Open Access Article
Jan Dismas Buriánek
a,
Martin Prokop
a,
Tomas Bystron
*a,
Martin Veselý
b,
Lukáš Koláčný
b,
Bruna Ferreira Gomes
c,
Carlos Manuel Silva Loboc,
Matija Gatalo
de,
Luka Pavko
de,
Nejc Hodnik
d,
Martin Paidar
a,
Christina Roth
c,
Miran Gaberscek
d and
Karel Bouzek
a
aDepartment of Inorganic Technology, University of Chemistry and Technology Prague, Technická 5, 166 28 Praha, Czech Republic. E-mail: bystront@vscht.cz
bDepartment of Organic Technology, University of Chemistry and Technology Prague, Technická 5, 166 28 Praha, Czech Republic
cElectrochemical Process Engineering, University of Bayreuth, Universitätsstraße 30, Bayreuth 95447, Germany
dDepartment of Materials Chemistry, National Institute of Chemistry, Hajdrihova 19, p.p. 660, SI-1001 Ljubljana, Slovenija
eReCatalyst, Hajdrihova 19, 1001 Ljubljana, Slovenia
First published on 7th January 2026
In the emerging hydrogen energy economy, proton-exchange membrane fuel cells (PEMFCs) serve as a key enabling technology, yet their cost is among other things dominated by platinum group metals-based cathode catalysts. This paper is focused on investigation of intermetallic Pt–Co and Pt–Ni nanoparticles supported on carbon (Ketjen black, reduced graphene oxide) as low-Pt-load candidates for high-temperature PEMFCs (HT-PEMFCs) operated at elevated temperature ∼180 °C in the presence of concentrated phosphoric acid. Catalytic activity toward the oxygen reduction reaction (ORR) was quantified by rotating electrode measurements (exchange current densities, Tafel slopes), and stability was probed by leaching in 97.6 wt% H3PO4 at 180 °C followed by post-exposure characterisation. A suite of techniques – XAS, XRD, TEM/EDS, XRF, Raman spectroscopy and ICP-OES – was used to study changes in composition and structure during degradation. All alloy catalysts showed in HClO4 at 25 °C higher ORR activity than commercial Pt/C. However, exposure to concentrated H3PO4 at 180 °C caused electrochemically active surface area loss, reduced ORR activity and supported Pt crystallite growth, Co/Ni dissolution, and surface reorganisation. Comparatively, reduced graphene oxide-supported catalyst was more resistant to ripening and dealloying than its Ketjen black analogue, and Pt–Ni alloy was more stable than Pt–Co. Overall, the results disentangle the roles of the carbon support and alloy composition and outline activity – stability trade-offs that guide the design of low-Pt loading cathodes for HT-PEMFCs.
Broader contextThe transition to a low-carbon energy economy increasingly relies on hydrogen as an energy carrier for storage, decarbonization of existing technologies, or transportation applications. Among hydrogen conversion technologies, fuel cells attract significant attention of study. High-temperature proton exchange membrane fuel cells (HT-PEMFCs) offer distinct advantages over their low-temperature counterparts, including efficient recovery of reaction heat, simplified cooling, and reduced hydrogen purity requirements. However, their long-term viability and commercialisation critically depend on the development of catalysts that remain durable and active under harsh operating conditions (phosphoric acid, 180 °C). Alloying Pt with transition metals in combination with suitable carbon supports provides a strategy to lower Pt loading while maximising efficiency and durability. In this study, three alloy catalysts – Pt–Co on Ketjenblack carbon, Pt–Co on reduced graphene oxide, and Pt–Ni on Ketjenblack carbon – are evaluated against a commercial Pt/C catalyst. The comparison focuses on changes in composition, morphology, and oxygen reduction reaction activity after exposure to hot phosphoric acid, aiming to identify optimal alloy/support combinations for stable HT-PEMFC performance. |
Low-temperature PEMFCs operate with a proton-exchange membrane, typically based on a perfluorinated sulfonated polymer (e.g., Nafion®), serving as a polymer electrolyte. Most LT-PEMFCs utilise H2 and O2, resulting in the production of water. The key advantages of LT-PEMFCs are their high energy efficiency and power density per unit weight (this applies to the FC itself, and not the heat exchange unit, see below). The relatively low operating temperature of LT-PEMFCs, around 70 °C, offers the benefit of reasonable stability of materials during prolonged operation (slow degradation of materials) and enables quick starts, including the possibility of cold starts.5
However, the low operating temperature of LT-PEMFCs does present severe practical challenges. These cells are extremely sensitive to impurities like CO, which act as catalyst poisons.6 Moreover, effective water management is essential because the perfluorinated sulfonated polymer membranes require sufficient humidity for proper operation. On the other hand, too much humidity in the system leads to catalyst layer flooding which causes severe mass transport limitations. Most importantly, the heat exchange unit of LT-PEMFC requires a significant heat exchange surface making it heavy and bulky. Finally, the waste heat generated by LT-PEMFCs remains underutilised.7
These disadvantages can be effectively addressed by HT-PEMFCs, which operate at temperatures ranging from 160 to 180 °C. At these temperatures, perfluorinated sulfonated polymers dry out and lose their ionic conductivity. Instead, polybenzimidazole (PBI)-based membranes doped with concentrated H3PO4 are usually used.8–14 HT-PEMFCs are not sensitive to the presence of CO at concentrations of up to 3 vol%, allowing for the use of H2 derived from sources like CH3OH, natural gas, or propane.5 HT-PEMFCs also do not require complicated water management, as the water produced exits the cell in gaseous form. The higher operating temperature permits the use of a more compact and lighter cooling system, facilitates temperature control, and enables the utilisation of waste heat, e.g., for steam cogeneration.15,16 Nevertheless, the presence of H3PO4 leaching from the membrane induces corrosion of other cell components, and H3PO4 (including its impurities) can block the catalyst surface, leading to mediocre cell performance during operation.17–21
