Next Article in Journal
The Influence of Residual Sodium on the Catalytic Oxidation of Propane and Toluene over Co3O4 Catalysts
Previous Article in Journal
Comparison of Different Metal Doping Effects on Co3O4 Catalysts for the Total Oxidation of Toluene and Propane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Sb-Doped SnO2 Nanostructures on Electrocatalytic Performance of a Pt Catalyst for Methanol Oxidation Reaction

1
Department of Chemical Engineering, Soongsil University, Seoul 156-743, Korea
2
Boyaz Energy, 165 Gasandigital 2-ro, Geumcheon-gu, Seoul 08504, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(8), 866; https://doi.org/10.3390/catal10080866
Submission received: 10 July 2020 / Revised: 28 July 2020 / Accepted: 29 July 2020 / Published: 3 August 2020

Abstract

:
In this study, antimony-doped tin oxide (ATO) support materials for a Pt anode catalyst in direct methanol fuel cells were prepared and electrochemically evaluated. When the heating temperature was increased from 300 to 400 °C, the ATO samples exhibited a slightly decreased specific surface area and increased electrical conductivity. In particular, the ATO sample heated at 350 °C in an air atmosphere showed improved electrical conductivity (1.3 S cm−1) with an optimum specific surface area of ~34 m2 g−1. The supported Pt catalysts were synthesized using a polyol process with as-prepared and heated ATO samples and Vulcan XC-72R as supports (denoted as Pt/ATO, Pt/ATO-350, and Pt/C, respectively). In the methanol oxidation reaction (MOR), compared to Pt/C and Pt/ATO, Pt/ATO-350 exhibited the best electrocatalytic activity and stability for MOR, which could be attributed to Pt nanoparticles on the relatively stable oxide support with high electrical conductivity and interaction between the Pt catalyst and the heated ATO support.

Graphical Abstract

1. Introduction

Direct methanol fuel cells (DMFCs) have attracted attention due to the ease of storage and transport of methanol as a liquid fuel and its high energy density, as well as the simplicity of the system for various applications from portable devices to vehicle power sources [1,2,3,4]. However, poor methanol oxidation reaction (MOR) activity at the anode and the high cost of noble metal catalysts are major issues for the practical development of DMFCs [5,6,7,8,9,10,11]. The main challenge for DMFCs is to find a catalyst with excellent catalytic activity and stability. In particular, the cost and limited availability of Pt as a catalyst hamper the commercialization of DMFCs as power sources for fuel cell vehicles and portable devices [11,12]. Thus, to reduce the utilization of Pt catalysts, various research attempts have been made, including the fabrication of Pt-based alloys, nanostructure design of pure Pt or Pt-based alloys, and synthesis of supported catalysts [13].
Supported catalysts, in particular, have been utilized as an effective catalyst structure by depositing Pt nanoparticles on supporting materials [14,15]. The essential characteristics of these supporting materials are high specific surface area to support the nano-sized Pt nanoparticles, high electrical conductivity for efficient electron transport in electrochemical reactions, electrochemical stability, and porous structure to effectively transport product and by-products [12,16,17,18]. Carbon nanostructured materials, such as carbon black, carbon nanotubes, carbon nanofibers, graphene, and graphite nanosheets, have been extensively utilized as supports [9,16,19]. However, the oxidation of carbon-based materials as a degradation mechanism of the electrocatalyst (Equation (1)) needs to be prevented in order to improve electrode stability [17,20]. Although the electrochemical reaction at 25 °C is likely to be sluggish in that potential range, the oxidation rate of carbon at potentials above 0.9 V vs. SHE, i.e., Standard Hydrogen Electrode, can be rapidly accelerated even at 25 °C [21]. Thus, the electrocatalytic activity and stability of carbon-supported catalysts can deteriorate due to a collapse in their structure [22].
C + 2H2O → CO2 + 4H+ + 4e     E° = 0.207 V vs. SHE
To overcome the electrochemical oxidation of carbon materials in supported catalysts, transition metal oxides, such as WO3, TiO2, and SnO2, have been proposed as supports in proton-exchange membrane fuel cells (PEMFCs) [19,23]. Among these oxides, SnO2 is attractive because of its electrochemical stability, non-toxicity, and abundance [23,24]. However, a wide energy band gap of > 3.6 eV for SnO2 as a semiconductor can deteriorate its electrochemical performance in PEMFCs because of relatively poor electrical conduction during the electrochemical reactions [25,26,27]. In general, the introduction of dopants in a semiconductor is the best way to increase its electrical conductivity of a semiconductor [23,26,28]. Thus, antimony (Sb5+) utilized as a dopant in SnO2 (antimony-doped tin oxide, ATO) can produce extra electrons, thereby improving electrical conductivity in the SnO2 structure [23,26,29,30]. In this study, mesoporous ATO nanostructures as supports for a Pt anode catalyst were synthesized using the Adams fusion method and a heating process at different temperatures. The Pt anode catalysts for the MOR were deposited on the ATO supports using the polyol process. Furthermore, the electrocatalytic activity and stability of the anode catalysts for MOR were evaluated using typical electrochemical cells. Specifically, compared to literature, in this study, we found that the electrocatalytic performance of the catalysts for the MOR could be attributed to Pt nanoparticles on the relatively stable oxide support with electrical conductivity and interaction between the Pt catalyst and the ATO support heated at a varying temperature [31,32,33].

