Next Article in Journal
Design and Microwave-Assisted Synthesis of TiO2-Lanthanides Systems and Evaluation of Photocatalytic Activity under UV-LED Light Irradiation
Next Article in Special Issue
Novel Complex Titanium NASICON-Type Phosphates as Acidic Catalysts for Ethanol Dehydration
Previous Article in Journal
Potential and Scale-Up of Pore-Through-Flow Membrane Reactors for the Production of Prebiotic Galacto-Oligosaccharides with Immobilized β-Galactosidase
Previous Article in Special Issue
Inversion of the Photogalvanic Effect of Conductive Polymers by Porphyrin Dopants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Electrocatalytic Activity of Cobalt-Doped Ceria Embedded on Nitrogen, Sulfur-Doped Reduced Graphene Oxide as an Electrocatalyst for Oxygen Reduction Reaction

by
Manickam Sridharan
1,
Thandavarayan Maiyalagan
1,*,
Gasidit Panomsuwan
2,3 and
Ratchatee Techapiesancharoenkij
2,3
1
Electrochemical Energy Laboratory, Department of Chemistry, SRM Institute of Science and Technology, Kattankulathur, Chennai 603203, Tamil Nadu, India
2
International Collaborative Education Program for Materials Technology, Education and Research (ICE-Matter), Bangkok 10900, Thailand
3
Department of Materials Engineering, Faculty of Engineering, Kasetsart University, Bangkok 10900, Thailand
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(1), 6; https://doi.org/10.3390/catal12010006
Submission received: 4 December 2021 / Revised: 16 December 2021 / Accepted: 17 December 2021 / Published: 22 December 2021
(This article belongs to the Special Issue Catalysts in Energy Applications)

Abstract

:
N, S-doped rGO was successfully synthesized and embedded Co-doped CeO2 via hydrothermal synthesis. The crystal structure, surface morphology and elemental composition of the prepared catalyst were studied by XRD, Raman spectra, SEM, TEM and XPS analyses. The synthesized electrocatalyst exhibits high onset and halfwave potential during the ORR. This result shows that a combination of N- and S-doped rGO and Co-doped CeO2 leads to a synergistic effect in catalyzing the ORR in alkaline media. Co–CeO2/N, S–rGO displays enhanced ORR performance compared to bare CeO2. The superior stability of the prepared catalyst implies its potential applications beyond fuel cells and metal–air batteries.

1. Introduction

Developing an electrocatalyst for the oxygen reduction reaction by using nonprecious metal to replace the Pt group metal-based catalyst is an important strategic task for fuel cells or metal–air batteries because of its crucial role in both [1,2,3]. Currently, extensive research has been carried out to find effective nonprecious metal catalysts for the ORR [4,5]. Metal oxide-based catalysts are used for support in the ORR. Metal oxides such as CeO2 [6,7,8,9], MnO2 [10,11,12], TiO2 [13,14] and ZnO [15] have been used as alternatives for the Pt-based electrocatalyst for the ORR. Metal oxides act as promoters during the ORR due to their facile oxygen release and storage capacity, as well as their well-known corrosion resistance [16,17].
Cerium oxide-based materials have shown promising electrocatalytic performance because of their unique structural features. CeO2 has been widely used in fuel cells (MOR, ORR) [18,19,20]. Precious metals like Pt and Pd supported on CeO2 exhibit good catalytic activity toward the ORR [21,22,23,24], but precious metals are quite challenging for commercialization [25,26]. Although there is much research dealing with CeO2-supported catalysts for the ORR, there are very few studies focusing on the importance of oxygen vacancy in doped CeO2 [27,28,29]. Furthermore, the abundance of oxygen vacancies in CeO2 favors O2 adsorption and oxygen conductivity due to the coexistence of Ce4+ and Ce3+, which contribute to the CeO2-supported electrocatalyst in the ORR [30,31]. However, the poor electronic conductivity of CeO2 limits its ORR performance. Depositing CeO2 on conducting carbon-based materials has proven to be the most effective method for improving its dispersibility and conductivity [32,33].
In recent years, a number of reduced graphene oxides with metal oxide hybrid catalysts have been investigated as ORR catalysts [34,35,36]. Among those, heteroatom-doped graphene oxide with metal oxide hybrid materials has emerged as a promising ORR electrocatalyst with huge potential [37,38,39,40,41]. In these cases, N doping in graphene oxide provides a positive charge on neighboring carbon atoms through structural distortion, which leads to an increase of O2 adsorption and higher defect density, resulting in more active sites to enhance the ORR performance [42,43,44]. Moreover, the atomic size of sulfur is larger than carbon and may create more defects and strain, which enhance the O2 adsorption and accelerate the weakening of the O–O bond [45,46]. Therefore, the synergistic impact of nitrogen and sulfur co-doping on rGO enhances the electrocatalytic performance towards the ORR. According to the findings of several studies, it can be concluded that transition metal-doped CeO2/N, S–rGO support could be an effective strategy to improve electrocatalytic activity toward the ORR.
Herein, we prepared a Co–CeO2/N, S-doped rGO composite with the hydrothermal method. The synthesized sample’s crystal structure, morphology and elemental oxidation state were confirmed by XRD, Raman spectra, SEM, TEM and XPS analyses. Then, the electrochemical performance of the Co–CeO2/N, S–rGO composite was performed in alkaline media.

2. Results and Discussion

2.1. XRD

The XRD patterns of Co-doped CeO2 and Co–CeO2/N, S–rGO composites are shown in Figure 1. The diffraction pattern of CeO2 2 θ values at 28.47 ° , 33.0°, 47.46°, 56.31°, 59.12°, 69.41° and 76.70° correspond to the (111), (200), (220), (311), (322), (400) and (331) planes of FCC CeO2 (JCPDS no: 34-0394) [47], respectively. The XRD patterns of Co–CeO2 reveal that during the Co doping, the peak intensity decreases, and the peaks slightly shift to a higher angle in comparison with bare CeO2, which indicates the doping of the Co ion into the CeO2 lattice. The peak intensity of the Co-doped CeO2 is much weaker than CeO2, which can be created by the ionic radii difference between CeO2 and Co-doped CeO2. Additionally, the ionic radius of Co2+/3+ is 0.75−0.9 Å and Ce4+ is 0.97Å [48]. However, the further increase in the Co concentration to 4.5% of the peak intensity also increases, which indicates grain size increase after the 4.5% Co ion is doped into the CeO2. From Figure 1c, GO exhibits a typical characteristic plane of (002) appearing at 10.1°. After the hydrothermal treatment, the synthesized N, S–rGO shows a characteristic peak of (002) at 24.5 and the strong peak (002) at 10.1° disappears, which indicates the reduction of rGO. Figure 1b exhibits the diffraction peaks of both CeO2 and N,S–rGO, which confirm the formation of Co-doped CeO2 and the N, S–rGO hybrid.

