Next Article in Journal
Low-Temperature Esterification to Produce Oleic Acid-Based Wax Esters Catalyzed by 4-Dodecylbenzenesulfonic Acid
Previous Article in Journal
CO2 Reforming of CH4 Using Coke Oven Gas over Ni/MgO-Al2O3 Catalysts: Effect of the MgO:Al2O3 Ratio
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon-Supported Pt-SnO2 Catalysts for Oxygen Reduction Reaction over a Wide Temperature Range: Rotating Disk Electrode Study

by
Ruslan M. Mensharapov
1,
Nataliya A. Ivanova
1,
Dmitry D. Spasov
1,2,
Elena V. Kukueva
1,
Adelina A. Zasypkina
1,
Ekaterina A. Seregina
1,
Sergey A. Grigoriev
1,2,3,* and
Vladimir N. Fateev
1
1
National Research Center “Kurchatov Institute”, 1, Akademika Kurchatova sq., 123182 Moscow, Russia
2
National Research University “Moscow Power Engineering Institute”, 14, Krasnokazarmennaya St., 111250 Moscow, Russia
3
HySA Infrastructure Center of Competence, Faculty of Engineering, North-West University, Potchefstroom 2531, South Africa
*
Author to whom correspondence should be addressed.
Submission received: 30 October 2021 / Revised: 22 November 2021 / Accepted: 28 November 2021 / Published: 30 November 2021

Abstract

:
Pt/C and Pt/x-SnO2/C catalysts (where x is mass content of SnO2) were synthesized using a polyol method. Their kinetic properties towards oxygen reduction reaction were studied by a rotating disk electrode (RDE) technique in a temperature range from 1 to 50 °C. The SnO2 content of catalyst samples was 5 and 10 wt.%. A quick evaluation of the catalyst activity, electrochemical behavior and average number of transferred electrons were performed using the RDE technique. It has been shown that the use of x-SnO2 (through modification of the carbon support) in a binary system together with Pt does not reduce the catalyst activity in the temperature range of 1–30 °C. The temperature rising up to 50 °C resulted in composite catalyst activity reduction at about 30%.

Graphical Abstract

1. Introduction

Increasing interest in renewable energy sources and hydrogen energy makes the research and development of polymer electrolyte membrane fuel cells (PEMFCs) an important task. A key element of PEMFCs is a membrane electrode assembly (MEA), which includes the anode and cathode catalytic layers. Since the kinetics of the hydrogen oxidation reaction on the anode catalytic layer is rather rapid, the rate-limiting process in the PEMFC is the oxygen reduction reaction (ORR) at the cathode (Figure 1), which is a slow four-electron transfer reaction. The low ORR rate calls for high-effective cathode catalysts showing high activity in this process.
Pt-based carbon-supported catalysts are commonly used at the PEMFC cathode [1,2]. Such electrocatalysts have high activity and sufficient stability under standard operating conditions of PEMFCs. However, widening of the range of external operating conditions (temperature and humidity) involves a number of challenges. First, drying of the ion-exchange polymer is observed at high temperatures and low humidity, which leads to a deterioration in proton conductivity both in the catalytic layer ionomer and in the membrane [3,4,5,6]. Second, at negative temperatures, ice crystals are formed inside MEA components, and destruction of the catalytic layers and membrane occurs during freeze-thaw cycles [7,8,9,10,11,12,13]. These problems can be solved by applying additives with high water retention ability, such as tin dioxide [14,15,16,17,18].
However, the addition of tin dioxide particles may have a negative effect on the activity of the platinum catalysts due to a possible blockage of Pt active sites or formation of large agglomerates of particles [19,20]. Therefore, a detailed study of activity of composite tin dioxide catalysts in the ORR is important for evaluation of their applicability in PEMFCs over a wide range of operating conditions.
The technique of a rotating disk electrode (RDE) in a three-electrode electrochemical cell is commonly used for the determination of electrocatalyst activity towards ORR. This technique allows for estimating kinetic and diffusion components of the reaction current. An extension of this method to low and high temperatures may aid in evaluating the efficiency of the catalysts over a wide temperature range and determining the activation energy in the ORR.
Most studies [21,22,23,24] are dedicated to the evaluation of the activity of tin dioxide composite catalysts in the alcohol oxidation reaction since Pt-SnO2 hetero-clusters enhance oxidation of COads intermediates. Fewer studies [25,26,27] have focused on investigation of the activity in the ORR of hybrid Pt-SnO2 catalyst with a high tin dioxide loading, co-catalyst properties of SnO2 have been reported.
As reported in our previous work [20], the composite electrocatalysts with the tin dioxide content of 5 and 10 wt.% demonstrate a significantly improved stability in the course of the accelerated stress testing and good efficiency within the PEMFC. This effect is achieved due to formation of Pt-SnO2 hetero-clusters, providing an improvement in the electrocatalyst durability. However, the presence of tin dioxide led to a reduction in the electrochemically active surface area (EASA), and may also affect the ORR kinetics.
This paper is dedicated to further investigation of composite Pt-SnO2 catalysts. The kinetics of different electrocatalysts in the ORR was investigated using the Koutecky–Levich approach. The study of the kinetics in the temperature range from 1 to 20 °C allows for evaluating the catalyst efficiency in conditions of starting the PEMFC at low ambient temperatures (“cold start”). The upper limit of the measurement temperature of 50 °C is explained by the increased evolution of acid vapors and accelerated corrosion of the RDE unit at higher temperatures.