2. Results and Discussion

Figure 1 shows XRD patterns for the bare ATO and the samples heated at different temperatures in air. All the samples exhibited the main characteristic peaks corresponding to the (110), (101), (200), and (211) planes at 26°, 34°, 38°, and 52°, respectively, in a tetragonal SnO2 structure with a = b = 4.74 Å and c = 3.19 Å (P42/mn(136) group, Powder Diffraction File (PDF) 88-0287) [24,29]. This demonstrated that the heated samples maintained a crystal structure identical to that of the ATO before heating, without other phases. Compared to pure SnO2 and bare ATO, the diffraction peak shift towards higher angles demonstrated the substitution of Sn with Sb in the SnO2 structure without other Sb-related structures [34,35]. The average crystallite sizes of ATO, ATO-300, ATO-350, and ATO-400 were determined to be 8.6, 8.9, 8.9, and 8.7 nm, respectively, by the Scherer equation. Figure 2 shows the TEM images of the as-prepared and heated ATO samples. Overall, the samples appeared spherical in shape. The average crystallite sizes of ATO, ATO-300, ATO-350, and ATO-400 were 8.6, 8.8, 8.9, and 8.7 nm, respectively, which was in accordance with those obtained by the XRD results. Furthermore, as the heating temperature was increased from 300 to 400 °C, the size of the ATO samples slightly decreased due to a slight aggregation of the ATO particles. From the high-resolution TEM images, the ATO samples were indexed as tetragonal SnO2 with an interplanar spacing of 1.4 Å corresponding to the dominant (110) plane, which was in agreement with the XRD analysis.
To confirm the specific surface area and pore structure of the samples, N2 adsorption/desorption characteristic curves were obtained, as shown in Figure 3 and Table 1. All the samples exhibited type IV curves, indicating a mesoporous structure for the ATO samples synthesized by the Adams fusion method. The specific surface areas of ATO, ATO-300, ATO-350, and ATO-400 were 37, 36, 34, and 31 m2 g−1, respectively. Furthermore, the dominant pore sizes of ATO, ATO-300, ATO-350, and ATO-400 were 19.4, 20.0, 21.2, and 22.5 nm, respectively. As the heating temperature increased, the specific surface area decreased, and the pore size increased due to a slight aggregation of the ATO particles. To investigate the chemical state of Sb in the ATO-samples, XPS analysis was performed (Figure 4). The amount of Sb measured by XPS in the ATO samples was determined to be ~5%, which is close to an optimum value for Sb [36]. The Sb 3d spectrum consists of two characteristic peaks for 3d3/2 and 3d5/2, located at ~540 and ~531 eV, respectively, because of spin-orbit interactions [37,38,39]. Practically, it is hard to analyze the accurate valence state of antimony by fitting the Sb 3d5/2 peak because of overlap with the O 1s peak. Thus, the Sb 3d3/2 peaks in the samples were fitted into the two chemical states of Sb3+ and Sb5+ at ~539.74 and ~540.71 eV, respectively. As shown in Table 2, with increasing heating temperature, the ratio of Sb5+ in the entire Sb state increased from 85% to 95%. The content of Sb5+ as a dopant in the SnO2 structure could be significantly related to its electrical behavior, i.e., conductivity. In other words, Sb5+ might play the role of the donor to enhance electrical conductivity, producing extra electrons in the conduction band of the Sb-doped SnO2 structure as an n-type semiconductor [38,39]. The electrical conductivities in the Sb-doped SnO2 samples heated at different temperatures were compared (Figure 5). The average conductivity values of ATO, ATO-300, ATO-350, and ATO-400 were 0.32, 0.68, 1.30, and 0.86 S cm−1, respectively. Compared to the ATO, the heated samples exhibited improved electrical conduction due to more activation of Sb5+ as a dopant in the SnO2 structure. Furthermore, the conductivity of the ATO samples increased with increasing heating temperature. The improved electrical conductivity of the doped ATO samples could be attributed to an increased Sb5+ ratio, i.e., an increased concentration of the activated Sb donors (Table 2). However, since annealing in air at 400 °C might lead to a decreased Sb5+ ratio due to an air-rich atmosphere, resulting in decreased electrical conductivity of the ATO-400, thus, the ATO-350 heated at 350 °C in the air could have an optimum conductivity.
In general, the essential requirements for the supporting materials are high electrical conductivity, specific surface area, and stability. Specifically, as shown in Figure 5, among these ATO samples, ATO-350 exhibited a high electrical conductivity with the optimum specific surface area. Thus, to investigate the effect of the heated ATO support on the electrocatalytic performance of Pt catalyst, we prepared Pt nanoparticles as catalysts deposited on ATO and ATO-350 supports (Pt/ATO and Pt/ATO-350). Figure 6 shows XRD patterns for the Pt nanoparticles as catalysts deposited on ATO and ATO-350 as supports, denoted as Pt/ATO and Pt/ATO-350, respectively. All the samples exhibited the main characteristic XRD peaks corresponding to the (111), (200), and (220) planes at 39.7°, 46.2° and 67.4°, respectively, in a face-centered-cubic (fcc) structure in Pt. Furthermore, the ATO structure in the Pt/ATO was the same as that of the ATO, without any change during the synthesis of the Pt nanoparticles. As shown in Figure 7 (TEM images), the Pt nanoparticles formed by the polyol process were deposited on the supports. Compared to Pt/C, Pt/ATO and Pt/ATO-350 exhibited an aggregation of Pt nanoparticles on the ATO supports because of the relatively low specific surface area of the ATO samples. The average crystallite sizes of the Pt catalysts in the Pt/ATO, Pt/ATO-350, and Pt/C were 2.2, 2.5, and 2.7 nm, respectively. From the high-resolution TEM images, the Pt nanoparticles were found to have an interplanar spacing of 2.2 Å corresponding to the (111) plane of an fcc structure.
