Fe, N-C@carbon-900.

Furthermore, RRDE measurements were employed to assess the electrocatalytic activity of the catalysts (shown in Fig. 4a and b). If oxygen is reduced through a two-electron pathway, hydrogen peroxide (H2O2) would generate as the intermediate. The Fe, NC@carbon-900 exhibited an ORR process through a four-electron transfer pathway with the H2O2 yield of 9.3% and the average n value of 3.83 under the potential range from 0.55 to 0.90 V. By comparison, Pt/C produced 3.2% H2O2 with the average n of 3.94 under the same condition.

Download:
Figure 4. (a) RRDE test; (b) extent of peroxide production and the electrontransfer number of Pt/C and Fe, N-C@carbon-900 electrodes.

The heat-treated temperature played a crucial role on the electrocatalytic activity of the catalysts [49, 50]. The precursors of the catalysts were treated at 700 ℃, 800 ℃ and 900 ℃ (denoted as Fe, N-C@carbon-700, Fe, N-C@carbon-800, Fe, N-C@carbon-900, respectively) to confirm the proper pyrolysis temperature for the Fe, N-C@carbon catalysts. The LSVs studies at different rotation rates of Fe, N-C@carbon-700 and Fe, N-C@carbon-800 were performed. Because too high pyrolysis temperature was not beneficial to incorporate N with C, but helped to increase the electrical conductivity for the carbon matrix. The ORR activities for the catalyst Fe, N-C@carbon-X (X = 700, 800, 900) increased with increasing the treatment temperature. The ORR Eonset and E1/2 in the RDE polarization curves at 1600 rpm (Fig.S1 in Supporting information) were 0.67 and 0.50 V for Fe, N-C@carbon-700 and 0.72 and 0.56 V for Fe, N-C@carbon-800, which were both inferior to that of Fe, N-C@Carbon-900, which agreed well with the CV results. Thus we chose 900 ℃ as the optimal pyrolysis temperature for the catalysts.

Next, we also carried out the comparison of the catalyst NC@carbon-900 of which the precursors contained N but Fe and the catalyst carbon-900 without Fe or N at different rotation rates (Fig. S2 in Supporting information). Apparently, the catalyst carbon-900 proceeds mainly by a two-electron ORR mechanism. After introducing N, the catalyst N-C@carbon-900 were shown through a four-electron process with more positive potential and higher limiting current density. The higher activity of the N-C@carbon-900 than carbon-900 were attributed to that electron-rich N atom induced charged sites of the adjacent carbon favorable adsorption of O2 [5153]. In association with the ORR performance of Fe, NC@carbon-900, we could reach a conclusion that Fe and N exhibited synergistically promotion in the ORR activity forthe catalyst Fe, N-C@carbon-900.

3. Conclusion

In summary, we have developed a high-performance ORR catalyst with a novel RBC-like structure through pyrolysis treatment of the PANI@PS composites. The unique flat sphere shells afforded abundant catalytic sites of the catalyst Fe, NC@carbon-900 and facilitated the mass and electron transfer process during the oxygen reduction process. The Fe, N-C@carbon-900 exhibited the ORR Eonset at 0.87 V, E1/2 at 0.78 V and high diffusion-limiting current density (5.20 mA/cm2), which were approaching that of Pt/C in alkaline medium. Furthermore, this work indicated that both N and Fe accounted for the high activity of the catalyst Fe, N-C@carbon-900 toward the oxygen reduction process. The precursor PANI was heat treated at 900 ℃ to ensure the nitrogen sites distributed uniformly on the carbon matrix. It could come to the conclusion that Fe and N exhibited synergistically promotion in the ORR activity for the catalyst Fe, N-C@carbon-900. We also provided a rational design of electrocatalysts with high ORR activity to further clarify the essential ORR sites of heteroatom doped carbon materials for fuel cells and metal–air battery applications.

