Open Access Article
Gonzalo Montielabc,
Eduardo Fuentes-Quezadabc,
Mariano M. Bruno*d,
Horacio R. Cortibc and
Federico A. Viva
*bc
aInstituto Nacional de Tecnología Industrial, Av. General Paz 5445 San Martín, Buenos Aires, Argentina
bDepartamento de Física de la Materia Condensada, Comisión Nacional de Energía Atómica, Av. General Paz 1499, San Martín, Buenos Aires, Argentina. E-mail: viva@tandar.cnea.gov.ar
cInstituto de Nanociencia y Nanotecnología, CNEA-CONICETCentro Atómico Constituyentes, Av Gral Paz 1499, San Martin, Buenos Aires, Argentina
dInstituto de Investigaciones en Tecnologías Energéticas y Materiales Avanzados (IITEMA), Universidad Nacional de Río Cuarto, Facultad de Cs. Exactas Físico Química y Naturales, Departamento de Química, 5800 Río Cuarto, Argentina. E-mail: mbruno@exa.unrc.edu.ar
First published on 18th August 2020
Mesoporous carbons (MCs) with different pore sizes were synthesized and evaluated as a catalyst support for fuel cells. The MCs were obtained from resorcinol–formaldehyde precursors, polymerized in the presence of polydiallyldimethylammonium chloride (cationic polyelectrolyte) as a structuring agent and commercial silica (Sipernat® or Aerosil®) as the hard template. The MC obtained with Aerosil® shows a broad pore size distribution with a maximum at 21 nm. On the other hand, the MCs with Sipernat® show a bimodal pore size distribution, with a narrow peak centered at 5 nm and a broad peak with a maximum ca. 30 nm. All MCs present a high specific surface area (800–1000 m2 g−1) and total pore volume ranging from 1.36 to 1.69 cm3 g−1. PtRu nanoparticles were deposited onto the MC support by an impregnation–reduction method with NaBH4 at 80 °C in basic media. The electrochemical characterization reveals improved electrocatalysis towards the methanol oxidation for the catalyst deposited over the carbon with the highest total pore volume. This catalyst also presented the highest CO2 conversion efficiency, ca. 80%, for the methanol oxidation as determined by differential electrochemical mass spectroscopy analysis. Moreover, the catalyst as a fuel cell anode showed the best performance, reaching a power density of 125 mW cm−2 at 90 °C with methanol as fuel and dry O2.
The metal catalysts in the CL are employed as nanoparticles dispersed over a conductive support. The most common support for fuel cell catalysts are carbon-based materials.11–14 The structure of the catalyst support can determine the catalyst nanoparticles stability and activity toward methanol oxidation.15,16 The support can maximize the particle dispersion as well as the electroactive catalyst area, while improves the mass transfer of reactants and products.3,6 Due to the mentioned cost of noble metals, the DMFC anode and cathode catalyst loadings must drop below 1.0 mg cm−2 from the present 2.0–4.0 mg cm−2 while maintaining the cell performance.12 Recently, different routes for the preparation of advanced nanostructured carbon materials have emerged, providing extra fine-tuning of the supported catalyst electroactivity.12,13,17–19
An ideal carbon support should allow the preparation of highly dispersed catalytic nanoparticles, whereas the porosity should ensure the ionomer penetration for proton transport while allowing a facile access path for reactants and by-products.17,20 It was shown that pores over 20 nm can ensure the formation of the triple phase boundary (TPB) region by allowing an optimal contact between the catalyst and the ionomer (Nafion).21,22 Additionally, the carbon surface defects, surficial groups and the microporosity have been identified as anchoring sites for metal nanoparticle, improving particle dispersion and catalytic activity.23,24 The porosity can also influence the residence time of the reactant and the by-products near of the catalyst nanoparticles.12,25–27 In this sense, a carbon material with hierarchical pore size distribution could offer a plausible way to guarantee the mentioned features. In previous works, we have demonstrated that carbon with micropores/mesopores obtained by using polyelectrolyte as a structuring agent can improve the catalyst performance in a fuel cell.17,22,23,28 More recently, we have presented a new method for the synthesis of carbon with dual mesopores size which can satisfy the requirements of an appropriate porous carbon support.29 This study analyzes the effect of the pore size distribution and the pore volume fraction of the carbon support on the fuel cell performance. The strategy was to produce carbons with a different pore size distribution and mesopore volume, preserving the rest of the textural properties. The set of properties were tuned by adding pore forming agents in the resorcinol formaldehyde polymerization media.29–31