Cyclic voltammograms (CVs) of the catalysts were measured in Ar-saturated 0.5 M H2SO4 at a scan rate of 50 mV s−1 (Figure 8a). All the catalysts showed typical CVs of a polycrystalline Pt electrode containing electrochemical characteristic peaks corresponding to hydrogen oxidation/reduction and oxygen oxidation/reduction. In particular, the electrochemically active surface area (EASA) of the Pt catalysts could be calculated by the following equation [9]:
EASA   [ m 2 g P t 1 ] = Q H 0.21 × M P t
where QH is the charge amount of hydrogen adsorbed on the Pt catalyst, 0.21 mC cm−2 is the Coulombic charge (QH) associated with hydrogen adsorption of the polycrystalline Pt electrode, and MPt is an amount of the Pt loading on the working electrode. The EASAs of the Pt/C, Pt/ATO, and Pt/ATO-350 were determined to be ~9.5, ~4.3, and ~18.6 m2 gPt−1, respectively. To compare the electrocatalytic activity of the catalysts for the MOR, CVs were measured in Ar-saturated 0.5 M H2SO4 + 2 M CH3OH at a scan rate of 50 mV s−1 in the potential range 0.1–1.3 V (vs. SHE) (Figure 8b). The specific activities of Pt/C, Pt/ATO, and Pt/ATO-350 measured at 0.65 V, close to the onset potential for the MOR, were 0.26, 0.22, and 0.47 A cm−2, respectively (Table 3). Furthermore, the mass activities of Pt/C, Pt/ATO, and Pt/ATO-350 were 0.39, 0.27, and 0.58 A gPt−1, respectively. Overall, the electrocatalytic activity of Pt/ATO-350 was superior to that of Pt/ATO because of the improved electrical conductivity of ATO heated at 350 °C with an unchanged specific surface area of ATO. Compared to Pt/C, the electrochemical performance of Pt/ATO-350 could be attributed to the supported catalyst structure of the Pt nanoparticles deposited on the doped oxide support with high electrical conductivity and an optimum specific surface area [19]. Compared to oxide-supported Pt catalysts reported in the literature, Pt/ATO-350 exhibited improved electrocatalytic activity toward the MOR (Table 4) [7,24,40]. Furthermore, the electrocatalytic stability of the supported catalysts was evaluated by measuring the variation of current density at a constant potential of 0.7 V for 3600 s in Ar-saturated 0.5 M H2SO4 + 2 M CH3OH (Figure 9). The initial current values of Pt/C and Pt/ATO-350 were 0.91 and 4.37 A g−1Pt, respectively. After 3600 s, the MOR current densities of Pt/C and Pt/ATO-350 dropped to 0.12 and 0.73 A g−1Pt, respectively. Compared to Pt/C, Pt/ATO-350 exhibited superior electrocatalytic stability for the MOR, which could be attributed to Pt nanoparticles on the relatively stable oxide support with high conductivity and interaction between the Pt catalyst and the heated ATO support [24]. As shown in Figure 10, the average crystallite sizes of Pt/C and Pt/ATO-350 after the stability tests were ~4.4 and ~3.5 nm, respectively, increasing by 160% and 140%, respectively, from the values before the stability tests. However, compared to Pt/C, the relatively less increased catalyst size of Pt/ATO-350 indicated the superior stability of Pt/ATO-350 during the MOR. Figure 11 shows the Pt4f XPS spectra of Pt/C and Pt/ATO-350. Compared to Pt/C, the relatively lower binding energy shift of Pt4f in Pt/ATO-350 demonstrated a strong coupling effect between Pt and ATO with electron transfer between the ATO and the Pt [41,42,43]. Such an electronic effect between Pt and ATO could lead to the favorable adsorption of OH on Pt sites by facilitating the cleavage of intermediates during the MOR, resulting in an improved MOR activity. In addition, the strong metal-oxide support interaction could prevent aggregation of Pt nanoparticles during the MOR, resulting in an enhanced MOR stability.
To confirm the stability of the supports without Pt catalysts, CVs were compared before and after potential cycling between 0.1 and 1.3 V for 100 cycles in Ar-saturated 0.5 M H2SO4 (Figure 12). The variation of the area in the CVs before and after potential cycling demonstrates the degradation of the stability of the supporting material [18]. The carbon support in Pt/C showed an increased area in the CV due to surface oxidation of the carbon, whereas the CV area for ATO-350 was maintained before and after cycling, thereby indicating high stability of the Sb in ATO [36,44,45]. Thus, compared to carbon support, Pt/ATO-350 showed superior electrochemical stability, which could be attributed to the excellent electrochemical stability of ATO-350 between 0.1 and 1.3 V in an acid solution. The superior electrocatalytic activity and stability of Pt/ATO-350 could be attributed to the relatively stable conductive oxide support and interaction between the Pt catalyst and the ATO support.