In the case of PEMFCs, electrochemical reactions occur within the catalytic layers of anodes and cathodes, which are supplied with H2 and air/O2, respectively, and are in contact with the proton-exchange membrane. It is widely accepted that selection of especially the cathode electrocatalyst catalysing the O2 reduction reaction (ORR) remains the most challenging task. First, sluggish ORR kinetics usually leads to a high activation overpotential and significant energy efficiency losses.22–24 Secondly, the low stability of cathode catalysts represents an additional issue.25,26 Currently, Pt nanoparticles immobilised on a carbon support represent the state-of-the-art ORR electrocatalyst in an acidic environment.27,28 Interestingly, various Pt alloys with Co, Ni, or Cu show potential to enhance the electrocatalytic activity towards ORR in aqueous environments.29–38 The presence of an alloying metal modifies the catalyst surface, potentially increasing its intrinsic catalytic activity. Therefore, a partial replacement of Pt with less rare metals could lead to a significant reduction of catalyst cost.39,40
The electrochemical behaviour of an alloy catalyst significantly depends on its composition and morphology. It is important to note that its structure and composition may evolve during PEMFC operation. This phenomenon is not yet fully understood and is currently extensively investigated.21,41–43 The influence of H3PO4 itself and its impurities (e.g. H3PO3) on the cell performance is substantial.18,41,44–48 It is also widely accepted that during the operation of Pt–Co(Ni) alloy catalysts in an acidic environment, Co(Ni) dissolves from the surface, thereby exposing the electrochemically active sites of the Pt catalyst.49–51 This process, occurring due to the limited stability of alloy catalysts, is known as (electro)chemical dealloying and/or leaching and represents one of the inevitable phenomena associated with their use.52 While initial catalyst activity is enhanced by this process, further degradation usually leads to a significant decrease in specific activity of catalysts.41–43 The amount of Co(Ni) in the alloy, as well as the conditions under which (electro)chemical dealloying occurs (electrode polarisation, temperature, impurities present in the H3PO4, etc.), affect significantly the catalyst's electrochemical behaviour. This is manifested by changes, e.g., in surface morphology, the distribution of metal nanoparticle sizes or porosity of the support.50
Catalyst activity is also influenced by the type of its support. Carbon-black-based materials currently represent the standard materials. Other studied materials with potential for application are graphene derivatives such as reduced graphene oxide (rGO), which possess distinct physicochemical properties compared to carbon black. They exhibit superior electronic conductivity, higher 2D crystallinity, and fewer structural defects due to the higher carbon content in sp2 hybridisation.53 Some observations suggest that these properties might result in improved thermodynamic stability and, consequently, potentially prolonged catalyst lifetime during fuel cell operation.54,55
The present work deals with alloy catalysts synthetises by a novel method, namely Pt–Co on amorphous Ketjen black (KB) carbon support (Pt–Co/KB), Pt–Co on reduced graphene oxide (rGO) support (Pt–Co/rGO) and Pt–Ni on KB support (Pt–Ni/KB).56 The catalysts are compared with a commercial Pt catalyst on carbon black support (Pt/C). The aim of this work is to determine the ORR activity of the studied catalysts in aqueous and H3PO4 media at ambient and elevated temperature, respectively. For a deeper understanding, the changes of the ORR activity of the catalysts during interaction with H3PO4 at elevated temperature and the composition and structural changes of the catalysts after interaction with concentrated H3PO4 at elevated temperature, were analysed.
H2O2 was used for purification and concentration of H3PO4.44,57 The 85 wt% H3PO4 was mixed with the 30 wt% H2O2 in volumetric ratio of 2
:
1. The mixture was poured in a PTFE beaker, slowly heated to 160 °C and left at this temperature for 24 h. Thus prepared concentrated H3PO4 was poured in a glass storage bottle while still hot and its concentration was determined by acid–base titration using automatic titrator (Easy Pro, Mettler Toledo). 0.1 M NaOH (Standard solution, Penta) was used as a titration agent. The concentration of H3PO4 was 97.6 ± 0.6 wt% (i.e., 70.9 wt% P4O10) with volumetric density of 1.849 ± 0.006 g cm−3 corresponding to the expected literature value.58,59 This H3PO4 was used for all measurements.
Measurements were performed in a glass electrochemical cell (50 cm3). The electrolyte was aqueous solution of HClO4, the cell was maintained at 25 °C using a thermostat (PC14 DC 10, Haake, Germany). To enable broader comparison with literature data and ensure robust validation of the measured values, experiments were conducted at two HClO4, concentrations: 0.5 mol dm−3 and 0.1 mol dm−3. The results were consistent across both conditions; therefore, subsequent measurements were performed exclusively in 0.1 mol dm−3 HClO4. Measurements were performed on a potentiostat (PGSTAT302N, Autolab Metrohm, Switzerland), RDE revolution was controlled by a rotor (AFMSRCE, Pine Research, USA). Athree-electrode setup was used. The counter electrode was a Pt wire of about 260 mm2 placed in a glass tube separated from the main compartment of the electrolyte by a porous frit. The reference electrode was a fresh reversible hydrogen electrode (RHE, Hydroflex, Gaskatel GmbH, Germany), placed directly into the electrolyte solution. The working electrode was carried in a shaft (Pine Research, USA) and placed approximately 5 mm below the electrolyte level.