2. Results and Discussion

2.1. Catalyst Characterization

High quality RDE measurements require thin and uniform films over the entire surface of the electrode [28]. Figure 2 shows micrographs of catalyst films obtained using catalyst ink of the proposed composition. The film in the left image was produced by the method of rotational drying at a certain angle, which is detailed below in Section 3.2, and the film in the right image was produced by stationary drying. Figure 2a shows the film with a uniform structure over the entire surface of the glassy carbon (GC) electrode. The catalyst film in Figure 2b is unevenly distributed over the GC electrode; defects and voids are observed. Thus, according to the micrographs, the proposed method for catalytic ink drying allows obtaining films with a high degree of uniformity.
Figure 3 shows the surface of the GC electrode coated with Pt40/C (40 wt.% of Pt on carbon carrier) before and after the RDE measurements in the HClO4 electrolyte. The distribution of the catalyst particles is very even, free from formation of agglomerates and defects, despite unavoidable formation of some islands (Figure 3a,c) at the sub-micrometer level during drying [29]. The image of the layer surface after the RDE measurements (Figure 3c) did not show significant changes in the surface structure in comparison with the initial layer structure (Figure 3a).
The image of the film before the RDE measurements taken at a higher magnification (Figure 3b) shows thin μm-sized patches of the ionomer layer (circled in red), which is described in [30]. The ionomer performs several important functions such as attaching catalyst particles to one another, obtaining homogenous and well-dispersed ink, and uniform application of the ink [31]. Such ionomer patches disappeared after the series of RDE measurements (Figure 3d), which points to the dissolution of the ionomer film in the electrolyte. The dissolution of the ionomer may lead to an increase in the catalyst layer activity due to better mass transport of oxygen to active sites and lower electrical resistance [31,32]. However, the amount of ionomer in the film and the number of patches were rather small, so the ionomer dissolution should not have a significant effect on the quality of measurements.
The absence of significant changes in the layer surface structure allows for obtaining high-quality results during the entire series of RDE measurements in the wide temperature range.

2.2. Electrochemical Studies

Figure 4 presents cyclic voltammograms (CVs) of different catalysts. The EASA values are 42, 83, 55 and 57 m2 g−1 Pt for Pt40/C, Pt20/C, Pt20SnO25/C and Pt20SnO210/C, respectively.
The CVs demonstrate well-defined hydrogen adsorption/desorption peaks in the potential range of 0.05–0.40 V vs. RHE. Two peaks in the hydrogen desorption region at 0.13 and 0.20 V correspond to platinum (110) and (100) active sites, respectively, [26,33]. The sample with SnO2 content of 10 wt.% demonstrates a shift of the hydrogen desorption peak to more positive potentials, which may be attributed to the effect of the presence of tin dioxide on the surface structure of platinum nanoparticles and various distributions of crystal orientations [34]. The smaller EASA for composite catalysts can be explained by the more facile adsorption of OH species, which impedes hydrogen adsorption and by possible blockage of Pt sites by SnO2 nanoparticles. A small peak at 0.73 V for the CVs of the modified catalysts may be attributed to the oxygen adsorption from the dissociation of water on the surface of tin dioxide particles [35].
Measurement of the electrolyte solution resistance is an important factor for obtaining correct values from processing the data acquired by the RDE technique. An almost two-fold decrease in resistance with increasing temperature was observed (Figure 5). The obtained values of electrolyte solution resistance were used in further calculations and plotting of polarization curves.
Figure 6 shows the polarization curves for a Pt20/C-coated electrode at different electrolyte temperatures and at the electrode rotation speed of 1600 rpm. The current density on the diffusion-limited plateau of the polarization curves increases with temperature, indicating that the decrease in oxygen solubility with temperature was lower than the increase in the oxygen diffusion coefficient [36]. A shift of the half-wave potential of the polarization curves to the high-potential region is also observed with increasing temperature, which characterizes the positive dependence of the reaction rate on temperature.
Table 1 shows the values of EASA and activities of the catalysts under the study at 20 °C and the values obtained by the other groups. The polyol method is the most commonly used method for the synthesis of platinum catalysts.
The electrochemical surface area and activity values for the Pt40/C catalyst proved to be lower than the values reported in literature. The relatively low values can be explained by the catalyst preparation technique [20], in which the carbon support was preliminarily impregnated with a precursor followed by reduction. At the same time, with an increase in the platinum content, the average size of platinum nanoparticles exceeded 3.5 nm, which was described in our previous work [40]. According to [41], for particles with a size less than 3.5 nm, predominant for catalysts with a lower platinum content, the crystal orientation of the surface (110) dominates, and its activity in ORR is higher than for (100) Pt sites [42].
The values of EASA and mass activity of the Pt20/C are comparatively higher than the values reported in the literature. The difference in the values may be due to different techniques used for recording polarization curves and catalyst film preparation. However, a comparative analysis of the obtained values is possible.
Catalysts with hybrid support and SnO2 loading of 20 wt.% have the lower values of EASA and mass activity, which can be attributed to the increase in the particle size with an increase in the concentration of tin dioxide [20].
Figure 7 shows the Koutecky–Levich (K-L) plots for catalysts with a platinum content of 20 wt.% at 50 °C. The plots remain almost parallel up to a potential of 0.9 V, pointing to a weak dependence of the number of transferred electrons on the potential and applicability of the K-L theory for increased temperatures. The number of transferred electrons for all samples is close to 4, which suggests high selectivity of samples up to elevated temperatures.
Table 2 shows the values of kinetic current and catalyst activity at various temperatures.
The Pt40/C catalyst demonstrates low activity values over the whole temperature range due to the larger platinum nanoparticle size and the smaller fraction of the platinum surface involved in the ORR.
The polarization curves for catalysts with platinum content of 20 wt.% have close values of kinetic current density and activity up to 30 °C. The structure of the sample with the lowest tin dioxide content Pt20SnO25/C is mainly represented by individual highly dispersed Pt and SnO2 particles, which participate as a co-catalyst in ORR. Thus, Pt20SnO25/C catalyst has close values of activity in comparison with Pt20/C up to 50 °C. The Pt20SnO210/C sample shows mass activity reduction at about 30% at the temperature of 50 °C, which can be explained by an agglomeration of tin dioxide particles, as well as the formation of Pt-SnO2 hetero-clusters.
Based on the obtained values, the sample with tin dioxide content of 5 wt.% has activity similar to that of the standard catalyst over the entire temperature range, which together with high stability of this sample [20], makes it promising for use in PEMFCs under a wide range of operating conditions. The sample with tin dioxide content of 10 wt.% also demonstrates high activity, however, at 50 °C, the activity was comparatively lower. Nevertheless, under conditions of increased temperature and low humidity, the use of this catalyst in PEMFC is justified taking into account the high water retention ability of tin dioxide nanoparticles.
The Arrhenius plots for the Pt20/C and Pt20SnO210/C samples under study at the potential of 0.9 V are shown in Figure 8b,d. Some plots demonstrate non-linear behavior. This could be due to activation of additional platinum catalytic centers at increased temperatures and larger kinetic current error than the calculated one.
According to the slope of the Arrhenius plots, activation energies for the ORR are 20 ± 1, 25 ± 1, 30 ± 1 and 25 ± 2 kJ mol−1 for Pt40/C, Pt20/C, Pt20SnO25/C and Pt20SnO210/C, respectively. The activation energies obtained in our study are −28 and 21 ± 3 kJ mol−1, respectively, which agrees well with the values for the catalyst with Pt content of 20 wt.% reported in papers [43,44]. Close activation energy values for the composite catalysts and reference platinum samples is indicative of high efficiency of catalysts with tin dioxide in the ORR. Lower values of the activation energy for Pt20SnO210/C than for Pt20SnO25/C may be attributed to the presence of the Pt-SnO2 hetero-clusters and higher activation energy of the SnO2 active centers of the Pt20SnO25/C sample.
Thus, note should be made of the low dependence of the ORR kinetics on temperature and high efficiency of the electrocatalysts down to temperatures close to 0 ℃. Such catalysts allow maintaining the PEMFC performance during the “cold start” from low ambient temperatures to standard operating conditions. Catalysts modified with tin dioxide have activity comparable to that of Pt20/C, despite the partial blockage of Pt sites by SnO2 nanoparticles. The high activity of the composite catalysts suggests that tin dioxide participates in the ORR as a co-catalyst, and high water adsorption ability of SnO2 makes these catalysts efficient at low temperatures due to prevention of formation of ice crystals in MEA, and at increased temperatures by preventing the PEMFC components from overdrying.