In the present work, we describe the preparation and characterization of PtRu catalyst, synthesized by the impregnation–reduction method, supported on three different MCs. The carbon supports were obtained by carbonization of a resorcinol–formaldehyde polymer combining commercial silica as hard template and polydiallyldimethylammonium chloride as structuring agent. The MCs were characterized by N2 adsorption–desorption isotherms. The supported catalysts were characterized by powder X-ray diffraction (PXRD), transmission electron microscopy (TEM), and energy-dispersive X-ray spectroscopy (EDS). Stripping of CO was used for the determination of the catalyst electrochemical surface area (ECSA), whereas the electrocatalytic activity was determined by cyclic voltammetry (CV), chronoamperometry, and potentiodynamic differential electrochemical mass spectrometry (DEMS) measurements. Finally, the performance of the MEA with the prepared materials as anode catalyst were evaluated with methanol as fuel.
000–200
000 g mol−1, Sigma-Aldrich), while the commercial silica powders Aerosil® 200 and Sipernat® 50 (EVONIK) were employed as the HT. The Aerosil® 200 is a nonporous fumed silica with a specific surface area between 50 and 500 m2 g−1, and a particles with sizes between 5 and 50 nm.32 The Sipernat® 50 are porous silica particles of 70 μm in diameter and a surface area of 475 m2 g−1.29 Briefly, two solutions were prepared, the solution A containing 2 g of resorcinol (R) (99.0% ACS, Sigma-Aldrich), 1 g of the SA solution, and 0.25 g of sodium acetate (trihydrate PA, Ciccarelli) were dissolved in 50 mL of milli-Q water. In the solution B, the HT was completely dispersed in 4 g of methanol (Biopack 99.8% wt), 5 g of glycerol (Biopack 99.5% wt), and 50 mL of milli-Q water. Both solutions were mixed and heated in a reflux system with magnetic stirring. The polymerization began by adding 1.4 g of a formaldehyde solution (F) (37% wt, Sigma-Aldrich) to the dispersion. After 45 min, another 3 g of F were added. The heating and magnetic stirring was maintained for 20 minutes and cooled it to 25 °C. Glycerol and methanol were added to the aqueous polymerization media as dispersant and wetting agents, respectively, rendering the polymer in a powder form. By this synthesis procedure, three different MC were obtained; MC A05 was synthesized by using 0.50 g of Aerosil® 200, while MC S15 and MC S30 were prepared by using 1.50 g and 3.00 g of Sipernat® 50, respectively. The composite resin was vacuum filtered from the solution, dried in a vacuum oven at 100 °C overnight and then carbonized under a N2 stream of 1 L min−1 in a tubular furnace (Indef model T-150) from 20 °C to 1000 °C at a heating rate of 3 °C min−1 and finally held at 1000 °C for 120 minutes. The HT from the carbon was removed by etching with 3 M NaOH solution at 60 °C for 24 h under constant stirring. MCs were washed with milli-Q water in Soxhlet apparatus until a neutral pH was obtained and finally dried in a vacuum oven at 110 °C overnight.
:
1 (NaBH4 to metal salt) to the suspension. Heating was maintained for 2 h, followed by stirring for 24 h at room temperature. The powder obtained was washed with milli-Q water in a Soxhlet apparatus and finally dried in a vacuum oven at 80 °C for 24 h.
PXRD pattern of the catalysts were obtained using a Siemens D5000 diffractometer with a Cu Kα source operating at 40 kV and 30 mA. The angle extended from 20 to 100° with a step size of 0.02° and a counting time of 2 s. TEM images were acquired with JEOL 100 CX II, meanwhile EDS was performed using an SEM Philips 505 with EDAX detector to quantify the atomic ratio of Pt and Ru in the catalysts. Thermogravimetric analyses (TGA) were performed with a Q600 SDT Thermal Analyzer from TA Instruments controlled by Q Series software. The experiments were carried out using approximately 10 mg of sample in alumina pans under air atmosphere (Praxair A10.0XD-T 99.99%), with a heating rate of 10 °C min−1 and a gas flow rate of 50 mL min−1. The metal content on the supported catalysts were calculated from the difference between the initial and final weights.