2.2. Raman Analysis

The doping of the Co ion in CeO2 induces oxygen defects in CeO2 (Figure 2a), which is further characterized by the Raman analysis. The main characteristic peak at 462 cm−1 refers to the F2g vibration mode cubic CeO2, and a shoulder peak at 595 cm−1 corresponds to the oxygen vacancy. Compared to the bare CeO2, the peak intensity of 462 cm−1 decreases and the peak of 595 cm−1 is increased after the doping of Co. These results suggest that the Co doping has led to the enhancement of the oxygen vacancy in CeO2. According to the literature, the Raman peak at 830 cm−1 corresponds to the peroxide species [49]. Figure 2b shows the Raman spectra of GO, N, S–rGO and the Co–CeO2/N, S–rGO composite. All samples exhibit two characteristic peaks of the D band at 1350 cm−1 and the G band at 1579 cm−1 [50,51]. The D band is assigned to the sp3 structural defect sites, and the G band is associated with the sp2-bonded graphitic carbon atom. The intensity ratio of the D and G bands (ID/IG) is typically used to measure defects in carbon materials. The ID/IG values were 0.941, 1.100 and 1.114 for GO, N, S–rGO and 3% Co–CeO2/N, S–rGO, respectively. The 3% Co–CeO2/N, S–rGO exhibits a higher ID/IG value than the rGO, suggesting that the N, S-doping and embedded Co–CeO2 create more defects in the carbon matrix and also create more active sites during the ORR. Additionally, the Raman spectra Co–CeO2/N, S–rGO exhibit the typical peak of CeO2, located at 462 cm−1, and the D and G bands of N, S–rGO, which conform to the coexistence of CeO2 and N, S–rGO. This result demonstrates the interaction between rGO and Co–CeO2 during the synthesis.

2.3. FESEM and TEM Analyses

The morphologies of 3% Co–CeO2/N, S–rGO were studied by SEM and TEM analyses. Figure 3a shows the sheet-like morphology of rGO. Figure 3b shows that the Co–CeO2 nanoparticles are uniformly anchored on the N, S–rGO to form the Co–CeO2/N, S–rGO. The corresponding TEM image shows that CeO2 nanoparticles were anchored on the sheet-like structure of N, S–rGO. In the HRTEM image (Figure 3c), the lattice fringes of 0.31 nm and 0.26 nm correspond to the CeO2 (111) and (002) N, S–rGO, respectively.

2.4. XPS Analysis

The XPS analysis was performed to further measure the elemental composition and chemical valence of the 3% Co–CeO2/N, S–rGO. Figure 4a shows the Ce 3d core level spectra of CeO2. The spin orbit doublets of Ce 3d3/2 and Ce 3d5/2 were labelled as U and V, respectively. The characteristic binding energy peaks of U‴ (916.2 eV), V‴ (899.7 eV), U𠌲 (907.6 eV), V″ (889.2 eV), U (901.7 eV) and V (883 eV) are ascribed to Ce4+ species. While the peaks centered at Uo (899.1 eV), Vo (881.1 eV), U′ (904.9 eV) and V′ (886.1 eV) were attributed to the Ce3+ [52,53]. Figure 4b shows the core level spectra of Co, which exhibits the core level of 2p3/2 (800 eV) and 2p1/2 (796.5 eV) for Co–O bonding. Additionally, the higher energy level of Co 2p3/2 and Co 2p1/2 corresponds to bivalent cobalt ions. As displayed in Figure 4c, the O 1s core level spectra split into three peaks, with the binding energy peak of 529.1 eV assigned as the lattice oxygen, the peak of 530.5 eV corresponding to the chemisorbed oxygen and the binding energy peak around 532.2 eV corresponding to the oxygen vacancy in CeO2 [54,55]. The C 1s (Figure 4d) core level spectra exhibit three different binding energy peaks, 284.7, 286.0 and 289.1 eV, which correspond to the C = C–O, C–N/C–S and C–O, respectively. The N 1s core level spectrum (Figure 4e) displays three peaks located at 398.3 eV for pyridinic N, 400 eV for pyrrolic and 400.8 eV for graphitic N. Figure 4f is the core level spectra of S 2p, where the characteristic peaks at 163.25 eV and 164.80 eV are assigned to the S 2p3/2 and S 2p1/2 of thiophene S, and the peak of 167.8 eV corresponds to the SOx group [56].