3. Materials and Methods

3.1. Preparation of Catalysts

Pt/x-SnO2/C and Pt/C catalysts were synthesized using a polyol method in ethylene glycol described in our previous works [20,24].

3.2. Electrode Preparation

Thick catalyst films on a polished GC disk electrode (0.102 cm2; Volta, Russia) were prepared by dropping catalyst ink with an Eppendorf micropipette. The catalyst ink included Nafion® used as a binder in the amount of 7 wt.% of the specified catalyst weight.
The catalyst inks were prepared by ultrasound treatment. The mixture consisting of 12 mg of catalyst powder, 0.18 mL of isopropanol and 0.2 mL of 0.5% Nafion® solution in DI water [45] was treated in ultrasonic homogenizer (the catalyst amount was ca. 60 mg mL−1) for 1 h. Five microliters of the prepared catalyst ink were dropped with an Eppendorf micropipette onto the surface of polished GC electrode. The drying method was obtained from experiments with the used ink composition. The GC electrode was neatly tilted at an angle of 45° and evenly rotated at 60 rpm until the ink became dry. The electrode was dried in air at room temperature. The final catalyst loading was about 0.2 mg cm−2.

3.3. Electrochemical Studies

The cyclic voltammograms (CVs) were measured in 0.1M HClO4 at 25 °C using a three-electrode glass cell equipped with a polished GC working electrode, a Pt wire counter electrode placed in a fritted glass tube, and RHE as a reference electrode connected to the electrochemical cell by a Luggin capillary. We used the RHE because it has a low level of impurities and does not require potential correction [27].
The electrode was activated in an N2 saturated 0.1 M HClO4 solution at the potential range of 0.05 to 1.20 V at a 50 mV s−1 sweep rate for about 30 cycles until a stable CV was obtained. The CVs registered at the potential range of 0.05 to 1.20 V and at a 20 mV s−1 sweep rate were used to characterize the catalyst.
The measurements were performed using a CorrTest CS350 electrochemical workstation (Wuhan, Corrtest Instruments Corp., Ltd., Hubei, Wuhan, China). The catalyst EASA was calculated using the hydrogen adsorption-desorption peaks at 0.05–0.40 V vs. RHE as described in [24,46].
After taking the CV measurements, the background current was measured by running the ORR sweep profile in an N2-purged electrolyte solution at the electrode rotation speed of 1600 rpm in the potential range of 0.05–1.05 V and at the 10 mV s−1 sweep rate. Then the electrolyte was purged with oxygen for 1.5–2.0 h. The ORR polarization curves were recorded at the electrode rotation speeds of 500, 700, 900, 1200 and 1600 rpm at the 10 mV s−1 sweep rate and in the potential range of 0.05–1.05 V. The electrolyte was purged by O2 for at least 5 min between the measurements.
To obtain polarization curves at different temperatures (1, 10, 20, 30 and 50 °C), a thermostatically controlled chamber of the three-electrode cell was connected to a water thermostat. The upper temperature was limited by increased acid evaporation and accelerated corrosion of the RDE unit, and the lower temperature limit was determined by freezing of the water used as cooling/heating medium in the thermostat. To accurately control the electrolyte temperature, a thermocouple was installed inside the cell. The polarization curves were recorded after reaching thermal equilibrium. At each temperature the impedance spectroscopy curves were recorded to take the electrolyte solution resistance into consideration in the Koutecky–Levich calculations. The electrochemical impedance spectroscopy curves were plotted with the CorrTest CS350 electrochemical workstation at frequencies of 105 to 0.1 Hz.