4. Experimental 4.1. Materials and chemicals

Aniline, hydrochloric acid and ferric chloride were purchased and used as received. PS (polystyrene) hollow spheres were purchased from Rhom & Haas Company. Nafion 117 solution (5 wt. %) was purchased from Alfa Aesar (USA). Pt/C catalyst (20 wt.% Pt on Vulcan XC-72R) was obtained from ETEK Inc (USA). All the aqueous solutions were prepared by the ultra-pure water ( > 18 MΩ cm) from a Millipore system.

4.2. Materials characterization

The microstructures of PANI@PS composites and the nitrogen doped RBC-like carbon catalyst were represented by field-emission scanning electron microscopy (SEM; HITACHI, SU-8010) and transmission electron microscopy (TEM; JEOL, JEM-2100). The crystalline phases were further confirmed by X-ray diffraction (XRD) pattern (XRD; SHIMADZU, Lab X XRD-6000). Thermogravimetric analysis (TGA) was conducted on a PerkinElmer TGA7 thermogravimetric analyzer with a heating rate of 10 ℃/ min under nitrogen flowing from 30 ℃ to 900 ℃. Raman scattering spectra were conducted on a HORIBA JOBIN YVON Instrument HR 800 with a laser wavelength of 514 nm. The FT-IR of PANI layer on PS surface was performed by Fourier transform infrared spectroscopy (FT-IR, BRUKER, TENSOR 27).

4.3. Electrochemical measurements

Electrochemical measurements were performed on a Pine biopotentiostat (Pine Instrument Co., USA) using a conventional three-electrode system with the platinum wire as the counter electrode, Ag/AgCl as the reference electrode and glassy carbon electrode (D = 5 mm) modified with a certain load of 4 mg/mL catalyst suspension as the working electrode. The ORR activity and four-electron selectivity of catalysts were evaluated using a rotating disk electrode (RDE) and a rotating ring disk electrode (RRDE), respectively.

A mixture of 4 mg catalysts in 1 mL of water/ethanol (4/1, v/v) and 10 mL 5% Nafion was sonicated for 30 min to form a homogeneous suspension. 20 mL of the suspensionwas cast on glassy carbon electrode surface and then dried at room temperature. As comparison, a commercial Pt/C catalyst suspension was prepared following the same procedure. The rotation rate of RDE was set from 400 rpm to 2025 rpm. The electron transfer number (n) was calculated according to the Koutechy–Levich equation:

Where J stands for the measured current density, JL the limiting diffusion current density, JK the kinetic current density, ω the RDE rotation rate, n the electron transfer number, F the Faraday constant (96485C/mol), DO2 the diffusion coefficient of O2 in 0.1 mol/L KOH (1.73 ×10-5 cm2/S), ν the kinematic viscosity of 0.1 mol/L KOH (0.01 cm2/S) and CO2 the bulk concentration of the dissolved O2 (1.2 ×10-6mol/cm).

The RRDE (D = 5.61 mm) measurements was conducted at a rotating rate of 1600 rpm. The ring potential was 0.5 V (vs. Ag/ AgCl). The electron transfer number (n) and percentage content (X) of the intermediate production HO2- were calculated according to the following equation:

Where Id is the disk current, Ir is the ring current and N is the current collection efficiency (37%) of the Pt ring.

All measurements were carried out in 0.1 mol/L KOH aqueous solution under Oxygen (or nitrogen) atmosphere. O2 (or N2) was bubbled into the electrolyte solution for at least 30 min to ensure the saturation of oxygen (or nitrogen). For CV measurements, at least 10 cycles of CV, ranging from -0.8 V to 0.2 V (vs. Ag/AgCl), were performed before the data collected. The Ag/AgCl electrode was calibrated against the reversible hydrogen electrode (RHE) according to the previously reported procedure [33]. ERHE = EAg/ AgCl + 1.0225 V in 0.1 mol/L KOH.