3. Materials and Methods

3.1. Synthesis of ATO Supported Pt Catalysts

The ATO sample was synthesized by the Adams fusion method [27,46]. To prepare ATO, the desired amount of SnCl4·5H2O (Sigma-Aldrich, 98%) was dissolved in 100 mL of de-ionized (DI) water with stirring for 30 min, and 5% SbCl3 was then added, calculated based on tin chloride, and continually stirred for 24 h. Subsequently, excessive NaNO3 (Sigma-Aldrich, 99%) was added to the solution under sonication with stirring for 30 min. The solvent in the solution was evaporated at 80 °C for 4 h. The remaining powder was calcined under an air atmosphere at different temperatures from 300 to 400 °C for 5 h. The calcined powders were washed and filtered with a 0.5 M HCl solution and then dried at 50 °C for 12 h. To prepare ATO supported Pt (20 wt%) catalysts, Pt nanoparticles were formed using the polyol process and deposited on the ATO samples. Ethylene glycol (Sigma-Aldrich, 99.8%) was used as both the solvent and reductant in the polyol process [14]. An amount of 0.225 g of H2PtCl6 (Sigma-Aldrich, 99.9%) was dissolved in 95 mL of ethylene glycol (Samchun, 99.5%) with continuous stirring. NaOH (Sigma-Aldrich, 97%) was added until a pH of 11.5 was reached in the precursor solution, and stirring was continued for 30 min. Then, the mixed solution was heated to 180 °C, and the temperature was maintained at 180 °C for 1 h with stirring for complete reduction of the Pt precursor. The support materials (0.427 g) were added to the colloidal solution of Pt nanoparticles at 25 °C with sonication for 30 min and overnight stirring, and the ξ-potential of the ATO samples was adjusted to pH 2.0 using 1 M H2SO4. The supported catalysts (Pt/ATO) were washed with ethanol and DI water and dried in a 50 °C oven for 12 h. For comparison, Pt (20 wt%) catalyst deposited on Vulcan XC-72R as support (Pt/C) was prepared using the same procedure as the Pt/ATO.

3.2. Structural Analysis of ATO Supported Pt Catalysts

The crystal structure of the samples was analyzed by an X-ray diffractometer (XRD, Bruker D2 Phaser system, Geumcheon, Korea) with Cu Ka radiation (λ = 0.15418 nm) and a Ni filter operating at an accelerating voltage of 30 kV and a current of 10 mA. The specific surface area and pore structure of the samples preheated under a pressure of 100 mmHg at 473 K for 360 min were characterized by nitrogen adsorption/desorption analyzer (Micromeritics ASAP 2020 instrument, Norcross, GA, USA). The morphology of the samples was observed using a Cs-corrected spherical aberration-corrected scanning transmission electron microscope (JEM-ARM 200F microscope, Akishima, Tokyo, Japan) working at an accelerating voltage of 200 kV. To characterize the surface composition and chemical species of the samples, X-ray photoelectron spectroscopy (XPS, K-Alpha, Waltham, MA, USA) was carried out with an Al Kα X-ray of 1468.8 eV and beam power of 200 W under a chamber pressure of 7.8 × 10−9 Torr. The sample was loaded in a Teflon-coated stainless steel container, and its electrical conductivity was measured under a pressure of ~5 bar as the following equation:
σ = L A × I I 0 V V 0   ( S · c m 1 )
where σ is electrical conductivity, L is the length of the pressed powder sample, A is the area of the pressed powder sample, V0 is the initial voltage, V is the final voltage, I0 is the initial current, and I is the final current.

3.3. Electrochemical Analysis of ATO Supported Pt Catalysts

The electrochemical properties of the samples were characterized in a three-electrode cell at 25 °C. Pt wire and Ag/AgCl (in 3 M KCl) were used as counter and reference electrodes, respectively. The working electrode was prepared by loading the prepared sample on the glassy carbon (GC) electrode of a rotating disk electrode system. The ink was prepared by mixing the sample powder with DI water, isopropyl alcohol (Sigma-Aldrich, 99.7%), and Nafion (Sigma-Aldrich, 5 wt% Nafion® ionomer) and then dropped on the GC electrode. After drying in an oven at 50 °C for 10 min, the loading amount of the catalyst deposited on the GC electrode was 80 μgPt cm−2. The electrochemical properties of the catalysts were characterized in Ar-saturated 0.1 M H2SO4 using cyclic voltammetry (CV, Metrohm Autolab PGSTAT302N instrument, Netherlands) in the potential range 0–1.2 V (vs. Ag/AgCl) with a scan rate of 50 mV s−1. To evaluate the electrocatalytic stability of the MOR, the electrocatalysts were kept at 0.7 V (vs. Ag/AgCl) for 3600 s in Ar-saturated 0.1 M H2SO4 + 2.0 M CH3OH. All electrode potentials were converted to the SHE potentials using the Nernst equation.

4. Conclusions

In summary, we synthesized Sb-doped SnO2 nanostructure supports for Pt catalyst for the MOR in an acid solution. The ATO-350 heated at 350 °C in air exhibited the best electrical conductivity (1.30 S cm−1) caused by a high Sb5+ ratio (95%) and the optimum specific surface area (34 m2 gPt−1) with a mesoporous structure. Compared to Pt/C and Pt/ATO, Pt/ATO-350 showed the best electrochemical performance, i.e., electrocatalytic activity (0.58 A gPt−1 at 0.65 V) and stability, for the MOR, resulting from the relatively stable conductive oxide support and interaction between the Pt catalyst and ATO support.

Author Contributions

Conceptualization, methodology, formal analysis, writing—original draft preparation, S.-G.L.; data curation, S.-B.H.; formal analysis, W.-J.L.; writing—review and editing, supervision, project administration, K.-W.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Technology Development Program to Solve Climate Changes of the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT (NRF-2017M1A2A2086648).

Acknowledgments

This research was supported by the National Research Foundation of Korea.

Conflicts of Interest

There is no conflict of interest.