2.5. Electrochemical Performance of Co–CeO2 and Co–CeO2/N, S–rGO Composites

The ORR performance of the synthesized catalyst was studied by using a cyclic voltammogram (CV), which was performed at a scan rate of 50 mV/s with O2-saturated KOH. Figure 5 shows the CV measurements of the prepared catalysts with a different percent of Co dopant. The cathodic oxygen reduction onset potential increases in the following order as 3% Co–CeO2/N, S–rGO > 1.5% Co–CeO2/N, S–rGO > 4.5% Co–CeO2/N, S–rGO > CeO2/N, S–rGO. Moreover, the 3% Co–CeO2/N, S–rGO exhibits a higher current density than the samples. This result demonstrates an enhanced ORR activity due to the Co doping on CeO2 and composite formation with N, S–rGO.
The electrocatalytic ORR performances of the synthesized catalysts were performed at a scan rate of 10 mV/s in O2-saturated 0.1 M KOH. Figure 6a shows the ORR polarization curves of the bare CeO2 and the different percentage of Co-doped CeO2. At 3% Co–CeO2, the halfwave potential shifted more positively and Jlim increased. However, when the Co doping concentration increased to 4.5%, the ORR performance decreased, showing that the optimal concentration is 3%. This result shows that Co doping to CeO2 improves the ORR performance, which can be attributed to the oxygen vacancy creation through Co doping and the coexistence of Ce3+ and Ce4+. The optimized 3% Co–CeO2/N, S–rGO hybrid electrocatalyst shows better ORR performance (Table 1) compared to bare CeO2 and 3% Co–CeO2. After the combination of Co–CeO2 and N, S–rGO, the 3% Co–CeO2/N, S–rGO has a high limiting current density (JL = 4.2 mA/cm2) as well as a higher halfwave potential (E1/2 = 0.73 V vs RHE) and onset potential (E0 = 0.87 V vs RHE). In comparison with the 3% Co–CeO2/N, S–rGO (Figure 6b), other samples show electrocatalytic activity with limited current density (1.4 mA/cm2 for CeO2, 3.4 mA/cm2 for CeO2/N, S–rGO, 3.6 mA/cm2 for 1.5% Co–CeO2/N, S–rGO and 3.0 mA/cm2 for 4.5% Co–CeO2/N, S–rGO), halfwave potential (0.45 V for CeO2, 0.67 V for CeO2/N, S–rGO, 0.70 V for 1.5% Co–CeO2/N, S–rGO and 0.71 V for 4.5% Co–CeO2/N, S–rGO) and onset potential (0.67 V for CeO2, 0.82 V for CeO2/N, S–rGO, 0.86 V for 1.5% Co–CeO2/N, S–rGO and 0.89 V for 4.5% Co–CeO2/N, S–rGO). Based on the ORR results, the 3% Co–CeO2/N, S–rGO exhibits superior electrocatalytic activity among the synthesized catalysts.
In order to further understand the electron transfer kinetics of the 3% Co–CeO2/N, S–rGO and the Co–CeO2/N, S–rGO composites (Figure 6c and Figure 7a–c) during the ORR, RDE measurements were carried out at 10 mV s−1 and different electrode rotation speeds in O2-saturated 0.1 M KOH solution. When increasing the rotation speed from 400 rpm to 1600 rpm, the diffusion current densities increased, while the onset potential was unchanged. To estimate the electron transfer number per O2 molecule during the ORR by the Koutecky–Levich equation, as follows [59]:
1/j = 1/jk + 1/jd = 1/jk + 1/(Bω1/2)
B = 0.62nFCD2/3 ν−1/6
where jk, jd and jl is kinetic, diffusion limiting and total current density, respectively; B and ω are the proportionality coefficient and angular velocity of the disk, respectively; C is the bulk concentration of oxygen; D is the diffusion coefficient of oxygen in 0.1 M KOH; F is the Faraday constant (96,485 C·mol−1); ν is the kinematic viscosity, and n is the number of electrons transferred per oxygen molecule. Figure 6d shows the K–L plots of 3% Co–CeO2/N, S–rGO, which exhibit a linear and parallel relationship with different potentials and rpms, suggesting that it abides by a first-order kinetic reaction during the ORR.
The enhanced ORR performance can be presumed by considering the following facts: (1) The defects such as oxygen vacancy play an essential role in the ORR; (2) The synergistic impact of nitrogen and sulfur co-doping on rGO enhances the electrocatalytic performance towards ORR:
2CeO2→2CeO2−x + xO2 (0 ≤ x ≤ 1)
CeO2 possesses a remarkable ability to reversibly exchange oxygen. CeO2 is frequently used as an oxygen buffer due to the redox reaction between Ce4+/Ce3+ (Equation (3)). It can store oxygen in the oxygen-rich environment as well as produce oxygen in oxygen insufficient conditions. When oxygen escapes from the structure of CeO2, the reduction of Ce4+ to Ce3+ occurs. The oxidation of Ce3+ to Ce4+ occurs while oxygen is adsorbed on the CeO2 surface. Through N doping in graphene oxide, it creates a positive charge on neighboring carbon atoms and generates a structural distortion, leading to increased O2 adsorption and higher defect density, resulting in more active sites to improve the ORR performance. Furthermore, because sulfur has a greater atomic size than carbon, it may induce more flaws and strain, enhancing O2 adsorption while simultaneously enhancing the weakening of the O–O bond. As a result, the synergistic effect of nitrogen and sulfur co-doping on rGO and Co-doped CeO2 improves the ORR electrocatalytic performance. Moreover, oxygen buffering capacity is highly responsible for the enhanced ORR activity of the Co–CeO2/N, S–rGO composite.
Durability is a crucial factor for the ORR, which can be studied by ADT (accelerated durability test) 0.1 M KOH solution and a scan rate of 10 mV s−1 (vs RHE) with an electrode rotation speed of 1600 rpm. As shown in Figure 8, after 5000 cycles the 3% Ce–CeO2/N, S–rGO exhibits a much less negative halfwave potential shift ( Δ E1/2 = ~4 mV), which demonstrate the long-term stability of 3% Ce–CeO2/N, S–rGO.

3. Materials and Methods

3.1. Materials

Graphite powder, cerium nitrate hexahydrate and cobalt chloride hexahydrate were purchased from Sigma-Aldrich. Thiourea was purchased from Alfa Aesar. All required solutions were prepared from Milli-Q reagent water.

3.2. Synthesis of N, S–rGO

Graphene oxide (GO) was prepared by using the modified Hummers method [60,61]. Initially, GO suspension was prepared by dispersing 60 mg of GO in 40 mL of DI water ultrasonically for 2 h. To the above suspension, 250 mg of thiourea was slowly added and kept for stirring (6 h). Then, the solution was transferred to a Teflon-lined stainless steel, and then heated at 180 °C for 12 h. The obtained precipitate was washed several times with DI water and dried at 60 °C for 24 h.