4.4. Synthesis of the RBC-like PANI@PS composites

The PANI@PS was synthesized through ferric chloride induced oxidative polymerization of aniline [34, 35]. For a typical procedure, 0.04 mol of aniline was added into HCl aqueous solution (30 mL, 2 mol/L) under stirring. Afterwards 30 mg crosslinked sulfonated PS hollow spheres [36, 37] were added into this solution. The mixture was kept for 24 h at room temperature to ensure aniline diffusion into the PS hollow spheres via the channels within the shell. The aniline hydrochloride coated sulfonated PS were centrifuged and washed by deionized water for several times. In situ oxidative polymerization was triggered when the treated monomers were added into 100 mL FeCl3 aqueous solution (0.1 mol/L) and stirred for 24 h at room temperature. The emeraldine PANI@PS composites were obtained after centrifugation with deionized water and dried in air.

4.5. Preparation of the nitrogen-doped carbon catalyst

The obtained composites were centrifuged and washed with FeCl3 solution and dried in air. The FeCl3 adsorbed RBC-like PANI@PS composites were subjected to heat treatment at 700 ℃ (Fe, N-C/carbon-700), 800 ℃ (Fe, N-C/carbon-800) and 900 ℃ (Fe, N-C/carbon-900) for 2 h under nitrogen atmosphere, respectively. For comparison, the RBC-like PANI@PS composites without FeCl3 solution treatment and the crosslinked sulfonated PS template without PANI or Fe3+ were pyrolyzed at the same condition as the Fe, N-C/carbon-900 and were denoted as N-C/carbon-900 and carbon sphere-900, respectively.

Acknowledgment

This work was partially supported by National Natural Science Foundation of China (Nos. 51273158, 21303131).

Appendix A. Supplementary data

Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.cclet.2016.12.006.

