Wanderson O.
Silva
*a,
Alexandre
Mabillard
a,
Mathieu
Soutrenon
a,
Grégoire
Gschwend
b,
Yorick
Ligen
c,
Steve
Joris
a,
Luc
Bondaz
d,
Kumar Varoon
Agrawal
d and
Hubert H.
Girault
ef
aInstitute of Systems Engineering, HES-SO Valais- Wallis, CH-1950 Sion, Switzerland. E-mail: wanderson.oliveiradasilva@hevs.ch
bCentre Suisse d'Électronique et de Microtechnique, Battery Innovation Hub, Rue de Monruz 17, CH-2002 Neuchâtel, Switzerland
cGREENGT SA, Rue de Lausanne 57, CH-1110 Morges, Switzerland
dLaboratory of Advanced Separations, Ecole Polytechnique Fédérale de Lausanne (EPFL), Valais Wallis, Rue de l'Industrie 17, CH-1950 Sion, Switzerland
eInstitut des Sciences et Ingénierie Chimiques (ISIC), École Polytechnique Fédérale de Lausanne (EPFL) Station 6, CH-1015 Lausanne, Switzerland
fMaterial Science, Energy and Nanoengineering (MSN) Department, University Mohammed VI Polytechnic, 43150 Ben Guerir, Morocco
First published on 28th January 2025
The present work reports a simple approach to manufacture electrocatalytically active gas diffusion electrodes (GDEs) in two steps: (i) inkjet printing and (ii) flash light irradiation from a xenon flash lamp, a process called Print-light-synthesis (PLS). Pt/C PLS GDEs were manufactured from a Pt precursor ink printed directly over a carbon paper gas diffusion layer (GDL) with a microporous layer (MPL) of carbon with a metal precursor loading of 0.5 mgPt−1 cm−2, the precursor film was then exposed to flash light irradiation at 450 V-pulse for 100 ms. SEM images showed a uniform and thin Pt catalyst layer deposited on top of the GDL. XRD and XPS spectra evidenced metallic Pt with face-centered cubic crystalline structures. TEM analysis provided an average particle size of 5.0 ± 0.3 nm with uniform particle distribution over the MPL. Electrochemical characterization was performed on half-cell and fuel cell setups showing electrocatalytic performances comparable to that of a reference GDE Pt/C. Pt/C PLS shows even better fuel cell performance per gram of Pt catalyst compared to the reference Pt/C. This work shows that PLS is a very simple approach, e.g. to manufacture GDEs on a roll-to-roll basis for applications in energy conversion devices such as fuel cells, batteries, electrolyzers, etc.
To achieve optimal platinum loading, a crucial parameter for practical applications is to find a balance between maximizing fuel cell power density and minimizing overall material costs. This balance ensures robust fuel cell performance, durability, and efficiency throughout the practical PEMFC operation.39,40 Advanced catalyst designs, along with improved electrode structures and novel materials, are continuously explored to enhance catalytic efficiency and decrease platinum usage, aiming to meet the practical demands of commercial-scale PEMFC applications.20,34 The interactions between the Nafion ionomer and Pt/C catalyst play a pivotal role in determining the structure and performance of the catalyst layers in PEMFCs. Nafion ionomers can influence few catalyst layer properties such as porosity, ionomer distribution, and proton conductivity, which are crucial factors that affect the TPB and the overall electrochemical activity and fuel cell performance.19,20 As already highlighted, Pt loading plays an important role in the number of active sites and catalytic activity, which influence directly the reaction kinetics for the HOR and ORR and efficiency.41 In addition, the carbon support where the Pt nanoparticles (Pt-NPs) are supported also affects the catalyst's stability, dispersion, and electron transfer properties.15 Even the typical dispersing solvents such as water and isopropanol used in the catalyst ink formulations can have an impact on the morphology and structure of the catalyst layer, affecting catalyst distribution and PEMFC performance.42 Therefore, optimizing all those parameters is key to achieve an ideal catalyst layer structure, maximizing active sites, proton conductivity, and mass transport, which will ultimately impact the practical performance of the fuel cell in terms of power density, durability, and efficiency.