3.3. Synthesis of Co–CeO2/N, S–rGO Nanocomposites

Co–CeO2/N, S–rGO nanocomposites were synthesized through a hydrothermal method by dispersing 60 mg of GO in 30 mL of 2 M sodium hydroxide solution and kept stirring for about 1 h. To the above solution, 50 mM of cerium nitrate hexahydrate (8 mL) was added slowly and kept stirring for 30 min. To the mixture, cobalt nitrate hexahydrate was added by varying the Co (1.5% and 4.5%) concentration for doping. Then, the hydrothermal reaction was carried out at 120 °C for 12 h after transferring the solution into a Teflon-lined stainless steel autoclave. Subsequently, the precipitate was washed several times with DI water and dried at 60 °C for 24 h.

3.4. Characterization

Phase purity of the sample was examined by using powder X-ray diffraction (XRD) (PAN Analytical, Almelo, The Netherlands). Morphology of the sample was observed by using SEM (FEI Quanta FEG 200, Zürich. Switzerland). The JEM Fb-2000 instrument was used to obtain the transmission electron microscopy (TEM) images. For the XPS spectra, a Perkin Elmer PHI 550 spectrometer (Norwalk, CA, USA) was used. A Raman spectrometer (LAB RAMAN HR Evolution HORIBA, Paris, France) was used to analyze the Raman spectra of the sample.

3.5. Electrochemical Techniques

The electrochemical characterizations were studied by using a CHI 760E electrochemical workstation. The experimental setup consists of Ag/AgCl (sat.KCl) as reference electrode, a thin Pt wire as counter electrode and a glassy carbon-rotating disc electrode (GC-RDE, 0.196 cm2) as a working electrode. An amount of 20% Pt/C (standard catalyst) bought from Alfa Aesar, (Chennai, india)(Pt/C), was used for the comparison of the ORR studies. Then, the catalyst ink was prepared by dispersing 2.5 mg of electrocatalyst in 1 mL of ethanol and water (1:1), 20 µL Nafion (5 wt.% solution), followed by sonication for 30 min. Then, 10 µL of the catalyst ink was dropped on the surface of the polished RDE-GC and dried under an IR lamp.

4. Conclusions

In summary, we have synthesized the Co–CeO2/N, S–rGO composite with the hydrothermal method. The structural, morphological and elemental compositions have been performed on this system by XRD, Raman spectra, FESEM, TEM and XPS. The electrochemical ORR performance of the catalyst was studied by measuring CV and LSV in oxygen-saturated 0.1 M KOH solution. The electrochemical ORR performance measurements showed that the Co doping on CeO2 and the composite with N, S–rGO exhibited an apparent catalytic activity in alkaline media, and the synthesized 3% Co–CeO2/N, S–rGO exhibited the highest electrocatalytic performance. The enhancement is ascribed to the Co doping, offering oxygen vacancy and the utilization efficiency of active sites, with the N- and S-doped conductive rGO support aiding the charge transport during the ORR. We believe that the Co–CeO2/N, S–rGO composite will produce very promising and low-cost catalyst fuel cells and metal–air batteries.

Author Contributions

Conceptualization, M.S. and T.M.; methodology, M.S.; software, M.S. and R.T.; validation, T.M., G.P. and R.T.; formal analysis, M.S. and T.M.; investigation, M.S. and T.M.; resources, M.S. and T.M.; data curation, T.M., G.P. and R.T.; writing—original draft preparation, M.S.; writing—review and editing, M.S. and T.M.; visualization, M.S. and T.M.; supervision, T.M.; project administration, T.M., G.P. and R.T.; funding acquisition, T.M., G.P. and R.T. All authors have read and agreed to the published version of the manuscript.