References

  1. Feng, Y.Y.; Kong, W.Q.; Yin, Q.Y.; Du, L.X.; Zheng, Y.T.; Kong, D.S. Platinum catalysts promoted by in doped SnO2 support for methanol electrooxidation in alkaline electrolyte. J. Power Sources 2014, 252, 156–163. [Google Scholar] [CrossRef]
  2. Amani, M.; Kazemeini, M.; Hamedanian, M.; Pahlavanzadeh, H.; Gharibi, H. Investigation of methanol oxidation on a highly active and stable Pt-Sn electrocatalyst supported on carbon-polyaniline composite for application in a passive direct methanol fuel cell. Mater. Res. Bull. 2015, 68, 166–178. [Google Scholar] [CrossRef]
  3. Kwak, D.H.; Lee, Y.W.; Lee, K.H.; Park, A.R.; Moon, J.S.; Park, K.W. One-step synthesis of hexapod pt nanoparticles deposited on carbon black for improved methanol electrooxidation. Int. J. Electrochem. Sci. 2013, 8, 5102–5107. [Google Scholar]
  4. You, D.J.; Kwon, K.; Pak, C.; Chang, H. Platinum-antimony tin oxide nanoparticle as cathode catalyst for direct methanol fuel cell. Catal. Today 2009, 146, 15–19. [Google Scholar] [CrossRef]
  5. Cao, Y.; Yang, Y.; Shan, Y.; Huang, Z. One-Pot and Facile Fabrication of Hierarchical Branched Pt-Cu Nanoparticles as Excellent Electrocatalysts for Direct Methanol Fuel Cells. ACS Appl. Mater. Interfaces 2016, 8, 5998–6003. [Google Scholar] [CrossRef]
  6. Chang, J.; Feng, L.; Jiang, K.; Xue, H.; Cai, W.B.; Liu, C.; Xing, W. Pt-CoP/C as an alternative PtRu/C catalyst for direct methanol fuel cells. J. Mater. Chem. A 2016, 4, 18607–18613. [Google Scholar] [CrossRef]
  7. Zheng, Y.; Qiao, J.; Hu, J.; Song, F.; Huo, D.; Yuan, J.; Shen, J.; Niu, L.; Wang, A.-j. PtIr alloy nanowire assembly on carbon cloth as advanced anode catalysts for methanol oxidation. Int. J. Hydrogen Energy 2019, 44, 20336–20344. [Google Scholar] [CrossRef]
  8. Luo, F.; Zhang, Q.; Qu, K.; Guo, L.; Hu, H.; Yang, Z.; Cai, W.; Cheng, H. Decorated PtRu Electrocatalyst for Concentrated Direct Methanol Fuel Cells. ChemCatChem 2019, 11, 1238–1243. [Google Scholar] [CrossRef]
  9. Liu, J.; Jiao, M.; Lu, L.; Barkholtz, H.M.; Li, Y.; Jiang, L.; Wu, Z.; Liu, D.J.; Zhuang, L.; Ma, C.; et al. High performance platinum single atom electrocatalyst for oxygen reduction reaction. Nat. Commun. 2017, 8, 1–9. [Google Scholar] [CrossRef] [Green Version]
  10. Long, Z.; Gong, L.; Sun, Y.; Li, Y.; Xu, P.; Zhang, X.; Ge, J.; Liu, C.; Ma, S.; Jin, Z. In-situ precise electrocatalytic behaviors of Pt/C and PtRu/C for methanol oxidation of DMFCs via the designed micro-MEA. Int. J. Hydrogen Energy 2018, 43, 12413–12419. [Google Scholar] [CrossRef]
  11. Kang, Y.S.; Jo, S.; Choi, D.; Kim, J.Y.; Park, T.; Yoo, S.J. Pt-Sputtered Ti Mesh Electrode for Polymer Electrolyte Membrane Fuel Cells. Int. J. Precis. Eng. Manuf. Green Technol. 2019, 6, 271–279. [Google Scholar] [CrossRef]
  12. Sharma, S.; Pollet, B.G. Support materials for PEMFC and DMFC electrocatalysts—A review. J. Power Sources 2012, 208, 96–119. [Google Scholar] [CrossRef]
  13. Gasteiger, H.A.; Kocha, S.S.; Sompalli, B.; Wagner, F.T. Activity benchmarks and requirements for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Appl. Catal. B Environ. 2005, 56, 9–35. [Google Scholar] [CrossRef]
  14. Qi, J.; Jiang, L.; Jing, M.; Tang, Q.; Sun, G. Preparation of Pt/C via a polyol process - Investigation on carbon support adding sequence. Int. J. Hydrogen Energy 2011, 36, 10490–10501. [Google Scholar] [CrossRef]
  15. Elezovic, N.R.; Radmilovic, V.R.; Krstajic, N.V. Platinum nanocatalysts on metal oxide based supports for low temperature fuel cell applications. RSC Adv. 2016, 6, 6788–6801. [Google Scholar] [CrossRef]
  16. Wang, J.; Yin, G.; Shao, Y.; Zhang, S.; Wang, Z.; Gao, Y. Effect of carbon black support corrosion on the durability of Pt/C catalyst. J. Power Sources 2007, 171, 331–339. [Google Scholar] [CrossRef]
  17. Natarajan, S.K.; Hamelin, J. Electrochemical Durability of Carbon Nanostructures as Catalyst Support for PEMFCs. J. Electrochem. Soc. 2009, 156, B210. [Google Scholar] [CrossRef]
  18. Lee, J.M.; Kim, S.B.; Lee, Y.W.; Kim, D.Y.; Han, S.B.; Roh, B.; Hwang, I.; Park, K.W. Core-shell nanostructure electrodes for improved electrocatalytic properties in methanol electrooxidation. Appl. Catal. B Environ. 2012, 111–112, 200–207. [Google Scholar] [CrossRef]