The MEA preparation with the optimal Pt/C catalyst layer involves several steps to maximize the triple-phase boundary and one of those steps involves the catalyst layer fabrication. For the PEMFC, this step is typically done by one of these three main approaches: Catalyst Coated Substrate (CCS), Catalyst Coated Membrane (CCM), and Decal Transfer Method (DTM).22 Based on this, several ink deposition methods can be applied to create efficient and uniform catalyst layers such as: blade coating, brush printing, ultrasonic spraying, rolling, screen printing, inkjet printing, electrospinning, slot die and electrospraying.22 Here, particular attention will be given to inkjet printing since it offers several advantages compared to the other deposition methods such as: (i) a very precise control of the catalyst deposition, (ii) versatility allowing controlled deposition of catalyst layers with different loadings, surface areas, designs, catalyst layers or gradients with high deposition resolution and accuracy, (iii) low material cost since it significantly reduces the material waste by precisely placing droplets only where needed, optimizing the catalyst utilization, etc.43–46 Furthermore, the non-contact nature of inkjet printing minimizes any possible damage of the substrate and/or catalyst layer. These advantages make inkjet printing a suitable approach for catalyst deposition compared to conventional methods for GDE fabrication in PEMFCs.
The combination of inkjet printing, to deposit the catalyst precursor, with a post-printing process carried out through flash light irradiation from a xenon lamp introduces a new method called Print-light-synthesis (PLS).47,48 The key advantage of PLS is the ability to precisely control the catalyst precursor deposition while harnessing light as a trigger for the in situ catalyst synthesis. This approach also presents great flexibility, allowing for tailored catalyst patterns and fine-tuning of catalyst characteristics such as size, shape, and crystallinity.48,49 In addition, the process is expected to be faster and more eco-friendly compared to conventional high-temperature synthesis methods, reducing energy consumption and waste generation. Therefore, PLS emerges as a promising approach for efficient and customizable catalyst fabrication in PEMFCs, offering advantages in terms of control, speed, and environmental impact over traditional methods.
In this work, PLS was applied to manufacture electroactive gas diffusion electrodes, containing small Pt nanoparticles well dispersed over the carbon support. This is a very simple approach where a Pt precursor salt and citric acid as a reducing agent are initially dissolved in ethanol and then deposited by inkjet printing directly over a GDL containing a microporous layer (MPL). After drying, the precursor film was exposed to high intensity flash light irradiation from a xenon lamp. With the flash light exposure, the carbon black present in the GDL acts as light-to-heat absorber and then it heats up in a very short time (milliseconds). The citric acid is decomposed thermally and photochemically and acts as a sacrificial electron donor, which subsequently promotes the Pt salt conversion into the metallic nanoparticles. The Pt catalyst synthesized by PLS was later characterized by Inductively coupled plasma mass spectrometry (ICP-MS), X-ray diffraction (XRD), scanning electron microscopy (SEM), high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM), X-ray photoelectron spectroscopy (XPS), electrochemical measurements in a conventional and half-cell setup to analyze the oxygen reduction reaction and polarization curves in a PEM fuel cell H2/air.
SEM images in Fig. 2 show very uniform Pt nanoparticle deposition over a microporous layer, which evidences inkjet printing as a suitable deposition method that minimizes the precursor salt waste once the platinum content stay on the GDL surface, and the flash light irradiation process promotes the fast synthesis of Pt nanoparticles. Thin and uniform catalyst deposition is one of the key factors to improve PEMFC performance. Therefore, the proposed protocol can potentially maximize the triple phase boundary with the possibility to manufacture GDEs with less Pt loading keeping the same catalytic performance and then making the technology cheaper and economically viable for large-scale commercialization.
![]() | ||
Fig. 2 SEM-EDX images of the GDE cross-section of the GDL and Pt catalyst layer synthesized by print-light-synthesis. |
Fig. 3 presents the XRD diffraction patterns of the Pt/C synthesized by PLS in comparison with the bare GDL that contains only carbon black and PTFE, and commercial Pt/C. The XRD spectrum for Pt/C PLS presents characteristic peaks of the face-centered cubic crystalline structure which corroborate with the diffraction peaks of the Pt/C ref.53 All three diffraction patterns show a broad diffraction peak around 25° 2θ, corresponding to the (002) plane of carbon graphite. The diffraction peaks placed at 39.8, 46.1, 67.6 and 81.4° are ascribed to the Pt (111), (200), (220) and (311) planes,53 respectively. The Pt (111) plane is one of the most common crystallographic planes reported for Pt catalysts, and it presents a high atomic density, which is typically associated with high reactivity in catalysis for example the hydrogen oxidation reaction and oxygen reduction reaction. Therefore, the Pt (111) plane was used as a reference to estimate the crystallite size by using the Sherrer's equation: D = Kλ/βcos
θ. A crystallite size of 3.0 nm was estimated for Pt/C PLS and it is in the range comparable to the particle size by TEM.