:
1 ratio of Nafion to catalyst was deposited with a micropipette(1–10 μL, Rontaig) over the working electrode (WE), consisting of a glassy carbon disk (5 mm diam.) mounted on a Teflon rod. The WE was previously cleaned with isopropyl alcohol in an ultrasonic bath for 10 minutes and after deposition of the catalyst suspension, was dried in a vacuum oven at 80 °C for 10 minutes. CV, and chronoamperometry experiments were performed in a three-electrode cell. The counter electrode (CE) consisted of a coiled Pt wire 0.5 mm in diameter and 30 cm length, whereas a Ag/AgCl (sat. KCl) electrode was used as a reference electrode (RE). All potentials were converted against the standard hydrogen electrode (SHE). The upper potential limit for the voltammetric determinations was set to 0.8 V vs. SHE to avoid the formation of irreversible ruthenium oxides or Ru dissolution.36 The electrochemical surface area (ECSA) was measured by CO stripping voltammetry (ESI-Fig. 1†). The cell was filled with 0.5 M H2SO4 (95–97%, Merck) solution and saturated with CO (RG, Indura) for 45 min while the WE potential was maintained at 0.2 V vs. SHE. After the time elapsed, and while maintaining the potential, the solution was purged with N2 (RG, Indura) for 15 min to remove the unadsorbed CO, and immediately two scans between 0.05 and 0.8 V vs. SHE at a scan rate of 1 mV s−1 were performed. The catalysts ECSA was calculated based on the mass deposited onto the WE and the CV peak integral using the reference charge value of 420 μC cm−2 for the oxidation of a CO monolayer,37 and employed to convert the measured current (i) to current density (j). The catalysts CVs were performed on a 1 M methanol solution in 0.5 M H2SO4 by sweeping the potential between 0.05 and 0.8 V vs. SHE at 2, 5, 10 and 20 mV s−1 (ESI-Fig. 2†). The chronoamperometry were performed in the same electrolyte solution at 0.5 V vs. SHE for 1 h. These measures were employed to obtain the poisoning rate and the turn over frequency (TOF) as was calculated in previous reports.38–40 All electrochemical measurements were performed with an Autolab PGSTAT302N potentiostat.
The experiments were performed in a flow electrochemical cell designed for the DEMS.17,41 The WE consisted of a glassy carbon disk built ad hoc (6 mm in diameter) with a 1 mm diameter hole in the center through which the reactant flowed during the measurement. A suspension of the supported catalyst as described in Section 2.4 was spread over the working electrode (WE). The inlet port, from the cell to the DEMS pressure chamber, is through a stainless steel frit. The electrochemical cell opening is separated from the frit by a porous Teflon membrane (0.02 μm pore, 50 μm thickness, 50% porosity by Gore). The WE lies on top of the Teflon® membrane, separated from it by a 100 μm Teflon® gasket, allowing the formation of a thin liquid reacting layer. Volatile product species diffuse through the membrane to reach the mass spectrometer. The electrolyte flow rate (0.12 mL min−1) was controlled by a syringe pump (PC11U, Apema). Calibration by CO stripping and further quantification of the conversion efficiency was carried out as previously reported.17,42
:
10
:
7 mass proportion, respectively, and spread on one side of a 5 cm2 Toray C paper TGP-H 60 10% PTFE coated (Fuel Cell Technologies), for a final electrode loading ca. 3 mg cm−2. A Nafion 212 membrane (Ion Power) was placed in between the electrodes and hot pressed at 150 °C and 40 bar for 25 min. The Nafion membrane was previously treated by boiling in H2O2 3% wt (H2O2 30% wt, Bio-pack) followed by H2SO4 3% wt (95–97%wt, Merck). The MEAs were mounted in a standard single cell housing with serpentine flow fields (Fuel Cell Technologies, Inc.). Teflon gasket films (50–150 μm) were employed as seal and the cell uniformly bolted with a torque of 2.3 Nm. After assembly of the cell, the MEA was re-humidified by circulating water at 80 °C overnight. Galvanodynamic polarization test was performed with a test station (University Test Station model from Fuel Cell Technologies) at 90 °C, from the open circuit voltage (OCV) to a voltage close of short circuit (0.05 V) while circulating 1 M methanol (Merk, HPLC grade) through the anode and dry O2 (RG 4.8, Indura) through the cathode. A Gilson Minipuls 3 peristaltic pump was used to circulate the methanol solution and a digital mass flow meter (MC 200 from Alicat Scientific) to control the O2 flows. For all the measurements the methanol flow was set to 2.0 mL min−1 while the O2 flow was 200 SCCM.