3.4. Catalysts Surface Structure Characterization

The uniformity of the formed films was evaluated using a Levenhuk DTX 90 digital optical microscope (USA) with 10–300× magnification.
The electrode surface morphology was observed using a Versa 3D DualBeam scanning electron microscope (SEM) (FEI, Hillsboro, OR, USA) under vacuum. Low vacuum images were obtained using a low vacuum secondary electron detector (LVSED) and a concentric backscattered detector (CBS).

3.5. Calculation Methods

The Koutecky–Levich equation was used for calculation of the kinetic and mass-transfer components of the measured currents:
1 j = 1 j k + 1 j d = 1 | j k | 1 0.62 n F D O 2 2 3 ν 1 6 C O 2 b ω 1 2
where j (mA cm−2) is the measured current density, jk (mA cm−2) is the kinetic current density, jd (mA cm−2) is the diffusion-limited current density, n is the average number of electrons transferred per O2 molecule, F (C mole−1) is the Faraday constant, D O 2 (cm2 s−1) is the diffusion coefficient of O2 in the 0.1M HClO4 solution, ν (cm2 s−1) is the kinematic viscosity,   C O 2 b (mol cm−3) is the oxygen concentration in the electrolyte, and ω (rad s−1) is the electrode angular rotation rate.
The area-specific (Sa) and mass-specific (Ma) catalyst activity were determined by the following equations:
S a = i k E A S A × m P t ,
M a = i k m P t   ,
where mPt (mgPt) is the mass of Pt on the electrode and ik (mA) is the measured kinetic current.
To evaluate the average number of transferred electrons n at temperatures different from room temperature, it is important to take into consideration the temperature dependence of the diffusion coefficient, kinematic viscosity and concentration of dissolved oxygen. The concentration of dissolved oxygen at each temperature was calculated according to Henry’s law:
C O 2 b e [ Δ s o l H R ( 1 T ) ] = c o n s t
where Δ s o l H is the enthalpy of solution, R is the universal gas constant and Δ s o l H R is a constant, equal to 1700 K for oxygen [47]. The values of kinetic viscosity at different temperatures were taken from [48], and the oxygen diffusion coefficient was calculated by the Stokes−Einstein equation [49]:
D O 2 ν / T = c o n s t
The activation energy of the ORR according to the Arrhenius equation was estimated by multiplying the gas constant by the slope of the log (ik) vs. 1/T plot. The kinetic current was normalized with the respect to the concentration of dissolved oxygen at each temperature.

4. Conclusions

The ORR kinetics on composite carbon-supported Pt-SnO2 catalysts were investigated over a wide temperature range. The CVs show a reduction in the EASA for the composite catalysts due to possible blockage of Pt active sites by SnO2 nanoparticles. The RDE technique was used for determination of kinetic and mass components of the ORR current. K-L plots at 50 °C remain almost parallel up to the potential of 0.9 V, which suggests that the number of transferred electrons weakly depends on the potential, and that the K-L theory can be applied for increased temperatures. The number of transferred electrons for all samples is close to 4, which points to high selectivity of samples at temperatures up to 50 °C. The catalysts modified with tin dioxide demonstrate high activity despite the partial blockage of Pt active sites by SnO2 nanoparticles and agglomeration. The relatively high activity of the composite catalysts is explained by participation of tin dioxide in the ORR as a co-catalyst. The Pt20SnO25/C catalyst has close values of activity in comparison with Pt20/C up to 50 °C, but the Pt20SnO210/C sample shows mass activity reduction at about 30% at the temperature of 50 °C, which can be explained by the agglomeration of tin dioxide particles, as well as the formation of Pt-SnO2 hetero-clusters. According to the slope of the Arrhenius plots, the activation energies for the ORR lie in the range from 20 to 30 kJ mol−1 (at temperatures ranging from 1 to 50 °C). The values of activation energies obtained in our study are in a good agreement with the values of ~28 and 21 ± 3 kJ mol−1 reported in the literature. Thus, the approach used in our study to synthesize modified catalysts allows for producing hybrid catalysts having not only improved stability, but also high activity at the temperature up to 50 °C for the Pt20SnO25/C and up to 30 °C for the Pt20SnO210/C.

Author Contributions

Conceptualization, V.N.F., N.A.I., R.M.M., D.D.S.; methodology, R.M.M., N.A.I. and D.D.S.; software, R.M.M. and A.A.Z.; validation, N.A.I., V.N.F. and S.A.G.; investigation, R.M.M., E.V.K., D.D.S., N.A.I. and E.A.S.; resources, V.N.F.; writing—original draft preparation, R.M.M. and N.A.I.; writing—review and editing, R.M.M., N.A.I., E.A.S. and S.A.G.; visualization, D.D.S., A.A.Z. and R.M.M.; supervision, S.A.G.; project administration, V.N.F.; funding acquisition, V.N.F. All authors have read and agreed to the published version of the manuscript.