![]() | ||
Fig. 3 X-ray diffractograms of the synthesized Pt/C PLS compared with the bare GDL and commercial Pt/C. |
TEM images of the Pt/C PLS and its corresponding histogram from the average nanoparticle size distribution are shown in Fig. 4. The acquired TEM images show uniform nanoparticle distribution over the carbon support with sphere shape and average nanoparticle size mostly in the range of 2–6 nm. The TEM images were recorded from different areas as illustrated in Fig. S6† and evidence the absence of large nanoparticle agglomerations; therefore the present PLS method provides a satisfactory nanoparticle distribution and homogeneity. Based on these TEM images, 600 nanoparticles were selected randomly and measured individually by using ImageJ software and all those values were used to construct a histogram providing an average nanoparticle size of 5.0 ± 0.3 nm. In addition, the histogram shows that 0.5% of the Pt nanoparticles are smaller than 2 nm, and 79.2% of the nanoparticles have an average size in the range of 4–6 nm, followed by 16.8% in the range of 8–10 nm, and 2.3% in the range of 12–14 nm and only 1.2% of those nanoparticles have a particle size larger than 14 nm. Therefore, the controlled Pt salt deposition and distribution performed by inkjet printing, followed by the fast nanoparticle synthesis process carried out in 100 ms using a high-power light source delivered by a xenon lamp, not only generate small nanoparticles but also minimize the nanoparticle agglomeration, which could be attributed to the control of the particle nucleation and growth from the fast-heating and cooling process.54
![]() | ||
Fig. 4 Transmission electron microscopy images and average particle size distribution histogram for Pt/C PLS (Pt loading: 0.25 mgPt−1 cm−2). |
Chemical synthesis is one of the main protocols to synthesize Pt nanoparticles.55 This method normally uses a water-soluble Pt cation as a precursor, a reducing agent and caping agent, the last one is used to stop the nanoparticle growth and agglomeration, both agents act to control the particle size and shape, for example some colloidal methods can generate Pt particles with different shapes such as sphere, cube, cuboctahedra, octahedra, triangular plates, hexagonal plates, nanobars, etc.11,56 However, several caping agents have been applied to minimize the particles aggregation, such as poly-vinylpyrrolidone (PVP), poly vinyl alcohol (PVA), cetyltrimethylammonium bromide (CTAB), dodecylammonium bromide (DTAB), etc.56 However, on the other hand those caping agents also take place blocking active catalytic sites which limits the catalyst performance. Therefore, several post-treatment steps and processes are required to remove those caping agents, which is high time consuming and costly. Therefore, the PLS protocol proposed in this study can also bring new insights and advantages as a free-capping agent protocol to synthesize small Pt nanoparticles with uniform dispersion.
X-ray photoelectron spectra (XPS) of Pt from Pt/C PLS and Pt/C ref catalysts are illustrated in Fig. 5. Both Pt 4f spectra were calibrated using the adventitious C 1s peak at 284.8 eV. The high-resolution spectra related to the Pt 4f region can be deconvoluted into 2 doublets, with components Pt 4f7/2 and Pt 4f5/2 separated by 3.35 eV. The main deconvoluted peaks for both catalysts centered at 71.3 and 74.6 for Pt 4f7/2 and Pt 4f5/2, respectively, can be attributed to the metallic platinum Pt0, which corresponds to 92.7% for Pt/C ref and 82.0% for Pt/C PLS. Overall, both catalysts present similar chemical states. The deconvoluted peaks placed at 72.7 and 76.0 for Pt 4f7/2 and Pt 4f5/2, respectively, are ascribed to the chemical state +2 which is related to the PtO or Pt(OH)2 species. Therefore, Pt/C synthesized by PLS provided 18% of platinum oxides, which is almost three times higher compared to the Pt/C ref that presents only 6.3%. It is well known that a higher Pt0 fraction and lower Pt oxygenate species present a weaker oxyphilicity and then higher electrocatalytic performance for the oxygen reduction reaction, which can explain the slight difference in the electrochemical performance as will be discussed later.57 In addition, it is important to point out the absence of the Cl 2p peak (Fig. S7†) from metal chlorine for the Pt/C PLS catalyst, typically detected at 198.5–199 eV, which evidences a satisfactory Pt salt conversion into Pt metallic, and the remaining salt was then removed by washing several times with water and isopropanol.