Funding

The authors also graciously acknowledge the financial support provided by Scheme for Promotion of Academic and Research Collaboration (SPARC grant No. SPARC/2018-2019/P1122/SL) and the SRM Institute of Science and Technology for providing financial support under the research grant RU-SATU Joint Research 2021: SRMIST 15.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data availability sets could be obtained by contacting authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, G.; Zelenay, P. Nanostructured nonprecious metal catalysts for oxygen reduction reaction. Acc. Chem. Res. 2013, 46, 1878–1889. [Google Scholar] [CrossRef] [PubMed]
  2. Gewirth, A.A.; Varnell, J.A.; DiAscro, A.M. Nonprecious metal catalysts for oxygen reduction in heterogeneous aqueous systems. Chem. Rev. 2018, 118, 2313–2339. [Google Scholar] [CrossRef]
  3. Zhan, Y.; Xie, F.; Zhang, H.; Jin, Y.; Meng, H.; Chen, J.; Sun, X. Highly Dispersed nonprecious metal catalyst for oxygen reduction reaction in proton exchange membrane fuel cells. ACS Appl. Mater. Interfaces 2020, 12, 17481–17491. [Google Scholar] [CrossRef]
  4. Wang, Y.; Li, J.; Wei, Z. Transition-metal-oxide-based catalysts for the oxygen reduction reaction. J. Mater. Chem. A 2018, 6, 8194–8209. [Google Scholar] [CrossRef]
  5. Li, H.; Nørskov, J.K. Effects of a conductive support on the bonding of oxygen containing molecules to transition metal oxide surfaces. Phys. Chem. Chem. Phys. 2020, 22, 26216–26222. [Google Scholar] [CrossRef] [PubMed]
  6. Bhuvanendran, N.; Ravichandran, S.; Kandasamy, S.; Zhang, W.; Xu, Q.; Khotseng, L.; Maiyalagan, T.; Su, H. Spindle-shaped CeO2/biochar carbon with oxygen-vacancy as an effective and highly durable electrocatalyst for oxygen reduction reaction. Int. J. Hydrogen Energy 2021, 46, 2128–2142. [Google Scholar] [CrossRef]
  7. Masuda, T.; Fukumitsu, H.; Fugane, K.; Togasaki, H.; Matsumura, D.; Tamura, K.; Nishihata, Y.; Yoshikawa, H.; Kobayashi, K.; Mori, T.; et al. Role of cerium oxide in the enhancement of activity for the oxygen reduction reaction at Pt–CeOx nanocomposite electrocatalyst—An in situ electrochemical X-ray absorption fine structure study. J. Phys. Chem. C 2012, 116, 10098–10102. [Google Scholar] [CrossRef] [Green Version]
  8. Soren, S.; Hota, I.; Debnath, A.K.; Aswal, D.K.; Varadwaj, K.S.K.; Parhi, P. Oxygen reduction reaction activity of microwave mediated solvothermal synthesized CeO2/g-C3N4 Nanocomposite. Front. Chem. 2019, 7, 403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Hota, I.; Debnath, A.K.; Muthe, K.P.; Varadwaj, K.S.K.; Parhi, P. A synergistic approach of Vulcan carbon and CeO2 in their composite as an efficient oxygen reduction reaction catalyst. J. Appl. Electrochem. 2020, 50, 1069–1077. [Google Scholar] [CrossRef]
  10. Yang, J.; Wang, J.; Zhu, L.; Gao, Q.; Zeng, W.; Wang, J.; Li, Y. Enhanced electrocatalytic activity of a hierarchical CeO2@MnO2 core-shell composite for oxygen reduction reaction. Ceram. Int. 2018, 44, 23073–23079. [Google Scholar] [CrossRef]
  11. Zhuang, Q.; Ma, N.; Yin, Z.; Yang, X.; Yin, Z.; Gao, J.; Xu, Y.; Gao, Z.; Wang, H.; Kang, J.; et al. Rich surface oxygen vacancies of MnO2 for enhancing electrocatalytic oxygen reduction and oxygen evolution reactions. Adv. Energy Sustain. Res. 2021, 2, 2100030. [Google Scholar] [CrossRef]
  12. Ghosh, S.; Kar, P.; Bhandary, N.; Basu, S.; Maiyalagan, T.; Sardar, S.; Pal, S.K. Reduced graphene oxide supported hierarchical flower like manganese oxide as efficient electrocatalysts toward reduction and evolution of oxygen. Int. J. Hydrogen Energy 2017, 42, 4111–4122. [Google Scholar] [CrossRef]
  13. Barbosa, E.C.M.; Parreira, L.S.; De Freitas, I.C.; Aveiro, L.R.; De Oliveira, D.C.; Dos Santos, M.C.; Camargo, P.H.C. Pt-decorated TiO2 materials supported on carbon: Increasing activities and stabilities toward the ORR by tuning the Pt loading. ACS Appl. Energy Mater. 2019, 2, 5759–5768. [Google Scholar] [CrossRef]
  14. Sravani, B.; Chandrashekar, Y.; Chandana, P.S.; Maiyalagan, T.; Sarma, L.S. Bimetallic PtCu-decorated reduced graphene oxide (RGO)-TiO2 nanocomposite for efficient oxygen reduction reaction. Synth. Met. 2020, 266, 116433. [Google Scholar] [CrossRef]
  15. Sun, Y.; Shen, Z.; Xin, S.; Ma, L.; Xiao, C.; Ding, S.; Li, F.; Gao, G. Ultrafine Co-doped ZnO nanoparticles on reduced graphene oxide as an efficient electrocatalyst for oxygen reduction reaction. Electrochim. Acta 2016, 224, 561–570. [Google Scholar] [CrossRef]
  16. Sun, M.; Liu, H.; Liu, Y.; Qu, J.; Li, J. Graphene-based transition metal oxide nanocomposites for the oxygen reduction reaction. Nanoscale 2015, 7, 1250–1269. [Google Scholar] [CrossRef]
  17. Venkatesan, P.N.; Dharmalingam, S. Synthesis and characterization of Pt, Pt–Fe/TiO2 cathode catalysts and its evaluation in microbial fuel cell. Mater. Renew. Sustain. Energy 2016, 5, 1–9. [Google Scholar] [CrossRef] [Green Version]
  18. Zhu, Y.; Liu, S.; Jin, C.; Bie, S.; Yang, R.; Wu, J. MnOx decorated CeO2 nanorods as cathode catalyst for rechargeable lithium-air batteries. J. Mater. Chem. A 2015, 3, 13563–13567. [Google Scholar] [CrossRef]