| Sample | HT | HT/R | SBETa (m2 g−1) | Vmb (cm3 g−1) | Vsc (cm3 g−1) | Vld (cm3 g−1) | Vtotale (cm3 g−1) |
|---|---|---|---|---|---|---|---|
| a Specific surface area using the BET method.b Micropore volume from DR equation.c Volume of small mesopores (2 < d < 7 nm).d Volume of large mesopores (7 < d < 50 nm).e Total pore volume at P/P° = 0.99. | |||||||
| MC A05 | Aerosil | 0.25 | 796 | 0.33 | 0.11 | 0.77 | 1.50 |
| MC S15 | Sipernat | 0.37 | 787 | 0.32 | 0.34 | 0.54 | 1.36 |
| MC S30 | Sipernat | 0.75 | 1000 | 0.40 | 0.59 | 0.66 | 1.69 |
The aforementioned results suggest that the specific synthesis method yield samples that present similar textural properties despite the different mesopore distribution, monomodal for the MC A05 and bimodal for the MC S15. Additionally, bimodal samples can be prepared with different ratios of small and large mesopores.
![]() | ||
| Fig. 2 TEM images of the different supported catalysts (A) PtRu/MC A05, (B) PtRu/MC S15, (C) PtRu/MC S30. Inset: corresponding particle size distributions. | ||
The PXRD diffractograms of the synthesized catalyst are shown in Fig. 3. The diffraction peaks at 2θ angles ca. 40, 46, 67, and 81 are due to face-centered cubic (fcc) crystalline Pt, assigned to the planes (111), (200), (220), and (311), respectively, with a slight shifting to higher 2θ values due to the presence of Ru.43–45 The lattice parameters of the metal nanoparticles were calculated by indexing the first three peaks yielding 3.900 ± 0.002 Å for PtRu/MC A05, 3.895 ± 0.002 Å for PtRu/MC S15, and 3.896 ± 0.002 Å for PtRu/MC S30. Using the reference value of 3.923 Å, the lattice parameter was used to estimate the Ru atomic fraction alloyed with Pt.46 The values obtained were 19% for PtRu/MC A05, 23% for PtRu/MC S15, and 22% for PtRu/MC S30, with an error of ±2%. The alloying Ru percentages, which are of the same order of magnitude for the three catalyst, shows that most of the Ru is present in an amorphous phase. Nonetheless, the percentage values are high for a catalyst that has not been subjected to a thermal treatment after deposition of the metal nanoparticles.17,47
The EDS results obtained for the catalyst indicated a Pt
:
Ru atomic ratio of 52
:
48 for PtRu/MC A05, 49
:
51 for PtRu MC S15, and 51
:
49 for PtRu/MC S30 close to the 1
:
1 atomic nominal ratio. The thermogravimetric analysis of the metal catalyst indicated metal loadings of 49% for PtRu/MC A05, 61% for PtRu MC S15, and 62% for PtRu/MC S30. The metal catalyst percentage over MC S15 and S30 are close to the intended value of 60% while is slightly lower over MC A05. As discussed above, the lower fraction of small size mesopores provides the least favorable support for particle anchoring.