Funding

The synthesis of catalysts and morphological study of catalytic films were financially supported by the Russian Foundation for Basic Research (project No. 20-08-00927). The electrochemical studies of catalysts over a wide temperature range were financially supported by the Russian Foundation for Basic Research (project No. 18-29-23030).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Derendyaev, M.A.; Koryakin, D.V.; Filalova, E.M.; Yalmaev, A.B.; Galin, M.Z.; Gerasimova, E.V.; Antipov, A.E.; Levchenko, A.V.; Dobrovolsky, Y.A. Role of Platinum Loading on the Characteristics at the PEM Fuel Cell Cathode. Nanotechnol. Russ. 2020, 15, 797–806. [Google Scholar] [CrossRef]
  2. Pavlov, V.I.; Gerasimova, E.V.; Zolotukhina, E.V.; Don, G.M.; Dobrovolsky, Y.A.; Yaroslavtsev, A.B. Degradation of Pt/C electrocatalysts having different morphology in low-temperature PEM fuel cells. Nanotechnol. Russ. 2016, 11, 743–750. [Google Scholar] [CrossRef]
  3. Gebel, G. Structural evolution of water swollen perfluorosulfonated ionomers from dry membrane to solution. Polymer 2000, 41, 5829–5838. [Google Scholar] [CrossRef]
  4. Chandan, A. High temperature (HT) polymer electrolyte membrane fuel cells (PEMFC)—A review. J. Power Sources 2013, 231, 264–278. [Google Scholar] [CrossRef]
  5. Laribi, S.; Mammar, K.; Sahli, Y.; Koussa, K. Air supply temperature impact on the PEMFC impedance. J. Energy Storage 2018, 17, 327–335. [Google Scholar] [CrossRef]
  6. Luo, Z.; Chang, Z.; Zhang, Y.; Liu, Z.; Li, J. Electro-osmotic drag coefficient and proton conductivity in Nafion® membrane for PEMFC. Int. J. Hydrogen Energy 2010, 35, 3120–3124. [Google Scholar] [CrossRef]
  7. Rajbongshi, B.M.; Shaneeth, M.; Verma, A. Investigation on sub-zero start-up of polymer electrolyte membrane fuel cell using un-assisted cold start strategy. Int. J. Hydrogen Energy 2020, 45, 34048–34057. [Google Scholar] [CrossRef]
  8. Lin, R.; Weng, Y.; Lin, X.; Xiong, F. Rapid cold start of proton exchange membrane fuel cells by the printed circuit board technology. Int. J. Hydrogen Energy 2014, 39, 18369–18378. [Google Scholar] [CrossRef]
  9. Mu, Y.T.; He, P.; Ding, J.; Chen, L.; Tao, W.Q. Numerical study of the gas purging process of a proton exchange membrane fuel cell. Energy Procedia 2017, 105, 1967–1973. [Google Scholar] [CrossRef]
  10. Yao, L.; Ma, F.; Peng, J.; Zhang, J.; Zhang, Y.; Shi, J. Analysis of the failure modes in the polymer electrolyte fuel cell cold-start process—Anode dehydration or cathode pore blockage. Energies 2020, 13, 256. [Google Scholar] [CrossRef] [Green Version]
  11. Mishler, J.; Wang, Y.; Mukherjee, P.P.; Mukundan, R.; Borup, R.L. Subfreezing operation of polymer electrolyte fuel cells: Ice formation and cell performance loss. Electrochim. Acta 2012, 65, 127–133. [Google Scholar] [CrossRef] [Green Version]
  12. Pineri, M.; Gebel, G.; Davies, R.J.; Diat, O. Water sorption–desorption in Nafion® membranes at low temperature, probed by micro X-ray diffraction. J. Power Sources 2007, 172, 587–596. [Google Scholar] [CrossRef]
  13. Ivanova, N.A.; Spasov, D.D.; Grigoriev, S.A.; Kamyshinsky, R.A.; Peters, G.S.; Mensharapov, R.M.; Seregina, E.A.; Millet, P.; Fateev, V.N. On the influence of methanol addition on the performances of PEM fuel cells operated at subzero temperatures. Int. J. Hydrogen Energy 2021, 46, 18116–18127. [Google Scholar] [CrossRef]
  14. Mensharapov, R.M.; Fateev, V.N. The Membranes with Modified Surface to Stabilize Water Balance of Fuel Cell under Low Humidity Conditions: A Model Study. Nanotechnol. Russ. 2020, 15, 363–369. [Google Scholar] [CrossRef]
  15. Spasov, D.D.; Mensharapov, R.M.; Zasypkina, A.A.; Ivanova, N.A. Nanoscale Pt-SnO2 Heteroclusters in Electrocatalysts for Oxygen Reduction and Methanol Oxidation Reactions. Nanotechnol. Russ 2020, 15, 723–729. [Google Scholar] [CrossRef]
  16. Hou, S.; Chen, R.; Zou, H.; Shu, T.; Ren, J.; Li, X.; Liao, S. High-performance membrane electrode assembly with multi-functional Pt/SnO2–SiO2/C catalyst for proton exchange membrane fuel cell operated under low-humidity conditions. Int. J. Hydrogen Energy 2016, 41, 9197–9203. [Google Scholar] [CrossRef] [Green Version]
  17. Chen, F.; Mecheri, B.; d’Epifanio, A.; Traversa, E.; Licoccia, S. Development of Nafion/Tin oxide composite MEA for DMFC applications. Fuel Cells 2010, 10, 790–797. [Google Scholar] [CrossRef] [Green Version]
  18. Brutti, S.; Scipioni, R.; Navarra, M.A.; Panero, S.; Allodi, V.; Giarola, M.; Mariotto, G. SnO2-Nafion® nanocomposite polymer electrolytes for fuel cell applications. Int. J. Nanotechnol. 2014, 11, 882–896. [Google Scholar] [CrossRef]