  19. Mohanta, P.K.; Glökler, C.; Arenas, A.O.; Jörissen, L. Sb doped SnO2 as a stable cathode catalyst support for low temperature polymer electrolyte membrane fuel cell. Int. J. Hydrogen Energy 2017, 42, 27950–27961. [Google Scholar] [CrossRef]
  20. Oh, H.S.; Oh, J.G.; Haam, S.; Arunabha, K.; Roh, B.; Hwang, I.; Kim, H. On-line mass spectrometry study of carbon corrosion in polymer electrolyte membrane fuel cells. Electrochem. Commun. 2008, 10, 1048–1051. [Google Scholar] [CrossRef]
  21. Zhang, X.; Yang, Y.; Guo, L.; Liu, H. Effects of carbon corrosion on mass transfer losses in proton exchange membrane fuel cells. Int. J. Hydrogen Energy 2017, 42, 4699–4705. [Google Scholar] [CrossRef] [Green Version]
  22. Maass, S.; Finsterwalder, F.; Frank, G.; Hartmann, R.; Merten, C. Carbon support oxidation in PEM fuel cell cathodes. J. Power Sources 2008, 176, 444–451. [Google Scholar] [CrossRef]
  23. Qu, W.; Wang, Z.; Sui, X.; Gu, D. An efficient antimony doped tin oxide and carbon nanotubes hybrid support of Pd catalyst for formic acid electrooxidation. Int. J. Hydrogen Energy 2014, 39, 5678–5688. [Google Scholar] [CrossRef]
  24. Lee, K.S.; Park, I.S.; Cho, Y.H.; Jung, D.S.; Jung, N.; Park, H.Y.; Sung, Y.E. Electrocatalytic activity and stability of Pt supported on Sb-doped SnO2 nanoparticles for direct alcohol fuel cells. J. Catal. 2008, 258, 143–152. [Google Scholar] [CrossRef]
  25. Fang, L.; Luo, X.; Saipanya, S.; Huang, X.; Li, F. Preparation of Pt/ATO catalysts for electrocatalysis in direct alcohol fuel cell. Chiang Mai J. Sci. 2016, 43, 169–175. [Google Scholar]
  26. Guo, D.J. Electrooxidation of ethanol on novel multi-walled carbon nanotube supported platinum-antimony tin oxide nanoparticle catalysts. J. Power Sources 2011, 196, 679–682. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Wang, C.; Mao, Z.; Wang, N. Preparation of nanometer-sized SnO2 by the fusion method. Mater. Lett. 2007, 61, 1205–1209. [Google Scholar] [CrossRef]
  28. Zhang, J.; Gao, L. Synthesis and characterization of antimony-doped tin oxide (ATO) nanoparticles. Inorg. Chem. Commun. 2004, 7, 91–93. [Google Scholar] [CrossRef]
  29. Yin, M.; Xu, J.; Li, Q.; Jensen, J.O.; Huang, Y.; Cleemann, L.N.; Bjerrum, N.J.; Xing, W. Highly active and stable Pt electrocatalysts promoted by antimony-doped SnO2 supports for oxygen reduction reactions. Appl. Catal. B Environ. 2014, 144, 112–120. [Google Scholar] [CrossRef]
  30. Cavaliere, S.; Jiménez-Morales, I.; Ercolano, G.; Savych, I.; Jones, D.; Rozière, J. Highly Stable PEMFC Electrodes Based on Electrospun Antimony-Doped SnO2. ChemElectroChem 2015, 2, 1966–1973. [Google Scholar] [CrossRef] [Green Version]
  31. Xu, J.; Aili, D.; Li, Q.; Pan, C.; Christensen, E.; Jensen, J.O.; Zhang, W.; Liu, G.; Wang, X.; Bjerrum, N.J. Antimony doped tin oxide modified carbon nanotubes as catalyst supports for methanol oxidation and oxygen reduction reactions. J. Mater. Chem. A 2013, 1, 9737–9745. [Google Scholar] [CrossRef]
  32. Tiwari, J.N.; Tiwari, R.N.; Singh, G.; Kim, K.S. Recent progress in the development of anode and cathode catalysts for direct methanol fuel cells. Nano Energy 2013, 2, 553–578. [Google Scholar] [CrossRef]
  33. Pan, C.; Li, Y.; Ma, Y.; Zhao, X.; Zhang, Q. Platinum—Antimony doped tin oxide nanoparticles supported on carbon black as anode catalysts for direct methanol fuel cells. J. Power Sources 2011, 196, 6228–6231. [Google Scholar] [CrossRef]
  34. López Morales, F.; Zayas, T.; Contreras, O.E.; Salgado, L. Effect of Sn precursor on the synthesis of SnO2 and Sb-doped SnO2 particles via polymeric precursor method. Front. Mater. Sci. 2013, 7, 387–395. [Google Scholar] [CrossRef]
  35. Zhong, X.; Yang, B.; Zhang, X.; Jia, J.; Yi, G. Effect of calcining temperature and time on the characteristics of Sb-doped SnO 2 nanoparticles synthesized by the sol-gel method. Particuology 2012, 10, 365–370. [Google Scholar] [CrossRef]
  36. Jiménez-Morales, I.; Cavaliere, S.; Dupont, M.; Jones, D.; Rozière, J. On the stability of antimony doped tin oxide supports in proton exchange membrane fuel cell and water electrolysers. Sustain. Energy Fuels 2019, 3, 1526–1535. [Google Scholar] [CrossRef]