  19. Li, Q.; Song, L.; Liang, Z.; Sun, M.; Wu, T.; Huang, B.; Luo, F.; Du, Y.; Yan, C.-H. A Review on Ceo2-based electrocatalyst and photocatalyst in energy conversion. Adv. Energy Sustain. Res. 2021, 2, 2000063. [Google Scholar] [CrossRef]
  20. Altaf, F.; Batool, R.; Rehman, Z.U.; Majeed, H.; Hashmi, M.Z.; Bibi, A.; Ghani, L.; Abbas, G.; Jacob, K. Synthesis and performance study of Pd/CeO2 composite catalyst for enhanced MOR activity. J. Electron. Mater. 2021, 50, 7222–7229. [Google Scholar] [CrossRef]
  21. Chen, J.; Li, Z.; Chen, Y.; Zhang, J.; Luo, Y.; Wang, G.; Wang, R. An enhanced activity of Pt/CeO2/CNT triple junction interface catalyst prepared by atomic layer deposition for oxygen reduction reaction. Chem. Phys. Lett. 2020, 755, 137793. [Google Scholar] [CrossRef]
  22. Xu, F.; Wang, D.; Sa, B.; Yu, Y.; Mu, S. One-pot synthesis of Pt/CeO2/C catalyst for improving the ORR activity and durability of PEMFC. Int. J. Hydrogen Energy 2017, 42, 13011–13019. [Google Scholar] [CrossRef]
  23. Altamirano-Gutiérrez, A.; Fernández, A.; Varela, F.R. Preparation and characterization of Pt-CeO2 and Pt-Pd electrocatalysts for the oxygen reduction reaction in the absence and presence of methanol in alkaline medium. Int. J. Hydrogen Energy 2013, 38, 12657–12666. [Google Scholar] [CrossRef]
  24. Sridharan, M.; Maiyalagan, T. Enhanced oxygen reduction activity of bimetallic Pd–Ag alloy-supported on mesoporous cerium oxide electrocatalysts in alkaline media. New J. Chem. 2021, 45, 22181–22192. [Google Scholar] [CrossRef]
  25. Maiyalagan, T.; Jarvis, K.A.; Therese, S.; Ferreira, P.J.; Manthiram, A. Spinel-type lithium cobalt oxide as a bifunctional electrocatalyst for the oxygen evolution and oxygen reduction reactions. Nat. Commun. 2014, 5, 1–8. [Google Scholar] [CrossRef] [Green Version]
  26. Kumar, A.; Ibraheem, S.; Nguyen, T.A.; Gupta, R.K.; Maiyalagan, T.; Yasin, G. Molecular-MN4 vs atomically dispersed M−N4−C electrocatalysts for oxygen reduction reaction. Coord. Chem. Rev. 2021, 446, 214122. [Google Scholar] [CrossRef]
  27. Meng, G.; Chang, Z.; Cui, X.; Tian, H.; Ma, Z.; Peng, L.; Chen, Y.; Chen, C.; Shi, J. SnO2/CeO2 nanoparticle-decorated mesoporous ZSM-5 as bifunctional electrocatalyst for HOR and ORR. Chem. Eng. J. 2021, 417, 127913. [Google Scholar] [CrossRef]
  28. Chutia, B.; Hussain, N.; Puzari, P.; Jampaiah, D.; Bhargava, S.K.; Matus, E.V.; Ismagilov, I.Z.; Kerzhentsev, M.; Bharali, P. Unraveling the role of CeO2 in stabilization of multivalent Mn species on α-MnO2/Mn3O4/CeO2/C surface for enhanced electrocatalysis. Energy Fuels 2021, 35, 10756–10769. [Google Scholar] [CrossRef]
  29. Liu, D.; Tian, J.; Tang, Y.; Li, J.; Wu, S.; Yi, S.; Huang, X.; Sun, D.; Wang, H. High-power double-face flow Al-air battery enabled by CeO2 decorated MnOOH nanorods catalyst. Chem. Eng. J. 2021, 406, 126772. [Google Scholar] [CrossRef]
  30. Pi, L.; Jiang, R.; Cai, W.; Wang, L.; Wang, Y.; Cai, J.; Mao, X. Bionic preparation of CeO2-encapsulated nitrogen self-doped biochars for highly efficient oxygen reduction. ACS Appl. Mater. Interfaces 2020, 12, 3642–3653. [Google Scholar] [CrossRef] [PubMed]
  31. Hota, I.; Debnath, A.K.; Muthe, K.P.; Varadwaj, K.S.K.; Parhi, P. Towards synergy of rGO and Ni doped CeO2 in their copmposite as efficient catalyst for oxygen reduction reaction. ChemistrySelect 2020, 5, 6608–6616. [Google Scholar] [CrossRef]
  32. Soren, S.; Mohaptra, B.D.; Mishra, S.; Debnath, A.K.; Aswal, D.K.; Varadwaj, K.S.K.; Parhi, P. Nano ceria supported nitrogen doped graphene as a highly stable and methanol tolerant electrocatalyst for oxygen reduction. RSC Adv. 2016, 6, 77100–77104. [Google Scholar] [CrossRef]
  33. Parhi, P.; Soren, S.; Mohapatra, B.D.; Mishra, S.; Debanath, A.K.; Aswal, D.K.; Varadwaj, K.S.K. Nanoceria supported NrGO as highly stable electrocatalyst for oxygen reduction. Acta Cryst. 2017, A73, C1164. [Google Scholar] [CrossRef] [Green Version]
  34. Zhang, Y.; Zhang, X.; Huang, H.; Cai, J.; Huang, B.; Lin, S. Synthesis of TiO2/rGO composites with different morphologies and their electrocatalysis for the oxygen reduction reaction. New J. Chem. 2018, 42, 19755–19763. [Google Scholar] [CrossRef]
  35. Yu, J.; Liu, Z.; Zhai, L.; Huang, T.; Han, J. Reduced graphene oxide supported TiO2 as high performance catalysts for oxygen reduction reaction. Int. J. Hydrogen Energy 2016, 41, 3436–3445. [Google Scholar] [CrossRef]
  36. Boppella, R.; Lee, J.-E.; Mota, F.M.; Kim, J.Y.; Feng, Z.; Kim, D.H. Composite hollow nanostructures composed of carbon-coated Ti3+self-doped TiO2-reduced graphene oxide as an efficient electrocatalyst for oxygen reduction. J. Mater. Chem. A 2017, 5, 7072–7080. [Google Scholar] [CrossRef]
  37. Xiong, C.; Yang, Q.; Dang, W.; Li, M.; Li, B.; Su, J.; Liu, Y.; Zhao, W.; Duan, C.; Dai, L.; et al. Fabrication of eco-friendly carbon microtubes @ nitrogen-doped reduced graphene oxide hybrid as an excellent carbonaceous scaffold to load MnO2 nanowall (PANI nanorod) as bifunctional material for high-performance supercapacitor and oxygen reduction reaction catalyst. J. Power Sources 2020, 447, 227387. [Google Scholar] [CrossRef]