References
[1] Z.H. Wen, S.Q. Ci, Y. Hou, J.H. Chen, Facile one-pot, one-step synthesis of a carbon nanoarchitecture for an advanced multifunctonal electrocatalyst. Angew.Chem.Int.Ed. 53 (2014) 6496–6500. DOI:10.1002/anie.201402574
[2] Y.Y. Shao, S. Park, J. Xiao, Electrocatalysts for nonaqueous lithium-air batteries:status, challenges, and perspective. ACS Catal. 2 (2012) 844–857. DOI:10.1021/cs300036v
[3] Y. Nie, L. Li, Z.D. Wei, Recent advancements in Pt and Pt-free catalysts for oxygen reduction reaction. Chem.Soc.Rev. 44 (2015) 2168–2201. DOI:10.1039/C4CS00484A
[4] Y.H. Bing, H.S. Liu, L. Zhang, D. Ghosh, J.J. Zhang, Nanostructured Pt-alloy electrocatalysts for PEM fuel cell oxygen reduction reaction. Chem.Soc.Rev. 39 (2010) 2184–2202. DOI:10.1039/b912552c
[5] X.W. Yu, S.Y. Ye, Recent advances in activity and durability enhancement of Pt/C catalytic cathode in PEMFC:part Ⅰ.Physico-chemical and electronic interaction between Pt and carbon support and activity enhancement of Pt/C catalyst. J.Power Sources 172 (2007) 133–144. DOI:10.1016/j.jpowsour.2007.07.049
[6] G. Wu, K.L. More, C.M. Johnston, P. Zelenay, High-performance electrocatalysts for oxygen reduction derived from polyaniline, iron, and cobalt. Science 332 (2011) 443–447. DOI:10.1126/science.1200832
[7] B.C.H. Steele, A. Heinzel, Materials for fuel-cell technologies. Nature 414 (2001) 345–352. DOI:10.1038/35104620
[8] L. Birry, J.H. Zagal, J.P. Dodelet, Does CO poison Fe-based catalysts for ORR. Electrochem.Commun. 12 (2010) 628–631. DOI:10.1016/j.elecom.2010.02.016
[9] H.A. Gasteiger, S.S. Kocha, B. Sompalli, F.T. Wagner, Activity benchmarks and requirements for Pt Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Appl.Catal.B 56 (2005) 9–35. DOI:10.1016/j.apcatb.2004.06.021
[10] M.K. Debe, Electrocatalyst approaches and challenges for automotive fuel cells. Nature 486 (2012) 43–51. DOI:10.1038/nature11115
[11] J.P. Collman, J.I. Brauman, T.R. Halbert, K.S. Suslick, Nature of O2 and CO binding to metalloporphyrins and heme proteins. Proc.Natl.Acad.Sci.U.S.A. 73 (1976) 3333–3337. DOI:10.1073/pnas.73.10.3333
[12] S.Q. Song, Y.R. Liang, Z.H. Li, Effect of pore morphology of mesoporous carbons on the electrocatalytic activity of Pt nanoparticles for fuel cell reactions. Appl.Catal.B 98 (2010) 132–137. DOI:10.1016/j.apcatb.2010.05.021
[13] S.Q. Song, S.B. Yin, Z.H. Li, Effect of pore diameter of wormholelike mesoporous carbon supports on the activity of Pt nanoparticles towards hydrogen electrooxidation. J.Power Sources 195 (2010) 1946–1949. DOI:10.1016/j.jpowsour.2009.10.009
[14] K.Y. Chen, X.B. Huang, C.Y. Wan, H. Liu, Efficient oxygen reduction catalysts formed of cobalt phosphide nanoparticle decorated heteroatom-doped mesoporous carbon nanotubes. Chem.Commun. 51 (2015) 7891–7894. DOI:10.1039/C5CC02028J
[15] J. Tang, J. Liu, C.L. Li, Synthesis of nitrogen-doped mesoporous carbon spheres with extra-large pores through assembly of diblock copolymer micelles. Angew.Chem.Int.Ed. 54 (2015) 588–593.
[16] J. Wei, Y. Liang, X.Y. Zhang, Controllable synthesis of mesoporous carbon nanospheres and Fe-N/carbon nanospheres as efficient oxygen reduction electrocatalysts. Nanoscale 7 (2015) 6247–6254. DOI:10.1039/C5NR00331H
[17] K. Wan, G.F. Long, M.Y. Liu, Nitrogen-doped ordered mesoporous carbon: synthesis and active sites for electrocatalysis of oxygen reduction reaction. Appl.Catal.B 165 (2015) 566–571. DOI:10.1016/j.apcatb.2014.10.054
[18] R.F. Wang, H. Wang, T.b. Zhou, The enhanced electrocatalytic activity of okara-derived N-doped mesoporous carbon for oxygen reduction reaction. J. Power Sources 274 (2015) 741–747. DOI:10.1016/j.jpowsour.2014.10.049