The catalytic activity of the prepared catalysts was assessed in a 1 M methanol + 0.5 M H2SO4 aqueous solution. Fig. 4 shows the voltammograms for the three catalysts at 2 mV s−1. The peak current density for the methanol oxidation increases in the order PtRu/MC S30 > PtRu/MC S15 > PtRu/MC A05, while the onset potential from the CVs are 0.49 V for the MC A05, 0.39 V for the MC S15, and 0.37 for the MC S30, following the same order. This suggest a more facile methanol oxidation on PtRu/MC S30.
![]() | ||
| Fig. 4 Cyclic voltammograms for the three different catalysts in 1 M methanol + 0.5 M H2SO4 at 2 mV s−1. | ||
The chronoamperometry measurements of the catalysts at 0.5 V vs. SHE for a 30 min period are shown in Fig. 5. The rapid current decay shows a more restricted diffusion for PtRu/MC A05 than for PtRu/MC S30. The slope of the current transient in the chronoamperograms were employed for the determination of the catalyst poisoning rate (δ),39,40 while the steady-state current density was used for the calculation of the turnover frequency (TOF).17,38,40 The chronoamperometry slope between 500 and 1800 s has been attributed to the catalyst poisoning by CO, whereas at longer times the decay is related to anion adsorption.17 The values obtained for the poisoning rate and TOF, respectively, are presented in Table 2. The PtRu/MC S30 presents the highest number of reacting molecules per site, since TOF is directly proportional to the steady state current density.40 The TOF for the supported catalyst increases as the nanoparticle size decreases, i.e. as the support Vs increases. The results show the relationship between the turn over frequency and the nanoparticle size and therefore the relation with the support porosity. The poisoning rate shows similar results for the three catalysts. As the δ is related to the adsorption of intermediate species, the results would indicate that either the intermediates leave rapidly the catalyst surface, or the methanol oxidation produces a low amount of intermediates. The DEMS analysis, vide infra, shows that the catalyst conversion efficiency is high. Previous reported values of δ and TOF parameters suggest that mesoporous support allows the intermediates to rapidly leave the catalyst surface.17,48–51
![]() | ||
| Fig. 5 Chronoamperometry determination of the three different catalysts in 1 M methanol + 0.5 M H2SO4 at 0.5 V vs. SHE. | ||
| Catalyst sample | δ (% s−1) | TOF (molecules per site·per s) |
|---|---|---|
| PtRu/MC A05 | 0.0125 | 0.018 |
| PtRu/MC S15 | 0.0083 | 0.023 |
| PtRu/MC S30 | 0.0118 | 0.034 |
![]() | ||
| Fig. 6 Potentiodynamic DEMS measurement of PtRu/MC A05, PtRu/MC S15 and PtRu/MC S30 in 1 M methanol + 0.5 M H2SO4 with the signal corresponding to m/z = 44 (CO2). | ||
The polarization plots show a better performance of the catalyst deposited over the bimodal carbons compared to the monomodal support. The PtRu/MC S30, which has the highest amount of small mesopores (Fig. 1B and Table 1), shows the highest power density followed by PtRu/MC S15. The PtRu/MC A05 presents the lowest performance of the prepared catalysts in this report, but still it performs better than a previously reported PtRu catalyst deposited over a monomodal MC.17,22 The reported MC, synthesized by a process without hard template, exhibited a PSD with a narrow peak centered at 20 nm and a total pore volume of ca 1 cm3 g−1. The commercial catalyst, with Vulcan® carbon as support, displays the lowest performance. Vulcan® is a carbon black with a low surface area (252 m2 g−1) without mesoporosity or microporisity (VT = 0.63 cm3 g−1).14 The measured fuel cell polarizations indicates that the support's mesoporosity, particularly the small mesopores, have a positive influence on the methanol oxidation, improving the overall cell performance.
The electrochemical results and the fuel cell performance measurements show the effect of the support textural properties on the PtRu nanoparticles catalytic properties. The carbon support synthesized presents an adequate surface area and microporosity that provides good particle size and distribution across the surface. The presence of the mesopores also allows the formation of the TPB and a sponge like structure which facilitates the access of methanol onto the catalyst and the departure of CO2.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/d0ra05676f |
| This journal is © The Royal Society of Chemistry 2020 |