Each prepared thin film electrode was first stabilised by potential cycling (200 cycles between 1.2 and 0.05 V vs. RHE, sweep rate of 300 mV s−1) in the N2 saturated electrolyte before the LSV measurement. Subsequently, the LSVs on the prepared thin film electrodes were measured in the potential range (1.0 V to 0.05 V vs. RHE) at electrode revolution rates of 400, 900, 1600 and 2500 rpm and at a sweep rate of 10 mV s−1 in the N2 saturated electrolyte and then in O2 saturated electrolyte. The currents measured in the N2 saturated electrolyte were subtracted from the currents obtained in the O2 saturated electrolyte. To evaluate the electrochemically active surface area (ECSA), the electrolyte was, immediately after measuring the ORR (i.e., in the same experimental setup), for 60 s saturated with CO, and then 10 min purged with N2. This was followed by 2 cycles of CV (0.05 V to 1.0 V vs. RHE) at a sweep rate of 20 mV s−1 on stationary electrode. A numerical integration (trapezoidal method) was used to integrate the CO-stripping region. A value of 450 µC cm−2 was used as the specific charge for the electrochemical desorption of the CO monolayer. All experimental voltammograms were corrected for the IR drop during the measurement. The value of 1.170 V is used as
in 0.1 M HClO4, 25 °C, O2 saturation, O2 activity was corrected using H2O vapour pressure over the solution, see ref. 60. The current densities were in all cases determined with respect to the geometric surface area of the electrode. The Koutecky–Levich (K–L) analysis (eqn (1)) of voltammograms measured at various revolution rates provided the kinetic current density values (jk), which were subsequently used to evaluate the Tafel slope.
![]() | (1) |
The limiting current density jlim is taken as the median of the points lying within a 20 mV interval in the region of most negative current densities j. To ensure reproducible results, each measurement was performed in triplicate and with at least three different working electrodes. A representative set of voltammograms measured at RDE (revolution rate of 1600 rpm, 20 mV s−1, in 0.1 M HClO4) is shown in Fig. S2.1, see SI. In addition to that, K–L analysis using voltammograms measured at various revolution rates of (400–2500 rpm, sweep 10 mV s−1, in 0.5 M HClO4) was performed, see resulting Tafel plots shown in Fig. S2.2 in SI. Both methods (analysis of jk obtained from LSV at 1600 rpm and jk obtained from LSVs at multiple revolution rates) provided comparable parameters.
:
50 with propan-2-ol) was placed onto the rod disk. Subsequently, 20 mm3 of purified 97.6 wt% H3PO4 (diluted 1
:
9 with water) was dropped onto the thin film and the rods were placed in an oven (SP-55 Easy, Kambič, Slovenia) pre-heated to 180 °C for catalyst degradation. After defined heating time (1, 3, 8, 24 and 72 h), the rods were gradually removed from the oven and used for electrochemical tests.Electrochemical characterisation of thin film-modified RREs was performed in a glass electrochemical cell (50 cm3). The electrolyte was 0.1 mol dm−3 HClO4, the cell was maintained at 25 °C using a thermostat (PC14 DC 10, Haake, Germany). Measurements were performed on a potentiostat (CompactStat.e, IVIUM Technologies, Netherland), electrode revolution rate was controlled by an electrode rotator (AFMSRCE, Pine Research, USA). A three-electrode setup was used. The counter electrode was a graphite rod (SPK grade, Ultra “F” purity, Thermo Scientific, USA) of about 40 cm2 placed in a glass tube separated from the main compartment of the electrolyte (through a thin connection tube). The reference electrode was a fresh RHE (Hydroflex, Gaskatel GmbH, Germany), placed separately and connected with the measuring cell by a glass bridge filled with the same electrolyte. The working RRE was mounted into an in-house made shaft and its tip was immersed approximately 5 mm below the electrolyte level. The shaft was constructed from CNC axis with centering and adjusted to a precise diameter on a lathe.
After degradation, each prepared thin film-modified RRE was first cycled 50 times in the potential range of 1.0 to 0.05 V vs. RHE at a sweep rate of 300 mV s−1 in the N2 saturated electrolyte. Subsequently, the LSVs of the prepared thin film-modified RREs were measured in the potential range (1.0 V to 0.05 V vs. RHE) at a sweep rate of 20 mV s−1 in the O2 saturated electrolyte at revolving electrode (1600 rpm). To evaluate the ECSA immediately after measuring the ORR (i.e., in the same experimental setup), the same protocol as described in Section 2.2.1 was used. All experimental voltammograms were corrected for the IR drop. The value of 1.170 V is used as
(25 °C, O2 saturation, pH value of 0.1 M HClO4, O2 activity was corrected using H2O vapour pressure over the solution, see ref. 60). The current densities were in all cases determined with respect to the geometric surface area of the electrode. The Koutecky–Levich (K–L) analysis (eqn (1)) of voltammograms measured at various revolution rates provided the kinetic current density values (jk), which were subsequently used to evaluate the Tafel slope. The limiting current density jlim is taken as the median of the points lying within a 20 mV interval in the region of minimum current densities j. To ensure reproducible results, each measurement was performed in triplicate and with at least three different working electrodes.
700 rpm. The combined liquid fractions were diluted to a total volume of 20 cm3. The solid fraction was then decanted several times with demineralised water until neutral pH of the suspension was achieved and then dried under N2 at room temperature. The solid residue was split in two parts. The first part was analysed by the various instrumental methods (XRD, XRF, XAS, Raman spectroscopy, XPS and TEM), the second one (about 15 mg) was dissolved in aqua regia (4 cm3 conc. HNO3 + 12 cm3 conc. HCl, 24 h, boiling under reflux). The aqua regia solution was then diluted to 200 cm3 and (after decantation of undissolved carbon) analysed by IPC-OES. Triplicate measurements were taken only at time 0 h and 24 h, the other times are represented by only one sample.For XRF measurements performed with a Performix spectrometer (Thermo ARL, Switzerland) with Rh lamp (4.2 kW), the samples were deposited on Ge (111) crystal. Standard-free analysis using the UNIQUANT software integrated into software Oxsas was used for processing the obtained data.
Raman spectra were measured with a dispersive Raman spectrometer model DXR Raman Microscope (Thermo Scientific, USA) equipped with an Olympus confocal microscope. The excitation source was a diode-excited Nd:YAG laser with a wavelength of 532 nm and an input power of 10 mW. A lattice of 900 notches mm−1 was used. A multichannel thermoelectrically cooled CCD camera was used as detector. The samples were measured with a 50× objective with a measurement footprint of about 1 µm2. The samples were measured through a 50 µm slit aperture. Measurements were performed with a power of 0.3 mW, measurement time of 30 s and 10 spectral accumulations, with photobleaching for 30 s. Three spots from each sample were measured.