[19] F. Rodríguez-Reinoso, The role of carbon materials in heterogeneous catalysis. Carbon 36 (1998) 159–175. DOI:10.1016/S0008-6223(97)00173-5
[20] L. Calvillo, M. Lázaro, E. García-Bordejé, Platinum supported on functionalized ordered mesoporous carbon as electrocatalyst for direct methanol fuel cells. J.Power Sources 169 (2007) 59–64. DOI:10.1016/j.jpowsour.2007.01.042
[21] L.M. Dai, Y.H. Xue, L.T. Qu, H.J. Choi, J.B. Baek, Metal-free catalysts for oxygen reduction reaction. Chem.Rev. 115 (2015) 4823–4892. DOI:10.1021/cr5003563
[22] Z.H. Zhao, M.T. Li, L.P. Zhang, L.M. Dai, Z.H. Xia, Design principles for heteroatom-doped carbon nanomaterials as highly efficient catalysts for fuel cells and metal-air batteries. Adv.Mater. 27 (2015) 6834–6840. DOI:10.1002/adma.201503211
[23] J.J. Duan, S. Chen, M. Jaroniec, S.Z Qiao, Heteroatom-doped graphene-based materials for energy-relevant electrocatalytic processes. ACS Catal. 5 (2015) 5207–5234. DOI:10.1021/acscatal.5b00991
[24] L. Hao, S.S. Zhang, R.J. Liu, Electrocatalysts:bottom-up construction of triazine-based frameworks as metal-free electrocatalysts for oxygen reduction reaction(Adv.Mater.20/2015). Adv.Mater. 27 (2015) 3189. DOI:10.1002/adma.201570138
[25] J. Yang, H.Y. Sun, H.Y. Liang, A highly efficient metal-free oxygen reduction electrocatalyst assembled from carbon nanotubes and graphene. Adv.Mater. 28 (2016) 4606–4613. DOI:10.1002/adma.v28.23
[26] K.P. Gong, F. Du, Z.H. Xia, M. Durstock, L.M. Dai, Nitrogen-doped carbon nanotube arrays with high electrocatalytic activity for oxygen reduction. Science 323 (2009) 760–764. DOI:10.1126/science.1168049
[27] E. Proietti, F. Jaouen, M. Lefèvre, Iron-based cathode catalyst with enhanced power density in polymer electrolyte membrane fuel cells. Nat. Commun. 2 (2011) 416. DOI:10.1038/ncomms1427
[28] W.X. Yang, X.J. Liu, X.Y. Yue, J.B. Jia, S.J. Guo, Bamboo-like carbon nanotube/Fe3C nanoparticle hybrids and their highly efficient catalysis for oxygen reduction. J. Am.Chem.Soc. 137 (2015) 1436–1439. DOI:10.1021/ja5129132
[29] W. Ding, L. Li, K. Xiong, Shape fixing via salt recrystallization:a morphology-controlled approach to convert nanostructured polymer to carbon nanomaterial as a highly active catalyst for oxygen reduction reaction. J.Am.Chem.Soc. 137 (2015) 5414–5420. DOI:10.1021/jacs.5b00292
[30] Y. Hu, J.O. Jensen, W. Zhang, Hollow spheres of iron carbide nanoparticles encased in graphitic layers as oxygen reduction catalysts. Angew.Chem.Int. Ed. 53 (2014) 3675–3679. DOI:10.1002/anie.v53.14
[31] L. Wang, A. Ambrosi, M. Pumera, Metal-free catalytic oxygen reduction reaction on heteroatom-doped graphene is caused by trace metal impurities. Angew.Chem.Int.Ed. 52 (2013) 13818–13821. DOI:10.1002/anie.201309171
[32] Q. Wang, Z.Y. Zhou, Y.J. Lai, Phenylenediamine-based FeNx/C catalyst with high activity for oxygen reduction in acid medium and its active-site probing. J.Am.Chem.Soc. 136 (2014) 10882–10885. DOI:10.1021/ja505777v
[33] Y.Y. Liang, Y.G. Li, H.L. Wang, Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction. Nat.Mater. 10 (2011) 780–786. DOI:10.1038/nmat3087
[34] M. Yang, X.X. Yao, G. Wang, H.J. Ding, A simple method to synthesize sea urchin-like polyaniline hollow spheres. Colloids Surf.A 324 (2008) 113–116. DOI:10.1016/j.colsurfa.2008.04.004
[35] Z.K. Meng, Q. Wang, X.Z. Qu, Papillae mimetic hairy composite spheres towards lotus leaf effect coatings. Polymer 52 (2011) 597–601. DOI:10.1016/j.polymer.2010.12.020