  19. Rizo, R.; Sebastián, D.; Lázaro, M.J.; Pastor, E. On the design of Pt-Sn efficient catalyst for carbon monoxide and ethanol oxidation in acid and alkaline media. Appl. Catal. B 2017, 200, 246–254. [Google Scholar] [CrossRef] [Green Version]
  20. Spasov, D.D.; Ivanova, N.A.; Pushkarev, A.S.; Pushkareva, I.V.; Presnyakova, N.N.; Chumakov, R.G.; Presnyakov, M.Y.; Grigoriev, S.A.; Fateev, V.N. On the influence of composition and structure of carbon-supported Pt-SnO2 hetero-clusters onto their electrocatalytic activity and durability in PEMFC. Catalysts 2019, 9, 803. [Google Scholar] [CrossRef] [Green Version]
  21. Kuriganova, A.B.; Leontyeva, D.V.; Ivanov, S. Electrochemical dispersion technique for preparation of hybrid MOx–C supports and Pt/MOx–C electrocatalysts for low-temperature fuel cells. J. Appl. Electrochem. 2016, 46, 1245–1260. [Google Scholar] [CrossRef]
  22. Chen, Y.; Wang, J.; Meng, X. Pt-SnO2/nitrogen-doped CNT hybrid catalysts for proton-exchange membrane fuel cells (PEMFC): Effects of crystalline and amorphous SnO2 by atomic layer deposition. J. Power Sources 2013, 238, 144–149. [Google Scholar] [CrossRef]
  23. Li, M.; Cullen, D.A.; Sasaki, K. Ternary electrocatalysts for oxidizing ethanol to carbon dioxide: Making Ir capable of splitting C–C bond. J. Am. Chem. Soc. 2013, 135, 132–141. [Google Scholar] [CrossRef]
  24. Pushkarev, A.S.; Pushkareva, I.V.; Ivanova, N.A.; du Preez, S.P.; Bessarabov, D.; Chumakov, R.G.; Stankevich, V.G.; Fateev, V.N.; Evdokimov, A.A.; Grigoriev, S.A. Pt/C and Pt/SnOx/C catalysts for ethanol electrooxidation: Rotating disk electrode study. Catalysts 2019, 9, 271. [Google Scholar] [CrossRef] [Green Version]
  25. Guan, J.; Zan, Y.; Shao, R. Phase Segregated Pt-SnO2/C Nanohybrids for Highly Efficient Oxygen Reduction Electrocatalysis. Small 2020, 16, 2005048. [Google Scholar] [CrossRef] [PubMed]
  26. Hussain, S.; Kongi, N.; Erikson, H.; Rähn, M.; Merisalu, M.; Matisen, L.; Alonso-Vante, N. Platinum nanoparticles photo-deposited on SnO2-C composites: An active and durable electrocatalyst for the oxygen reduction reaction. Electrochim. Acta 2019, 316, 162–172. [Google Scholar] [CrossRef]
  27. Zhang, N. Pt/Tin Oxide/Carbon Nanocomposites as Promising Oxygen Reduction Electrocatalyst with Improved Stability and Activity. J. Electrochim. Acta 2014, 117, 413–419. [Google Scholar] [CrossRef]
  28. Garsany, Y.; Barurina, O.A.; Swider-Lyons, K.E.; Kocha, S.S. Experimental methods for quantifying the activity of platinum electrocatalysts for the oxygen reduction reaction. Anal. Chem. 2010, 82, 6321–6328. [Google Scholar] [CrossRef] [PubMed]
  29. Ke, K.; Hiroshima, K.; Kamitaka, Y.; Hatanaka, T.; Morimoto, Y. An accurate evaluation for the activity of nano-sized electrocatalysts by a thin-film rotating disk electrode: Oxygen reduction on Pt/C. Electrochim. Acta 2012, 72, 120–128. [Google Scholar] [CrossRef]
  30. Orfanidi, A.; Rheinländer, P.J.; Schulte, N.; Gasteiger, H.A. Ink solvent dependence of the ionomer distribution in the catalyst layer of a PEMFC. J. Electrochem. Soc. 2018, 165, F1254. [Google Scholar] [CrossRef]
  31. Kocha, S.S.; Zack, J.W.; Alia, S.M.; Neyerlin, K.C.; Pivovar, B.S. Influence of ink composition on the electrochemical properties of Pt/C electrocatalysts. ECS Trans. 2013, 50, 1475. [Google Scholar] [CrossRef]
  32. Shinozaki, K.; Pivovar, B.S.; Kocha, S.S. Enhanced oxygen reduction activity on Pt/C for Nafion-free, thin, uniform films in rotating disk electrode studies. ECS Trans. 2013, 58, 15. [Google Scholar] [CrossRef]
  33. Gómez, R.; Orts, J.M.; Álvarez-Ruiz, B.; Feliu, J.M. Effect of temperature on hydrogen adsorption on Pt (111), Pt (110), and Pt (100) electrodes in 0.1 M HClO4. J. Phys. Chem. B 2004, 108, 228–238. [Google Scholar] [CrossRef]
  34. Jerkiewicz, G. Electrochemical hydrogen adsorption and absorption. Part 1: Under-potential deposition of hydrogen. Electrocatalysis 2010, 1, 179–199. [Google Scholar] [CrossRef]
  35. Lim, D.H.; Choi, D.H.; Lee, W.D.; Lee, H.I. A new synthesis of a highly dispersed and CO tolerant PtSn/C electrocatalyst for low-temperature fuel cell; its electrocatalytic activity and long-term durability. Appl. Catal. B 2009, 89, 484–493. [Google Scholar] [CrossRef]
  36. Elezović, N.R.; Babić, B.M.; Krstajić, N.V.; Gojković, S.L.; Vračar, L.M. Temperature dependence of the kinetics of oxygen reduction on carbon-supported Pt nanoparticles. J. Serb. Chem. Soc. 2008, 73, 641–654. [Google Scholar] [CrossRef]