XPS measurements were performed on an Omicron ESCA Probe P (Scienta Omicron, Germany). Survey spectra were measured with 0.4 eV step and pass energy 50 meV. Excitation source was Al with energy Kα = 1486.7 eV, pressure in chamber around 3.5 × 10−10 mbar.
EFTEM Jeol 2200 FS (Jeol GmbH, Germany), 200 kV EFTEM with Schottky FEG and In-Column Omega Energy-filter was used for TEM analysis of samples of the leached catalysts. Energy dispersive spectroscopy (EDS, Oxford Instruments, UK) with a resolution of 1–2.4 nm was used for elemental analysis.
The ex situ XAS measurements were performed at the P64 beamline (DESY, Hamburg, Germany). Samples of the pristine catalysts and catalysts after leaching were deposited between two Kapton tapes (80 µm, DuPont, USA) to achieve loading of approximately 1 mgmetals cm−2. The measurements were performed in fluorescence (PIPS detector) and transmission mode. Although both detection modes yielded comparable results, fluorescence measurements provided a better signal-to-noise ratio. However, self-absorbance effects were assessed by comparison with transmission data and corrected when necessary. Spectra at the Pt L3-edge and Co and Ni K-edges were collected. The corresponding metallic foils were measured simultaneously with the samples to align the energy. The software Bessy_47b and SimXLite, both developed by the beam scientists at KMC-3 beamline (BESSY II – Berlin), were used for the data processing and EXAFS fitting, respectively. The ranges used for the EXAFS fit were k = (2.0; 12.0) Å−1 and R = (1.0; 3.0) Å for Pt L3-edge, and k = (4.0; 12.0) Å−1 and R = (1.0; 3.0) Å for Co K-edge, and k = (3.0; 11.0) Å−1 and R = (1.0; 3.0) Å for Ni K-edge. The metallic foils were used to determine the amplitude factor, S02, that was equal to 0.867, 0.796 and 0.787 for Pt L3-edge, Co K-edge and Ni K-edge, respectively. The scattering paths were obtained using the FEFF8 software.
The in situ XAS measurements were performed at the P65 beamline (DESY, Hamburg, Germany). Pristine catalyst was dispersed in 15 cm3 of N,N-dimethylformamide to achieve an overall catalyst concentration of 14 mg cm−3. Subsequently, a PBI solution in DMF was added to reach a catalyst-to-PBI mass ratio of 1
:
0.11. The resulting suspension was homogenised in an ice bath using an ultrasonic probe (Sonoplus MS73, Bandelin, Germany) operated at 12 W for 30 minutes. The ink was then deposited onto a square piece of gas diffusion layer (6 cm × 6 cm, GDL, Sigracet 38BC®, SGL, Germany) using a computer-controlled ultrasonic spray coater (CNC platform by CZRobotics, ultrasonic spray system by Cheersonic) equipped with a 0.5 mm ultrasonic tip (6 W, N2 flow 50 cm3 min−1). The GDL was placed on a hot plate preheated to 110 °C. The ink, stored in a 10 cm3 syringe, was fed to the ultrasonic spray tip via a hose using an automatic syringe pump (NE-1000, KFTechnology, Italy) at a flow rate of 0.5 cm3 min−1. Following deposition, the hot plate temperature was raised to 190 °C for 30 minutes to complete the drying process. The coated GDE was then allowed to cool on the plate and stored in a dry box under N2 atmosphere. The final catalyst loading (1 mgmetals cm−2) was verified by comparing the mass of the coated GDE to that of the pristine GDL.
For monitoring in situ degradation via XAS, a custom-designed in-house cell allowing spectra recording during cell heating was employed. Detailed cell description and photographs of the complete cell assembly are provided in the Section S1.3 and Fig. S1.1 in SI. This setup was chosen for its similarity to the H3PO4 leaching setup used in this study (i.e., simple leaching without polarisation). In future research, it would be advisable to use an in situ setup allowing electrode polarisation (for example, during the ORR process) while maintaining a temperature of 180 °C and an environment of concentrated H3PO4 (in-operando regime).
The measurements were carried out in a PTFE vessel (100 cm3) placed in a heating mantle (WiseTherm WHM, Witeg) heated to 120, 140, 160 or 180 °C using purified 97.6 wt% H3PO4 as an electrolyte solution. All measurements were performed using a potentiostat EC301 (Stanford Research Systems, UK). A rotator with rotation control (AFMSRXE, Pine Research, USA) was used in combination with an in-house made shaft. The shaft was constructed from CNC axis with centering and adjusted to a precise diameter on a lathe. The temperature of electrolyte was monitored by a thermocouple (5SRTC, Omega, USA) placed in a glass tube and controlled automatically by an in-house made temperature controller. The counter electrode was a roll of Pt foil with an area of about 7 cm2 separated from the main compartment of the electrolyte by a ceramic frit. MSE (Hg|Hg2SO4 in K2SO4sat. solution, Monokrystaly, Czechia) was used as the reference electrode. It was separated from the hot H3PO4 environment by a double junction, the first one filled with 97.6 wt% H3PO4, the second with saturated K2SO4 solution. The second junction was placed in a short Liebig cooler tempered to 25 °C by a cryostat (Julabo F12). The cell was sealed with PTFE tape (DuPont, USA).