[36] H.S. Byun, R.P. Burford, A.G. Fane, Sulfonation of cross-linked asymmetric membranes based on polystyrene and divinylbenzene. Appl.Polym.Sci. 52 (1994) 825–835. DOI:10.1002/app.1994.070520612
[37] O. Ozer, A. Ince, B. Karagoz, N. Bicak, Crosslinked PS-DVB microspheres with sulfonated polystyrene brushes as new generation of ion exchange resins. Desalination 309 (2013) 141–147. DOI:10.1016/j.desal.2012.09.024
[38] M. Yang, J. Ma, Z.W. Niu, Synthesis of spheres with complex structures using hollow latex cages as templates. Adv.Funct.Mater. 15 (2005) 1523–1528. DOI:10.1002/(ISSN)1616-3028
[39] A.C. Ferrari, J.C. Meyer, V. Scardaci, Raman spectrum of graphene and graphene layers. Phys.Rev.Lett. 97 (2006) 187401. DOI:10.1103/PhysRevLett.97.187401
[40] V.M. Dhavale, S.S. Gaikwad, L. George, R.N. Devi, S. Kurungot, Nitrogen-doped graphene interpenetrated 3D Ni-nanocages:efficient and stable water-to-dioxygenelectrocatalysts. Nanoscale 6 (2014) 13179–13187. DOI:10.1039/C4NR03578J
[41] S.K. Singh, V.M. Dhavale, S. Kurungot, Surface-tuned Co3O4 nanoparticles dispersed on nitrogen-doped graphene as an efficient cathode electrocatalyst for mechanical rechargeable zinc-air battery application. ACS Appl.Mater. Interfaces 7 (2015) 21138–21149. DOI:10.1021/acsami.5b04865
[42] N. Bahlawane, P.H.T. Ngamou, V. Vannier, Tailoring the properties and the reactivity of the spinel cobalt oxide. Phys.Chem.Chem.Phys. 11 (2009) 9224–9232. DOI:10.1039/b910707j
[43] J.W. Xiao, Q. Kuang, S.H. Yang, Surface structure dependent electrocatalytic activity of Co3O4 anchored on graphene sheets toward oxygen reduction reaction. Sci.Rep. 3 (2013) 2300. DOI:10.1038/srep02300
[44] D.H. Deng, L. Yu, X.L. Pan, Size effect of graphene on electrocatalytic activation of oxygen. Chem.Commun. 47 (2011) 10016–10018. DOI:10.1039/c1cc13033a
[45] W.J. Yuan, Y. Zhou, Y.R. Li, The edge-and basal-plane-specific electrochemistry of a single-layer graphene sheet. Sci.Rep. 3 (2013) 2248. DOI:10.1038/srep02248
[46] Q.X. Guo, Y. Xie, X.J. Wang, Characterization of well-crystallized graphitic carbon nitride nanocrystallites via a benzene-thermal route at low temperatures. Chem.Phys.Lett. 380 (2003) 84–87. DOI:10.1016/j.cplett.2003.09.009
[47] W.J. Liu, K. Tian, Y.R. He, H. Jiang, H.Q. Yu, High-yield harvest of nanofibers/mesoporous carbon composite by pyrolysis of waste biomass and its application for high durability electrochemical energy storage. Environ.Sci. Technol. 48 (2014) 13951–13959. DOI:10.1021/es504184c
[48] Y. Zheng, Y. Jiao, L. Ge, M. Jaroniec, S.Z. Qiao, Two-step boron and nitrogen doping in graphene for enhanced synergistic catalysis. Angew.Chem.Int.Ed. 125 (2013) 3192–3198. DOI:10.1002/ange.201209548
[49] Y.H. Su, Y.H. Zhu, H.L. Jiang, Cobalt nanoparticles embedded in N-doped carbon as an efficient bifunctional electrocatalyst for oxygen reduction and evolution reactions. Nanoscale 6 (2014) 15080–15089. DOI:10.1039/C4NR04357J
[50] Y. Zhao, K. Watanabe, K. Hashimoto, Self-supporting oxygen reduction electrocatalysts made from a nitrogen-rich network polymer. J.Am.Chem. Soc. 134 (2012) 19528–19531. DOI:10.1021/ja3085934
[51] L.J. Yang, S.J. Jiang, Y. Zhao, Boron-doped carbon nanotubes as metal-free electrocatalysts for the oxygen reduction reaction. Angew.Chem.Int.Ed. 123 (2011) 7270–7273. DOI:10.1002/ange.v123.31
[52] Z. Yang, Z. Yao, G.F. Li, Sulfur-doped graphene as an efficient metal-free cathode catalyst for oxygen reduction. ACS Nano 6 (2012) 205–211. DOI:10.1021/nn203393d
[53] Y. Zhao, L.J. Yang, S. Chen, Can boron and nitrogen co-doping improve oxygen reduction reaction activity of carbon nanotubes. J.Am.Chem.Soc. 135 (2013) 1201–1204. DOI:10.1021/ja310566z