  37. Sladkevich, S.; Mikhaylov, A.A.; Prikhodchenko, P.V.; Tripol’skaya, T.A.; Lev, O. Antimony tin oxide (ATO) nanoparticle formation from H2O2 solutions: A new generic film coating from basic solutions. Inorg. Chem. 2010, 49, 9110–9112. [Google Scholar] [CrossRef]
  38. Mao, W.; Xiong, B.; Li, Q.; Zhou, Y.; Yin, C.; Liu, Y.; He, C. Influences of defects and Sb valence states on the temperature dependent conductivity of Sb doped SnO2 thin films. Phys. Lett. Sect. A Gen. At. Solid State Phys. 2015, 379, 1946–1950. [Google Scholar] [CrossRef]
  39. Liu, S.M.; Ding, W.Y.; Chai, W.P. Influence of Sb doping on crystal structure and electrical property of SnO2 nanoparticles prepared by chemical coprecipitation. Phys. B Condens. Matter 2011, 406, 2303–2307. [Google Scholar] [CrossRef]
  40. Zhou, C.; Wang, H.; Liang, J.; Peng, F.; Yu, H.; Yang, J. Effects of RuO2 content in Pt/RuO2/CNTs nanocatalyst on the electrocatalytic oxidation performance of methanol. Cuihua Xuebao / Chinese J. Catal. 2008, 29, 1093–1098. [Google Scholar] [CrossRef]
  41. Matin, M.A.; Lee, E.; Kim, H.; Yoon, W.S.; Kwon, Y.U. Rational syntheses of core-shell Fe@(PtRu) nanoparticle electrocatalysts for the methanol oxidation reaction with complete suppression of CO-poisoning and highly enhanced activity. J. Mater. Chem. A 2015, 3, 17154–17164. [Google Scholar] [CrossRef]
  42. Timperman, L.; Alonso-Vante, N. Oxide Substrate Effect Toward Electrocatalytic Enhancement of Platinum and Ruthenium-Selenium Catalysts. Electrocatalysis 2011, 2, 181–191. [Google Scholar] [CrossRef]
  43. Jiménez-Morales, I.; Cavaliere, S.; Jones, D.; Rozière, J. Strong metal-support interaction improves activity and stability of Pt electrocatalysts on doped metal oxides. Phys. Chem. Chem. Phys. 2018, 20, 8765–8772. [Google Scholar] [CrossRef] [PubMed]
  44. Geiger, S.; Kasian, O.; Mingers, A.M.; Mayrhofer, K.J.J.; Cherevko, S. Stability limits of tin-based electrocatalyst supports. Sci. Rep. 2017, 7, 3–9. [Google Scholar] [CrossRef]
  45. Dubau, L.; Maillard, F.; Chatenet, M.; Cavaliere, S.; Jiménez-Morales, I.; Mosdale, A.; Mosdale, R. Durability of alternative metal oxide supports for application at a proton-exchange membrane fuel cell cathode—Comparison of antimony- And niobium-doped tin oxide. Energies 2020, 13, 403. [Google Scholar] [CrossRef] [Green Version]
  46. Han, S.B.; Mo, Y.H.; Lee, Y.S.; Lee, S.G.; Park, D.H.; Park, K.W. Mesoporous iridium oxide/Sb-doped SnO2 nanostructured electrodes for polymer electrolyte membrane water electrolysis. Int. J. Hydrogen Energy 2019, 45, 1409–1416. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the antimony-doped tin oxide (ATO) samples heated at different temperatures from 300 to 400 °C in an air atmosphere. (A tetragonal SnO2 structure with a = b = 4.74 Å, c = 3.19 Å (P42/mn(136) group, PDF 88-0287)).
Figure 1. XRD patterns of the antimony-doped tin oxide (ATO) samples heated at different temperatures from 300 to 400 °C in an air atmosphere. (A tetragonal SnO2 structure with a = b = 4.74 Å, c = 3.19 Å (P42/mn(136) group, PDF 88-0287)).
Catalysts 10 00866 g001
Figure 2. TEM images, size distributions, and high-resolution TEM images of ATO (a,e,i), ATO-300 (b,f,j), ATO-350 (c,g,k), and ATO-400 (d,h,l).
Figure 2. TEM images, size distributions, and high-resolution TEM images of ATO (a,e,i), ATO-300 (b,f,j), ATO-350 (c,g,k), and ATO-400 (d,h,l).
Catalysts 10 00866 g002
Figure 3. N2 adsorption-desorption isotherms and pore size distributions for ATO (a,e), ATO-300 (b,f), ATO-350 (c,g), and ATO-400 (d,h).
Figure 3. N2 adsorption-desorption isotherms and pore size distributions for ATO (a,e), ATO-300 (b,f), ATO-350 (c,g), and ATO-400 (d,h).
Catalysts 10 00866 g003
Figure 4. XPS Sb3d3/2 spectra of (a) ATO, (b) ATO-300, (c) ATO-350, and (d) ATO-400.
Figure 4. XPS Sb3d3/2 spectra of (a) ATO, (b) ATO-300, (c) ATO-350, and (d) ATO-400.
Catalysts 10 00866 g004
Figure 5. Comparison of the specific surface area and electrical conductivity of the samples.
Figure 5. Comparison of the specific surface area and electrical conductivity of the samples.
Catalysts 10 00866 g005
Figure 6. XRD patterns of Pt/C, Pt/ATO, and Pt/ATO-350.
Figure 6. XRD patterns of Pt/C, Pt/ATO, and Pt/ATO-350.
Catalysts 10 00866 g006