Each prepared thin film-modified electrode was first cycled 3 times in the potential range from 0.3 V to −0.1 vs. MSE (sweep rate 100 mV s−1) in the N2 saturated electrolyte. Subsequently, the LSVs were measured in the same potential range at a sweep rate of 5 mV s−1 in the N2 saturated electrolyte (minimum saturation time 15 min) and then in O2 saturated electrolyte (minimum saturation time 15 min) at revolving RRE electrodes (400, 900, 1600 and 2500 rpm). Currents measured in N2 saturated electrolyte were subtracted from the currents obtained in the O2 saturated electrolyte. All measurements were performed at various temperatures (120 °C then 140 °C, 160 °C and finally 180 °C, whole series typically takes from 3 to 5 hours).
The potentials measured using MSE reference electrode (at each temperature) were recalculated to RHE scale. For this purpose, an in-house made RHE (Pt wire coated with Pt black immersed directly in H2 saturated H3PO4 electrolyte) was used in the same setup, without the use of additional electrodes or gas saturation. Potentials between MSE and RHE were measured over the temperature range of 120–180 °C. This dependence can be approximated by the equation in eqn (2) (R2 = 0.975 for the temperature range of 120–180 °C).
| EMSE,t = −694.6 mV + t·0.70 mV °C−1 | (2) |
| Catalyst type | Pristine (wt%) | After 24 h leaching | ||||
|---|---|---|---|---|---|---|
| In liquid fraction (wt%) | Remaining in alloy (wt%) | |||||
| Pt | Co(Ni) | Pt | Co(Ni) | Pt | Co(Ni) | |
| Pt/C | 38.1 ± 0.9 | — | 0.1 ± 0.1 | — | 31.4 ± 0.7 | — |
| Pt–Co/KB | 25.3 ± 0.4 | 1.9 ± 0.1 | 3.3 ± 0.2 | 1.8 ± 0.1 | 11.7 ± 0.1 | 0.0 ± 0.2 |
| Pt–Co/rGO | 31.0 ± 0.1 | 3.9 ± 0.1 | 0.5 ± 0.1 | 1.0 ± 0.1 | 25.4 ± 0.1 | 1.3 ± 0.1 |
| Pt–Ni/KB | 28.2 ± 1.3 | 7.1 ± 0.6 | 1.7 ± 0.2 | 3.2 ± 0.9 | 19.7 ± 2.5 | 3.8 ± 0.2 |
Pt–Co/C and Pt–Ni/C catalysts exhibit lower values of Tafel slope (Fig. S2.1 in SI) than the used benchmark material Pt/C. On the other hand, the Tafel slope of Pt–Co/rGO is slightly higher than the benchmark. All experimental alloy catalysts also possess a higher ESCA value. In particular, the ECSA in case of Pt–Co/rGO is more than double compared to Pt/C. All alloy catalysts also achieve higher j at negative overpotentials observed in the LSV curve (Fig. 2) compared to Pt/C. For example, at 0.9 V vs. RHE, the values are as follows: Pt/C: j = −13.5 A m−2, Pt–Co/KB: j = −35.7 A m−2, Pt–Co/rGO: j = −20.0 A m−2, and Pt–Ni/KB: j = −34.5 A m−2.
![]() | ||
| Fig. 2 LSV, 1600 rpm; RDE electrode, 25 °C, 10 µgmetals cm−2 of catalyst, 0.1 mol dm−3 HClO4, 20 mV s−1, corrected for background currents and uncompensated resistance. | ||
Subsequently, the same parameters were evaluated for the catalysts after 1–72 h of exposure to concentrated H3PO4 at 180 °C (leaching) by RRE. Determined parameters are summarised in Fig. 1 and numerically in SI, Table S2.1, example of LSV curves for Pt/C are show in Fig. S2.2 in SI.
The measured values show that already after the first hour of catalysts leaching in concentrated H3PO4 at elevated temperatures, the Tafel slope value increases for all catalysts. However, this value subsequently decreases with prolonged leaching time, and after 72 h, it reaches similar values for all catalysts (∼90 mV dec−1 for Pt/C, Pt–Co/KB and Pt–Ni/KB, ∼100 mV dec−1 for Pt–Co/rGO). The ECSA decreases for all studied catalysts. The decrease in ECSA may be attributed to particle growth, which leads to a reduction in the overall surface area. However, it should also be noted that some metallic particles may detach due to support corrosion; this effect was not investigated and is assumed to be negligible. Given that purified H3PO4 was used for catalyst leaching, and ECSA was determined after potential cycling in excess of HClO4, the decrease in ECSA was also unlikely result of phosphorous and phosphate species adsorption.18,44,45 Interestingly, the lowest ECSA loss of about 20% (after 72 h of leaching) was observed for Pt–Ni/KB. In other cases, the ECSA dropped to about 50% of the original value. After 72 h of leaching, the best performing catalyst is Pt–Ni/KB as can be seen from the following j values at 0.9 V vs. RHE, (Pt–Ni/KB: j = −12.5 A m−2 > Pt–Co/KB: j = −10.8 A m−2 > Pt–Co/rGO: j = −8.9 A m−2 > Pt/C j = −8.2 A m−2). Focusing on jex, the changes during leaching vary within one order of magnitude, generally in a non-monotonic manner in all cases, although some trends can be observed. For example, in the case of Pt/C, an initial increase in jex is followed by a downward trend. Conversely, Pt–Co/KB and Pt–Ni/KB exhibit a rising trend in jex throughout the degradation period, except for the final measurement at 72 h.
| Catalyst type | mol%alloy Pt | mol%alloy Co(Ni) |
|---|---|---|
| Pt/C | — | — |
| Pt–Co/KB | 80.1 ± 0.9 | 19.9 ± 0.9 |
| Pt–Co/rGO | 70.6 ± 0.4 | 29.4 ± 0.4 |
| Pt–Ni/KB | 54.4 ± 1.8 | 45.6 ± 1.8 |
Focusing on the first two columns of Table 1, the Pt content (38.1 wt%) determined in pristine Pt/C agrees well with the information provided by the manufacturer (40 wt%). In the case of alloy catalysts, the highest Pt content (31 wt%) was found in the pristine Pt–Co/rGO, the highest content of non-Pt metal (7 wt%) was present in pristine Pt–Ni/KB. Regarding the alloy composition, Pt–Co/KB exhibits a molar ratio of 4
:
1 Pt
:
Co, Pt–Co/rGO shows a 7
:
3 Pt
:
Co ratio, and in the case of Pt–Ni/KB, the molar fraction of Pt in the alloy is around 55 mol%, see Table 2. Results of catalyst leaching in 97.6 wt% H3PO4 for 24 h (columns 3–6 of Table 1) clearly show that while almost no Pt is dissolved from Pt/C, a significantly higher amounts of Pt were dissolved from the alloy catalysts. In particular, the loss of Pt decreased in the following order Pt–Co/KB (13 wt%) < Pt–Ni/KB (6 wt%) < Pt/rGO (2 wt%).