  38. Wang, Z.; Zhou, H.; Xue, J.; Liu, X.; Liu, S.; Li, X.; He, D. Ultrasonic-assisted hydrothermal synthesis of cobalt oxide/nitrogen-doped graphene oxide hybrid as oxygen reduction reaction catalyst for Al-air battery. Ultrason. Sonochem. 2021, 72, 105457. [Google Scholar] [CrossRef]
  39. Ning, R.; Tian, J.; Asiri, A.M.; Qusti, A.H.; Al-Youbi, A.O.; Sun, X. Spinel CuCo2O4 nanoparticles supported on N-doped reduced graphene oxide: A highly active and stable hybrid electrocatalyst for the oxygen reduction reaction. Langmuir 2013, 29, 13146–13151. [Google Scholar] [CrossRef]
  40. Ramirez-Barria, C.S.; Fernandes, D.M.; Freire, C.; Villaro-Abalos, E.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Upgrading the properties of reduced graphene oxide and nitrogen-doped reduced graphene oxide produced by thermal reduction toward efficient ORR electrocatalysts. Nanomaterials 2019, 9, 1761. [Google Scholar] [CrossRef] [Green Version]
  41. Chanda, D.; Dobrota, A.S.; Hnát, J.; Sofer, Z.; Pašti, I.A.; Skorodumova, N.V.; Paidar, M.; Bouzek, K. Investigation of electrocatalytic activity on a N-doped reduced graphene oxide surface for the oxygen reduction reaction in an alkaline medium. Int. J. Hydrogen Energy 2018, 43, 12129–12139. [Google Scholar] [CrossRef]
  42. Khandelwal, M.; Chandrasekaran, S.; Hur, S.H.; Chung, J.S. Chemically controlled in-situ growth of cobalt oxide microspheres on N,S-co-doped reduced graphene oxide as an efficient electrocatalyst for oxygen reduction reaction. J. Power Sources 2018, 407, 70–83. [Google Scholar] [CrossRef]
  43. Ma, X.-X.; He, X.-Q. Cobalt oxide anchored on nitrogen and sulfur dual-doped graphene foam as an effective oxygen electrode catalyst in alkaline media. Appl. Mater. Today 2016, 4, 1–8. [Google Scholar] [CrossRef]
  44. Gao, Q.; Shi, Z.; Xue, K.; Ye, Z.; Hong, Z.; Yu, X.; Zhi, M. Cobalt sulfide aerogel prepared by anion exchange method with enhanced pseudocapacitive and water oxidation performances. Nanotechnology 2018, 29, 215601. [Google Scholar] [CrossRef]
  45. Xia, B.; Yan, Y.; Wang, X.; Lou, X.W.D. Recent progress on graphene-based hybrid electrocatalysts. Mater. Horiz. 2014, 1, 379–399. [Google Scholar] [CrossRef] [Green Version]
  46. Oh, T.; Kim, M.; Park, D.; Kim, J. Synergistic interaction and controllable active sites of nitrogen and sulfur co-doping into mesoporous carbon sphere for high performance oxygen reduction electrocatalysts. Appl. Surf. Sci. 2018, 440, 627–636. [Google Scholar] [CrossRef]
  47. Sridharan, M.; Kamaraj, P.; Vennila, R.; Huh, Y.S.; Arthanareeswari, M. Bio-inspired construction of melanin-like polydopamine-coated CeO2 as a high-performance visible-light-driven photocatalyst for hydrogen production. New J. Chem. 2020, 44, 15223–15234. [Google Scholar] [CrossRef]
  48. Tiwari, S.; Balasubramanian, N.; Biring, S.; Sen, S. Investigation of structural, optical and electronic properties of (Co, Ni) codoped CeO2 nanoparticles. IOP Conf. Ser. Mater. Sci. Eng. 2018, 390, 012001. [Google Scholar] [CrossRef]
  49. Daniel, M.; Loridant, S. Probing reoxidation sites by in situ Raman spectroscopy: Differences between reduced CeO2 and Pt/CeO2. Raman Spectrosc. 2011, 43, 1312–1319. [Google Scholar] [CrossRef]
  50. Naqvi, T.K.; Bharati, M.S.S.; Srivastava, A.K.; Kulkarni, M.M.; Siddiqui, A.M.; Rao, S.V.; Dwivedi, P.K. Hierarchical laser-patterned silver/graphene oxide hybrid SERS sensor for explosive detection. ACS Omega 2019, 4, 17691–17701. [Google Scholar] [CrossRef] [Green Version]
  51. He, C.; Qiu, S.; Sun, S.; Zhang, Q.; Lin, G.; Lei, S.; Han, X.; Yang, Y. Electrochemically active phosphotungstic acid assisted prevention of graphene restacking for high-capacitance supercapacitors. Energy Environ. Mater. 2018, 2, 88–95. [Google Scholar] [CrossRef] [Green Version]
  52. Bellardita, M.; Fiorenza, R.; D’Urso, L.; Spitaleri, L.; Gulino, A.; Compagnini, G.; Scirè, S.; Palmisano, L. Exploring the photothermo-catalytic performance of brookite TiO2-CeO2 composites. Catalysts 2020, 10, 765. [Google Scholar] [CrossRef]
  53. Bêche, E.; Charvin, P.; Perarnau, D.; Abanades, S.; Flamant, G. Ce 3d XPS investigation of cerium oxides and mixed cerium oxide (CexTiyOz). Surf. Interface Anal. 2008, 40, 264–267. [Google Scholar] [CrossRef]
  54. Zhu, Y.; Jain, N.; Hudait, M.K.; Maurya, D.; Varghese, R.; Priya, S. X-ray photoelectron spectroscopy analysis and band offset determination of CeO2 deposited on epitaxial (100), (110), and (111)Ge. J. Vac. Sci. Technol. B 2014, 32, 011217. [Google Scholar] [CrossRef]
  55. Deng, J.; Zhou, Y.; Li, S.; Xiong, L.; Wang, J.; Yuan, S.; Chen, Y. Designed synthesis and characterization of nanostructured ceria-zirconia based material with enhanced thermal stability and its application in three-way catalysis. J. Ind. Eng. Chem. 2018, 64, 219–229. [Google Scholar] [CrossRef]