Figure 7. TEM images, size distributions, and high-resolution TEM images of Pt/C (a,d,g), Pt/ATO (b,e,h), and Pt/ATO-350 (c,f,i).
Figure 7. TEM images, size distributions, and high-resolution TEM images of Pt/C (a,d,g), Pt/ATO (b,e,h), and Pt/ATO-350 (c,f,i).
Catalysts 10 00866 g007
Figure 8. Cyclic voltammograms (CVs) for Pt/C, Pt/ATO, and Pt/ATO-350 measured with a scan rate of 50 mV s−1 (a) in Ar-saturated 0.5 M H2SO4 and (b) Ar-saturated 0.5 M H2SO4 + 2 M CH3OH at 25 °C. (c) Comparison of specific and mass activities of the supported catalysts measured at 0.65 V.
Figure 8. Cyclic voltammograms (CVs) for Pt/C, Pt/ATO, and Pt/ATO-350 measured with a scan rate of 50 mV s−1 (a) in Ar-saturated 0.5 M H2SO4 and (b) Ar-saturated 0.5 M H2SO4 + 2 M CH3OH at 25 °C. (c) Comparison of specific and mass activities of the supported catalysts measured at 0.65 V.
Catalysts 10 00866 g008
Figure 9. Plots of current density vs. time for Pt/C and Pt/ATO.
Figure 9. Plots of current density vs. time for Pt/C and Pt/ATO.
Catalysts 10 00866 g009
Figure 10. TEM images and size distributions of Pt/C (a,c) and Pt/ATO-350 (b,d) after the methanol oxidation reaction (MOR) stability test.
Figure 10. TEM images and size distributions of Pt/C (a,c) and Pt/ATO-350 (b,d) after the methanol oxidation reaction (MOR) stability test.
Catalysts 10 00866 g010
Figure 11. XPS Pt4f spectra of (a) Pt/ATO-350 and (b) Pt/C.
Figure 11. XPS Pt4f spectra of (a) Pt/ATO-350 and (b) Pt/C.
Catalysts 10 00866 g011
Figure 12. CVs for (a) carbon and (b) ATO-350 after 100 cycles between 0.1 and 1.3 V with a scan rate of 50 mV s−1 in Ar-saturated 0.5 M H2SO4 at 25 °C.
Figure 12. CVs for (a) carbon and (b) ATO-350 after 100 cycles between 0.1 and 1.3 V with a scan rate of 50 mV s−1 in Ar-saturated 0.5 M H2SO4 at 25 °C.
Catalysts 10 00866 g012
Table 1. Comparison of the specific surface area, pore size, and particle size of the antimony-doped tin oxide (ATO) samples.
Table 1. Comparison of the specific surface area, pore size, and particle size of the antimony-doped tin oxide (ATO) samples.
SampleSpecific Surface Area (m2 g−1)Pore Size (nm)Crystalline Size (nm)
ATO3719.48.6
ATO-3003620.08.8
ATO-3503421.28.9
ATO-4003122.58.7
Table 2. Comparison of the Sb5+ ratio and electrical conductivity of the ATO samples.
Table 2. Comparison of the Sb5+ ratio and electrical conductivity of the ATO samples.
Sample S b 5 + S b 3 + + S b 5 + × 100 Conductivity (S cm−1)
ATO48%0.32
ATO-30049%0.68
ATO-35066%1.30
ATO-40064%0.86
Table 3. Comparison of the methanol oxidation reaction (MOR) activities and electrochemically active surface areas (EASAs) of the supported Pt catalysts.
Table 3. Comparison of the methanol oxidation reaction (MOR) activities and electrochemically active surface areas (EASAs) of the supported Pt catalysts.
SampleCurrent Density
(A cmGEO−2)
Current Density
(A cmEASA−2)
Current Density
(A gPt−1)
EASA
(m2 gPt−1)
Pt/C0.260.090.39 9.5
Pt/ATO0.220.900.274.3
Pt/ATO-3500.470.520.5818.6
Table 4. Comparison of the Pt catalysts supported by oxide supports.
Table 4. Comparison of the Pt catalysts supported by oxide supports.
SampleConditionMass Activity
(A gPt−1@0.65 V)
Ref.
Pt/ATO-3500.5 M H2SO4 + 2 M CH3OH
50 mV s−1
0.58In this study
Pt/SnO20.5 M KOH + 2 M CH3OH
20 mV s−1
0.307
Pt/InSnO20.5 M KOH + 2 M CH3OH
20 mV s−1
0.417
Pt/SbSnO20.5 M H2SO4 + 2 M CH3OH
50 mV s−1
0.6424
Pt/RuO2/CNT1 M HClO4 + 1 M CH3OH
100 mVs−1
0.4935

Share and Cite

MDPI and ACS Style

Lee, S.-G.; Han, S.-B.; Lee, W.-J.; Park, K.-W. Effect of Sb-Doped SnO2 Nanostructures on Electrocatalytic Performance of a Pt Catalyst for Methanol Oxidation Reaction. Catalysts 2020, 10, 866. https://doi.org/10.3390/catal10080866

AMA Style

Lee S-G, Han S-B, Lee W-J, Park K-W. Effect of Sb-Doped SnO2 Nanostructures on Electrocatalytic Performance of a Pt Catalyst for Methanol Oxidation Reaction. Catalysts. 2020; 10(8):866. https://doi.org/10.3390/catal10080866

Chicago/Turabian Style

Lee, Seul-Gi, Sang-Beom Han, Woo-Jun Lee, and Kyung-Won Park. 2020. "Effect of Sb-Doped SnO2 Nanostructures on Electrocatalytic Performance of a Pt Catalyst for Methanol Oxidation Reaction" Catalysts 10, no. 8: 866. https://doi.org/10.3390/catal10080866

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