The non-Pt metals leached significantly more than Pt. Based on ICP-OES results, it appears that 45 wt% of Ni, 60 wt% of Co and nearly 99.5 wt% of Co was dissolved from Pt–Ni/KB, Pt–Co/rGO and Pt–Co/KB, respectively. These values were estimated from the ratio of non-Pt metal content before and after leaching. If the increased mass of the carbon support was taken into account by assuming that the Pt mass balance (the Pt mass in the pristine catalyst minus the Pt mass in the liquid fraction must equal the Pt mass remaining in the alloy after leaching), then the non-Pt metal losses are 28 wt% of Ni, 60 wt% of Co and nearly 99 wt% of Co was dissolved from Pt–Ni/KB, Pt–Co/rGO and Pt–Co/KB, respectively. Thus, the Pt–Ni alloy exhibit under studied condition higher stability than Pt–Co alloy. A significant difference between the dissolution of Co from Pt–Co alloy in the case of the two different carbon supports can be attributed to the stabilising effect of the rGO mitigating catalyst dealloying.63
These results were also confirmed by XRF analyses. For relative comparison, the parameter ρPt/X is introduced, defined as the ratio between the mass fraction of Pt and that of the non-platinum metal. The specific method of its determination is indicated in the subscript (ρPt/X,ICP, ρPt/X,XRF or ρPt/X,EDS). Results for XRF are summarised in Table 3 and in the Table S3.1 in SI. These data also indicate a stable alloy composition in the case of Pt–Ni/KB and relative stability in the case of Pt–Co/rGO compared to the exponential increase of ρPt/M,XRF in the case of Pt–Co/KB. Interestingly, a superior stability of Pt–Co/rGO catalyst during short-term operation of HT-PEMFC was recently reported.63 Further information about the catalysts (pristine alloy catalysts and catalysts after 24 h of leaching) was obtained by TEM in combination with EDS spot analysis, see Table 3 and Fig. S3.3–S3.5 in SI. This enabled the investigation of the compositional uniformity of alloyed nanoparticles (wt% of Pt and non-Pt metal X; ρPt/X,EDS) before and after leaching in 97.6 wt% H3PO4 environment. As can be seen, the pristine catalysts already contains particles with a quite variable composition (expressed by ρPt/X,EDS). After H3PO4 leaching (24 h) a part of metallic particles contains only Pt, confirming the loss of a significant part of non-Pt metal from the alloy. This trend was the same for Pt–Co/KB and Pt–Ni/KB. On the other hand, in the case of Pt–Co/rGO, the differences in particle composition before and after leaching were less significant, which is again in good agreement with the already discussed improved stability of this catalyst. It is also worth noting, that leaching of all catalysts does not only lead to the change in metals content, but also to the increased heterogeneity in their distribution, i.e. while there are some particles with a composition similar to that of the pristine sample, more particles containing only Pt are newly present after leaching, see column of ρPt/X,EDS and its standard deviations in Table 3.
| Catalyst type | ρPt/X,ICP | ρPt/X,XRF | ρPt/X,EDS | |
|---|---|---|---|---|
| Pt–Co/KB | Pristine | 13.2 ± 0.7 | 14.5 ± 0.2 | 12.2 ± 7.8 |
| 24 h leaching | ∞ | 27.6 ± 0.6 | 45.3 ± 49.8 | |
| Pt–Co/rGO | Pristine | 8.4 ± 0.2 | 8.1 ± 0.1 | 5.8 ± 1.2 |
| 24 h leaching | 19.0 ± 1.5 | 13.8 ± 0.2 | 12.9 ± 8.9 | |
| Pt–Ni/KB | Pristine | 4.0 ± 0.4 | 5.3 ± 0.1 | 6.3 ± 3.6 |
| 24 h leaching | 5.1 ± 0.7 | 6.5 ± 0.1 | 9.0 ± 7.2 |
The Raman spectra (Fig. S3.6. in SI) also show the different behaviour of the KB and rGO supports. A summary of D and G band intensity ratios (ID/IG) in Table 4 shows that the ratio is lower for rGO than for amorphous carbon KB support, and that the ratio does not significantly change upon leaching.
| Catalyst | Ratio ID/IG/% |
|---|---|
| Pt–Co/rGO | 0.90 |
| Pt–Co/rGO 24 h | 0.91 |
| Pt–Co/KB | 1.13 |
| Pt–Co/KB 24 h | 1.08 |
Another monitored parameter was the average size of metallic crystallites, see Fig. 3 and Table S3.2 in SI, determined by analysis of XRD patterns shown in Fig. S3.7–S3.10 in SI.
![]() | ||
| Fig. 3 Crystallite size interpreted from Pt(111) reflection of the catalyst samples before and after leaching in 97.6 wt% H3PO4 at 180 °C for 1, 3, 8, 24 and 72 h determined by XRD. | ||
In the case of the Pt–Ni/KB catalyst, the XRD pattern (Fig. 4) shows distinct reflections corresponding to a minor Pt and dominant intermetallic tetragonal PtNi phase.64–72 The stoichiometry of PtNi is in agreement with ICP-OES measurement, see Table 2. However, the overlap of Pt and PtNi reflections does not allow accurate determination of the crystallite size. Consequently, at least TEM images were used to estimate the particle sizes in the pristine and leached Pt–Ni/KB samples. While the particle sizes determined by TEM are not directly comparable to the crystallite sizes derived from the Pt(111) XRD reflections, the analysis clearly showed that there is no significant change in the overall particle size distribution, see Fig. S3.11 in SI.