  37. Garsany, Y.; Junjie, G.; St-Pierre, J.; Rocheleau, R.; Swider-Lyons, K.E. Analytical procedure for accurate comparison of rotating disk electrode results for the oxygen reduction activity of Pt/C. J. Electrochem. Soc. 2014, 161, F628. [Google Scholar] [CrossRef]
  38. Garsany, Y.; Singer, I.L.; Swider-Lyons, K.E. Impact of film drying procedures on RDE characterization of Pt/VC electrocatalysts. J. Electroanal. Chem. 2011, 662, 396–406. [Google Scholar] [CrossRef]
  39. Sun, S.; Zhang, G.; Geng, D. A highly durable platinum nanocatalyst for proton exchange membrane fuel cells: Multiarmed starlike nanowire single crystal. Angew. Chem. 2011, 123, 442–446. [Google Scholar] [CrossRef]
  40. Pushkareva, I.V.; Pushkarev, A.S.; Kalinichenko, V.N.; Chumakov, R.G.; Soloviev, M.A.; Liang, Y.; Millet, P.; Grigoriev, S.A. Reduced Graphene Oxide-Supported Pt-Based Catalysts for PEM Fuel Cells with Enhanced Activity and Stability. Catalysts 2021, 11, 256. [Google Scholar] [CrossRef]
  41. Sattler, M.L.; Ross, P.N. The surface structure of Pt crystallites supported on carbon black. Ultramicroscopy 1986, 20, 21–28. [Google Scholar] [CrossRef] [Green Version]
  42. Markovic, N.; Gasteiger, H.; Ross, P.N. Kinetics of oxygen reduction on Pt (hkl) electrodes: Implications for the crystallite size effect with supported Pt electrocatalysts. J. Electrochem. Soc. 1997, 144, 1591. [Google Scholar] [CrossRef]
  43. Paulus, U.A.; Schmidt, T.J.; Gasteiger, H.A.; Behm, R.J. Oxygen reduction on a high-surface area Pt/Vulcan carbon catalyst: A thin-film rotating ring-disk electrode study. J. Electroanal. Chem. 2001, 495, 134–145. [Google Scholar] [CrossRef]
  44. Jäger, R.; Härk, E.; Steinberg, V.; Lust, E. Influence of Temperature on the Oxygen Electroreduction Activity at Nanoporous Carbon Support. ECS Trans. 2015, 66, 47. [Google Scholar] [CrossRef]
  45. Kirakosyan, S.A.; Alekseenko, A.A.; Guterman, V.E.; Novomlinskii, I.N.; Men’Shchikov, V.S.; Gerasimova, E.V.; Nikulin, A.Y. De-Alloyed PtCu/C catalysts of oxygen electroreduction. Russ. J. Electrochem. 2019, 55, 1258–1268. [Google Scholar] [CrossRef]
  46. Grigoriev, S.A.; Millet, P.; Fateev, V.N. Evaluation of carbon-supported Pt and Pd nanoparticles for the hydrogen evolution reaction in PEM water electrolysers. J. Power Sources 2008, 177, 281–285. [Google Scholar] [CrossRef]
  47. Karbowiak, T.; Gougeon, R.D.; Alinc, J.B.; Brachais, L.; Debeaufort, F.; Voilley, A.; Chassagne, D. Wine oxidation and the role of cork. Crit. Rev. Food Sci. Nutr. 2009, 50, 20–52. [Google Scholar] [CrossRef]
  48. Vanysek, P. CRC Handbook of Chemistry and Physics, 95th ed.; CRC Press: Boca Raton, FL, USA, 2011; pp. 6–231. [Google Scholar]
  49. Zhu, Z.; Liu, Q.; Liu, X.; Shui, J. Temperature impacts on oxygen reduction reaction measured by the rotating disk electrode technique. J. Phys. Chem. C 2020, 124, 3069–3079. [Google Scholar] [CrossRef]
Figure 1. Simplified scheme of the ORR with direct and indirect pathways. In the scheme: ki and ki are the reaction rate constants at various stages of the process in the forward and reverse directions, respectively.
Figure 1. Simplified scheme of the ORR with direct and indirect pathways. In the scheme: ki and ki are the reaction rate constants at various stages of the process in the forward and reverse directions, respectively.
Catalysts 11 01469 g001
Figure 2. Optical micrographs of a catalytic film on the surface of the electrode fabricated by rotational (a) and stationary (b) drying.
Figure 2. Optical micrographs of a catalytic film on the surface of the electrode fabricated by rotational (a) and stationary (b) drying.
Catalysts 11 01469 g002
Figure 3. SEM images of Pt40/C-coated electrodes before (a,b) and after (c,d) RDE measurements.
Figure 3. SEM images of Pt40/C-coated electrodes before (a,b) and after (c,d) RDE measurements.
Catalysts 11 01469 g003
Figure 4. CVs of 40 wt.%-Pt/C (Pt40/C), 20 wt.%-Pt/C (Pt20/C), 20 wt.%-Pt/5 wt.%-SnO2/C (Pt20SnO25/C), 20 wt.%-Pt/10 wt.%-SnO2/C (Pt20SnO210/C) recorded in the 0.1M HClO4 solution at 25 °C (scan rate 20 mV s−1).
Figure 4. CVs of 40 wt.%-Pt/C (Pt40/C), 20 wt.%-Pt/C (Pt20/C), 20 wt.%-Pt/5 wt.%-SnO2/C (Pt20SnO25/C), 20 wt.%-Pt/10 wt.%-SnO2/C (Pt20SnO210/C) recorded in the 0.1M HClO4 solution at 25 °C (scan rate 20 mV s−1).
Catalysts 11 01469 g004
Figure 5. Resistance of 0.1 M HClO4 solution at different temperatures in the 1−50 °C range.
Figure 5. Resistance of 0.1 M HClO4 solution at different temperatures in the 1−50 °C range.
Catalysts 11 01469 g005