  56. Feng, D.; Gao, T.-N.; Fan, M.; Li, A.; Li, K.; Wang, T.; Huo, Q.; Qiao, Z.-A. A general ligand-assisted self-assembly approach to crystalline mesoporous metal oxides. NPG Asia Mater. 2018, 10, 800–809. [Google Scholar] [CrossRef]
  57. Hota, I.; Soren, S.; Mohapatra, B.; Debnath, A.; Muthe, K.; Varadwaj, K.; Parhi, P. Mn-doped ceria/reduced graphene oxide nanocomposite as an efficient oxygen reduction reaction catalyst. J. Electroanal. Chem. 2019, 851, 113480. [Google Scholar] [CrossRef]
  58. Parwaiz, S.; Bhunia, K.; Das, A.K.; Khan, M.M.; Pradhan, D. Cobalt-doped ceria/reduced graphene oxide nanocomposite as an efficient oxygen reduction reaction catalyst and supercapacitor material. J. Phys. Chem. C 2017, 121, 20165–20176. [Google Scholar] [CrossRef]
  59. Ravichandran, S.; Bhuvanendran, N.; Xu, Q.; Maiyalagan, T.; Xing, L.; Su, H. Ordered mesoporous Pt-Ru-Ir nanostructures as superior bifunctional electrocatalyst for oxygen reduction/oxygen evolution reactions. J. Colloid Interface Sci. 2022, 608, 207–218. [Google Scholar] [CrossRef]
  60. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved synthesis of graphene oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  61. Lv, Y.; Wang, X.; Mei, T.; Li, J.; Wang, J. Single-step hydrothermal synthesis of N, S-dual-doped graphene networks as metal-free efficient electrocatalysts for oxygen reduction reaction. ChemistrySelect 2018, 3, 3241–3250. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of CeO2 and Co-doped CeO2; (b) CeO2/N, S–rGO and Co-doped CeO2/N, S–rGO; (c) GO and N, S–rGO.
Figure 1. (a) XRD patterns of CeO2 and Co-doped CeO2; (b) CeO2/N, S–rGO and Co-doped CeO2/N, S–rGO; (c) GO and N, S–rGO.
Catalysts 12 00006 g001
Figure 2. (a) Raman spectra of CeO2 and Co-doped CeO2; (b) Raman spectra of GO, N, S–rGO and 3% Co–CeO2/N, S–rGO.
Figure 2. (a) Raman spectra of CeO2 and Co-doped CeO2; (b) Raman spectra of GO, N, S–rGO and 3% Co–CeO2/N, S–rGO.
Catalysts 12 00006 g002
Figure 3. (a) SEM, (b) TEM and (c) HRTEM images of 3% Co–CeO2/N, S–rGO.
Figure 3. (a) SEM, (b) TEM and (c) HRTEM images of 3% Co–CeO2/N, S–rGO.
Catalysts 12 00006 g003
Figure 4. (a) Core level spectra of Ce 3d, (b) Co 2p, (c) O1s, (d) C 1s, (e) N 1s and (f) S 2p for 3% Co–CeO2/N, S–rGO.
Figure 4. (a) Core level spectra of Ce 3d, (b) Co 2p, (c) O1s, (d) C 1s, (e) N 1s and (f) S 2p for 3% Co–CeO2/N, S–rGO.
Catalysts 12 00006 g004
Figure 5. Cyclic voltammogram for (i) Co–CeO2/N, S–-rGO, (ii) 1.5% Co–CeO2/N, S–rGO, (iii) 3% Co–CeO2/N, S–rGO and (iv) 4.5% Co–CeO2/N, S–rGO.
Figure 5. Cyclic voltammogram for (i) Co–CeO2/N, S–-rGO, (ii) 1.5% Co–CeO2/N, S–rGO, (iii) 3% Co–CeO2/N, S–rGO and (iv) 4.5% Co–CeO2/N, S–rGO.
Catalysts 12 00006 g005
Figure 6. (a) LSV curves of CeO2 and Co-doped CeO2; (b) LSV curves of CeO2/N, S–rGO and Co-doped CeO2/N, S–rGO; (c) LSV curves of 3% Co–CeO2/N, S–rGO (at different rpm); (d) K–L plots of 3% CeO2/N, S–rGO.
Figure 6. (a) LSV curves of CeO2 and Co-doped CeO2; (b) LSV curves of CeO2/N, S–rGO and Co-doped CeO2/N, S–rGO; (c) LSV curves of 3% Co–CeO2/N, S–rGO (at different rpm); (d) K–L plots of 3% CeO2/N, S–rGO.
Catalysts 12 00006 g006
Figure 7. LSV curves of the (a) CeO2/N, S–rGO, (b) 1.5% Co–CeO2/N, S–rGO and (c) 4.5% Co–CeO2/N, S–rGO.
Figure 7. LSV curves of the (a) CeO2/N, S–rGO, (b) 1.5% Co–CeO2/N, S–rGO and (c) 4.5% Co–CeO2/N, S–rGO.
Catalysts 12 00006 g007
Figure 8. LSV curves of 3% Ce–CeO2/N, S–rGO before and after 5000 cycles.
Figure 8. LSV curves of 3% Ce–CeO2/N, S–rGO before and after 5000 cycles.
Catalysts 12 00006 g008
Table 1. Comparison of ORR activity of doped CeO2 with rGO based catalyst.
Table 1. Comparison of ORR activity of doped CeO2 with rGO based catalyst.
S.noCatalystOnset Potential (V vs Ag/AgCl)Halfwave Potential (V vs Ag/AgCl)Limiting Current Density (mA/cm2)References
1CeO2/N–rGO−0.15 V-4.0[32]
25% Mn–CeO2/rGO−0.136 V−0.336 V4.8[57]
35% Ni–CeO2/rGO−0.170 V−0.300 V3.6[31]
43% Co–CeO2/rGO−0.223 V−0.295 V-[58]
53% Co–CeO2/N, S–rGO−0.124 V−0.261 V4.2Present work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sridharan, M.; Maiyalagan, T.; Panomsuwan, G.; Techapiesancharoenkij, R. Enhanced Electrocatalytic Activity of Cobalt-Doped Ceria Embedded on Nitrogen, Sulfur-Doped Reduced Graphene Oxide as an Electrocatalyst for Oxygen Reduction Reaction. Catalysts 2022, 12, 6. https://doi.org/10.3390/catal12010006

AMA Style

Sridharan M, Maiyalagan T, Panomsuwan G, Techapiesancharoenkij R. Enhanced Electrocatalytic Activity of Cobalt-Doped Ceria Embedded on Nitrogen, Sulfur-Doped Reduced Graphene Oxide as an Electrocatalyst for Oxygen Reduction Reaction. Catalysts. 2022; 12(1):6. https://doi.org/10.3390/catal12010006

Chicago/Turabian Style

Sridharan, Manickam, Thandavarayan Maiyalagan, Gasidit Panomsuwan, and Ratchatee Techapiesancharoenkij. 2022. "Enhanced Electrocatalytic Activity of Cobalt-Doped Ceria Embedded on Nitrogen, Sulfur-Doped Reduced Graphene Oxide as an Electrocatalyst for Oxygen Reduction Reaction" Catalysts 12, no. 1: 6. https://doi.org/10.3390/catal12010006

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