For the Pt–Co/KB and Pt–Co/rGO catalysts, the situation is different. The reflections of relevant intermetallic phases (Pt3Co, PtCo, or even PtCo3) and Pt fall within the regions of the broad XRD peaks present in the diffractograms, see Fig. S3.7. It is, therefore, not possible to unambiguously distinguish whether any and which intermetallic phases are present in the samples alongside Pt or if the Pt–Co alloys are disordered solid solutions. The latter would not be unexpected, as the acid-leached Pt–Co alloys (as in our case, see Section S1.2 in SI) often exhibit a disordered structure, with the approximately three outermost atomic layers enriched in Pt.72,73 Considering the small particle sizes of investigated nanoparticles (around 3 nm) and given the atomic diameter of Pt (0.27 nm74), such Pt enrichment would correspond to more than half of the total particle volume. To facilitate comparison, the main reflection (40–50° 2θ) is interpreted as Pt(111) reflection and the corresponding crystallite sizes are shown in Fig. 3.
The results show that catalyst leaching in 97.6 wt% H3PO4 at 180 °C leads in the case of Pt/C, Pt–Co/KB to a pronounced increase in average crystallite size interpreted from Pt(111) reflection. On the other hand, nanoparticles in Pt–Co/rGO and Pt–Ni/KB do not seem to grow significantly. The observed growth can be attributed to Ostwald ripening at high open circuit potential (approximately 100 mV lower vs. estimated EORR, for suspension of catalyst with concentrated H3PO4 in contact with air, i.e. about 20% of O242,43). For the time range of 0–24 h, the most significant crystallite size growth of about 110% was observed for Pt/C. In the case of alloy catalysts, the crystallite growth was significantly lower (39% for Pt–Co/KB, and only 10% for Pt–Co/rGO). A similar behaviour has also been observed in the literature when the same catalyst was used in a single-cell setup.63 These results show that while almost no Pt leached from Pt/C, the catalyst changed significantly. This suggests that a significant part of the nanoparticles dissolved into Pt2+ ions, most of which redeposited on the larger crystallites within the catalyst. The more pronounced Pt loss and lower crystallite growth in the case of alloy catalysts can be interpreted as a greater difficulty of Pt2+ ion redeposition on distorted Pt alloy surfaces than on a pure Pt surface.51 Finally, the least pronounced crystallite growth and least Pt and Co dissolution in the case of Pt–Co/rGO confirms the stabilising effect of rGO support. Thus, from this point of view, the use of rGO as a support appears promising because a smaller crystallite size implies a higher ECSA for ORR. The insignificant growth of nanoparticles within Pt–Ni/KB can likely be attributed to the presence of intermetallic PtNi phase.
The coordination numbers of Pt and Co(or Ni) (labelled X) were determined from the EXAFS fit, see Fig. 5, XAS spectra in Fig. S3.12–S3.18 in SI. The notation CNPt–X refers to a coordination number where Pt is the central atom and X represents the (non-Pt) coordinating atom. It can be deduced from the Pt–Pt coordination number, CNPt–Pt, increase (the ratio of surface Pt atoms to bulk atoms decreases) that the particles increase in size. Another reason for the increase in value of CNPt–Pt may be the leaching of the non-Pt metal from the alloy. Based on CNPt–Pt, the observed trend in particle size growth (Pt–Ni/KB < Pt–Co/rGO < Pt–Co/KB < Pt/C) agrees with Fig. 3. In contrast, the CNX–X values are not significantly affected by the degradation. This could imply that statistically, in the nanoparticle surface layer, Ni and Co are more likely present as single atoms (which subsequently dissolve) surrounded/coordinated mainly by Pt.
On the other hand, deeper in the nanoparticles (protected from etching) the non-Pt metals are more concentrated, so that on average Co(Ni) is in direct contact with approximately 1 atom (2.5 atoms) of the same non-Pt atom. In the case of Pt–Ni/KB, the value of CNNi–Ni is relatively higher compared to Pt–Co alloys, which can be attributed to the fact that the molar concentration of Ni in the alloy is higher compared to Pt–Co alloys, see Table 2. This value can also be explained by the protecting effect that the Pt shell has on the Ni-rich core. It can be seen that, while the pristine samples have lower CNX–Pt (non-Pt central atom coordinated by Pt) values, after only 1 h of leaching this value increases and than is more or less stable. This can be explained by the fact that the surface atoms of the non-Pt metal are dissolved by leaching, leaving only the atoms of the non-Pt metal in the core protected by the Pt shell.
A similar trend was observed during the in situ measurements (see Fig. S3.19–S3.22 in SI). Fig. 6 summarises the Pt-edges for Pt/C, Pt–Co/KB and Pt–Ni/KB, and Fig. 7 shows the Co(Ni)-edges for Pt–Co/KB and Pt–Ni/KB.
In situ XAS measurements at the Pt edge show a stable trend in CNPt–Pt for the Pt/C sample at 180 °C, the coordination number remains approximately constant around 7.5, see Fig. 6A. For the Pt–Co/KB, at the Pt L3-edge, see Fig. 6B, the EXAFS fitting was restricted to Pt–Pt contributions, since the weaker Pt–Co signals could not be distinguished from the dominant Pt–Pt scattering paths due to the higher scattering amplitude of Pt. Conversely, at the Co and Ni K-edges, both X–Pt and X–X contributions could be fitted, as the strong contrast in backscattering amplitudes makes the heteroatomic coordination more clearly detectable. Pt–Ni/KB shows a stable value of CNPt–Pt (around 4) and CNPt–Ni (around 2) in agreement with ex situ degradation, see Fig. 6C. Similarly, CNCo–Co and CNCo–Pt values evaluated from Co-edge of Pt–Co/KB remain constant, see Fig. 7A. In contrast, the Ni-edge of Pt–Ni/KB, see Fig. 7B reveals a significant increase in CNNi–Ni during the 180 °C treatment. This pronounced increase suggests that the remaining Ni atoms become confined within a Ni-rich core.