Figure 6. RDE polarization curves (20 mV s−1) for the ORR at a Pt20/C-coated disk electrode rotating at 1600 rpm in a 0.1 M HClO4 solution at different temperatures in the range from 1 to 50 °C.
Figure 6. RDE polarization curves (20 mV s−1) for the ORR at a Pt20/C-coated disk electrode rotating at 1600 rpm in a 0.1 M HClO4 solution at different temperatures in the range from 1 to 50 °C.
Catalysts 11 01469 g006
Figure 7. K-L plots calculated from the Pt20/C (a), Pt20SnO25/C (b), Pt20SnO210/C (c) RDE polarization curves at 50 °C. Inset: n value as a function of the electrode potential.
Figure 7. K-L plots calculated from the Pt20/C (a), Pt20SnO25/C (b), Pt20SnO210/C (c) RDE polarization curves at 50 °C. Inset: n value as a function of the electrode potential.
Catalysts 11 01469 g007
Figure 8. Arrhenius plots for the ORR in 0.1 M HClO4 for the samples: Pt40/C (a) Pt20/C (b) Pt20SnO25/C (c) Pt20SnO210/C (d). The mass transfer is adjusted, the currents are normalized to the concentration of dissolved oxygen.
Figure 8. Arrhenius plots for the ORR in 0.1 M HClO4 for the samples: Pt40/C (a) Pt20/C (b) Pt20SnO25/C (c) Pt20SnO210/C (d). The mass transfer is adjusted, the currents are normalized to the concentration of dissolved oxygen.
Catalysts 11 01469 g008aCatalysts 11 01469 g008b
Table 1. Comparison of catalyst parameters: EASA, area-specific (Sa) and mass-specific (Ma) activity of catalysts at 20 °C and at potential 0.9 V.
Table 1. Comparison of catalyst parameters: EASA, area-specific (Sa) and mass-specific (Ma) activity of catalysts at 20 °C and at potential 0.9 V.
CatalystSynthesis MethodEASA, m2g−1Sa, mA cm2Ma, mA mg−1Scan Rate, mV s−1
Pt40/CPolyol method420.14 ± 0.0156 ± 110
Pt40/C [37]Polyol method49 ± 10.73 ± 0.03359 ± 1420
Pt40/C [38]Polyol method49 ± 10.48 ± 0.03230 ± 1020
Pt40/C [39]Polyol method430.219010
Pt20/CPolyol method830.25 ± 0.04207 ± 410
Pt20/C [26]Carbonyl route63 ± 20.37 ± 0.0175 ± 110
Pt20/C [26]Photo-deposition72 ± 20.41 ± 0.0294 ± 510
Pt20/C [38]Polyol method55 ± 100.33 ± 0.06180 ± 620
Pt20/C [28]-660.201605
Pt20SnO25/CPolyol method550.37 ± 0.01205 ± 210
Pt20SnO210/CPolyol method570.36 ± 0.01205 ± 510
Pt20SnO220/C [26]Photo-deposition33 ± 10.44 ± 0.0549 ± 410
Pt20SnO220/C [27]Polyol method530.315610
Table 2. Catalyst activity parameters: kinetic current density (jk), area-specific (Sa) and mass-specific (Ma) activity of catalysts in the temperature range from 1 to 50 °C at potential of 0.9 V.
Table 2. Catalyst activity parameters: kinetic current density (jk), area-specific (Sa) and mass-specific (Ma) activity of catalysts in the temperature range from 1 to 50 °C at potential of 0.9 V.
T, °C110203050
Pt40/C (EASA = 42 m2 g−1)
jk, mA cm−23.7 ± 0.24.3 ± 0.34.6 ± 0.15.2 ± 0.25.7 ± 0.2
Sa, mA cm−20.12 ± 0.010.13 ± 0.010.14 ± 0.010.16 ± 0.010.17 ± 0.01
Ma, mA mg−147 ± 251 ± 456 ± 162 ± 269 ± 2
Pt20/C (EASA = 83 m2 g−1)
jk, mA cm−24.2 ± 0.24.8 ± 0.25.6 ± 0.16.9 ± 0.111.5 ± 0.4
Sa, mA cm−20.19 ± 0.010.22 ± 0.010.25 ± 0.040.31 ± 0.010.52 ± 0.02
Ma, mA mg−1154 ± 7178 ± 8207 ± 4256 ± 2428 ± 15
Pt20SnO25/C (EASA = 55 m2 g−1)
jk, mA cm−22.9 ± 0.13.6 ± 0.14.5 ± 0.16.3 ± 0.17.9 ± 0.3
Sa, mA cm−20.24 ± 0.010.30 ± 0.050.37 ± 0.010.52 ± 0.010.65 ± 0.02
Ma, mA mg−1132 ± 1165 ± 3205 ± 2286 ± 6357 ± 13
Pt20SnO210/C (EASA = 57 m2 g−1)
jk, mA cm−23.8 ± 0.15.0 ± 0.15.9 ± 0.27.1 ± 0.28.0 ± 0.1
Sa, mA cm−20.24 ± 0.040.31 ± 0.010.36 ± 0.010.44 ± 0.010.49 ± 0.01
Ma, mA mg−1134 ± 2175 ± 2205 ± 5248 ± 6280 ± 4
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mensharapov, R.M.; Ivanova, N.A.; Spasov, D.D.; Kukueva, E.V.; Zasypkina, A.A.; Seregina, E.A.; Grigoriev, S.A.; Fateev, V.N. Carbon-Supported Pt-SnO2 Catalysts for Oxygen Reduction Reaction over a Wide Temperature Range: Rotating Disk Electrode Study. Catalysts 2021, 11, 1469. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121469

AMA Style

Mensharapov RM, Ivanova NA, Spasov DD, Kukueva EV, Zasypkina AA, Seregina EA, Grigoriev SA, Fateev VN. Carbon-Supported Pt-SnO2 Catalysts for Oxygen Reduction Reaction over a Wide Temperature Range: Rotating Disk Electrode Study. Catalysts. 2021; 11(12):1469. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121469

Chicago/Turabian Style

Mensharapov, Ruslan M., Nataliya A. Ivanova, Dmitry D. Spasov, Elena V. Kukueva, Adelina A. Zasypkina, Ekaterina A. Seregina, Sergey A. Grigoriev, and Vladimir N. Fateev. 2021. "Carbon-Supported Pt-SnO2 Catalysts for Oxygen Reduction Reaction over a Wide Temperature Range: Rotating Disk Electrode Study" Catalysts 11, no. 12: 1469. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11121469

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