Finally, Table S3.3 and Fig. S3.23–S3.25 summarises post-mortem XRD of used GDEs used for in situ XAS measurements, where growing particle sizes can be seen (of prepared GDE vs. pristine catalyst powder).
These measurements demonstrate the feasibility of using the XAS setup for in situ experiments. In the future, a similar approach could be employed to investigate catalyst degradation under identical conditions, but simultaneously with catalyst polarisation and ongoing ORR.
A representative set of LSV curves for various catalysts and temperatures of 120–180 °C (revolution rate of 2500 rpm) is shown in Fig. 8, and the corresponding Tafel plots are presented in Fig. S4.1, see SI. It is clear, that the jlim generally increases with increasing temperature.
![]() | ||
| Fig. 8 LSV, RRE using revolution rate 2500 rpm; GC electrode, 10 µgmetals cm−2, 97.6 wt% H3PO4, 5 mV s−1 corrected for background currents and uncompensated resistance, use of corresponding EORR,g,T. | ||
The pristine Pt–Co/KB showed the best ORR activity in aqueous media. It also performs best in the H3PO4 environment. Contrary, the lowest ORR activity was observed for Pt–Co/rGO. This can be attributed to an improved stability of the catalyst supported by rGO preventing more significant dealloying of the nanoparticles as discussed previously. A similar behaviour of the same catalyst has also been observed in the literature when used in a HT-PEMFC single-cell cathode.63
The obtained ORR LSV data at various temperatures were analysed to obtain the Tafel slope, charge transfer coefficient (α) and exchange current density values (jex). The results are summarised in Fig. 9 and numerically in Table S4.1, see SI.
Alloy catalysts generally exhibit a lower value of ORR Tafel slope than Pt/C. If the charge transfer coefficient α remained constant within the temperature range, the Tafel slope value would be expected to increase with temperature. In this study, the Tafel slope of all catalysts generally decreases with increasing temperature and approaches values of around −105 mV dec−1 at 180 °C, which is comparable to the value observed for ORR on Pt in concentrated H3PO4 at 150–180 °C.77,80
It should be noted that the measurement itself takes a finite amount of time, during which the catalyst can undergo changes as discussed in the previous section. Additionally, the catalysts are polarised during the experiments, which may affect the kinetics of such transformations. Nevertheless, in these particular experiments, the presence of a significant excess of H3PO4 and intensive electrode rotation likely suppress Ostwald ripening by rapidly removing dissolved Pt2+ from the diffusion layer and thus minimising redeposition.
In the case of pure Pt, a change in Tafel slope is still observed, which may stem from changes in the reaction mechanism, changes in surface energy, or temperature-induced surface restructuring. The behaviour observed for non-Pt components supports previous findings that nearly all non-Pt metals leach out from the Pt surface region where the ORR takes place. This suggests that, temperature-induced modifications (such as a decrease in phosphate coverage, restructuring of surface sites, or changes in surface energy) can significantly impact the reaction kinetics. These effects may act synergistically and could explain the observed decrease in Tafel slopes.
Overall, the convergence of Tafel slope values for all tested catalysts at elevated temperatures implies that the apparent electrochemical behaviour under these conditions is governed more by temperature-driven surface modifications than by the specific alloy composition.
From the jex values, activation energies were estimated using the methodology described in ref. 60; the values are provided in SI Table S4.2. These values are comparable with those reported in the literature.81–85 The values differ among the individual alloys; the value of the activation energy follows the increasing trend Pt < Pt–Co < Pt–Ni.
In summary, the newly developed alloy catalysts achieve Pt/C-level ORR activity at lower Pt loadings, offering a credible route to Pt-lean cathodes. These results suggest that finding suitable combination of carbonaceous support and alloy composition is a key to development of practically useful HT-PEMFC catalyst. To translate these materials’ promise into device-level durability, following experiments should focus on long-term polarisation/accelerated stress test campaigns in concentrated H3PO4 (140–180 °C), coupled with operando structural/composition analyses to relate activity to alloy composition and nanoparticle-size evolution.
| ECSA | Electrochemically active surface area |
| EDS | Energy dispersive spectroscopy |
| EXAFS | Extended X-ray absorption fine structure |
| FC | Fuel cell |
| GC | Glassy carbon |
| GDL | Gas diffusion layer |
| GO | Graphene oxide |
| HT | High temperature |
| ICP-OES | Inductively coupled plasma with optical emission spectroscopy |
| K–L | Koutecky–Levich |
| KB | Ketjen black |
| LSV | Linear sweep voltammetry |
| LT | Low temperature |
| MSE | Mercury-mercurous sulphate electrode (saturated) |
| ORR | Oxygen reduction reaction |
| PBI | Polybenzimidazole |
| PEMFC | Proton-exchange membrane fuel cell |
| PTFE | Polytetrafluorethylene |
| RDE | Rotating disk electrode |
| rGO | Reduced graphene oxide |
| RHE | Reversible hydrogen electrode |
| RRE | Rotating rod disk electrode |
| TEM | Transmission electron microscopy |
| XAS | X-ray absorption spectroscopy |
| XRD | X-ray diffraction |
| XRF | X-ray fluorescence |
| XPS | X-ray photoelectron spectroscopy |
Data for this article, including structured raw data accompanied with data description are available at Zenodo at https://doi.org/10.5281/zenodo.17131699.
| This journal is © The Royal Society of Chemistry 2026 |