Therefore, the Pt 4f spectrum for Pt/C PLS is related to the synthesized amount of Pt provided by flash light irradiation. The higher Pt oxide content was also observed by Cha et al.54 who synthesized high-entropy alloys by flash-thermal shock (FTS) also using a xenon lamp source, this similar approach was followed for carbon nanofiber (CNF) membranes impregnated with different metal precursors (Pt, Ir, Fe, Ni, Co, La, Ce, In, and Sr), which were exposed to flashlight irradiation for 20 ms and an energy density of 4.9 J cm−2. A similar approach was then applied in this work, however with the advantages to control the metal precursor loading deposition and layer thickness by inkjet printing. This evidences PLS as a suitable strategy to synthesize GDEs on a large scale for several energy storage applications such as fuel cells, batteries, electrolyzers, etc.
Catalyst activity of the Pt/C PLS was initially investigated by cyclic voltammetry (CV) from a standard three electrode electrochemical cell in HClO4 2 M electrolyte saturated with argon and a scan rate of 100 mV s−1 as shown in Fig. 6a. The CV profile presents three characteristic peaks related to the hydrogen adsorption–desorption region from 0.05 to 0.35 V (vs. RHE), double-layer region 0.35 to 0.8 V (vs. RHE) and Pt-oxide formation/reduction from 0.8 to 1.2 V (vs. RHE).11 The CV from Pt/C PLS presents a wider double layer region compared to Pt/C ref, which can be ascribed to the higher capacitive behavior, probably related to the higher surface area provided by the higher carbon support content.
The electrochemically active surface areas (ECSA) for both catalysts were measured by CO stripping following the equation:
ECSA (m2 gPt−1) = QCO/QML × mPt |
STEM-HAADF images presented in Fig. 7 show Pt nanoparticles dispersed on the carbon support (Fig. 7a), which is confirmed by STEM-EDXS Pt elemental mapping as illustrated in Fig. 7b. In addition, Pt single atoms dispersed on the carbon support are clearly observed in Fig. 7e and f. Therefore, this evidences Print-light-synthesis as a promising approach to successfully convert the Pt precursor into small nanoparticles and single dispersed atoms from two single steps: (I) inkjet printing and (ii) flash light irradiation without complex and timing consuming steps. In addition, the presence of single Pt single atoms could be indirectly correlated with the higher electrochemically active surface area for Pt/C synthesized by PLS compared to the reference Pt/C.
Electrochemical activity for the ORR catalyzed by Pt/C PLS in comparison with Pt/C ref and bare GDL was measured under more realistic conditions using a half-cell setup. This kind of setup (Fig. S4†) has already been reported50 and can predict more accurately the catalyst performance of a fuel cell when compared to the standard setup in a three-electrode cell using a rotating disk electrode (RDE) and oxygen-saturated electrolyte. Polarization curves for the ORR were hence recorded using linear sweep voltammetry (LSV) in HClO4 2 M/air at 5 mV s−1, as presented in Fig. 6b. Both polarization curves present a profile with the oxygen reduction reaction starting at 0.9 V (vs. RHE), and the slight difference can be attributed to the higher capacitive current from the higher carbon content for Pt/C PLS. Fig. S8† shows a typical Tafel plot curves for the ORR catalyzed by the Pt catalyst; both electrodes present similar Tafel slopes which confirm similar kinetics and rate determining steps for the ORR, however Pt/C ref has a slight lower overpotential (30 mV) compared to the Pt/C PLS, which might be attributed to the higher Pt loading and lower capacitive current. However, overall, the electrochemical activity of Pt/C PLS for the ORR is comparable to that of the commercial Pt/C catalyst.
Polarization curves from a PEMFC H2/air with Pt/C PLS and Pt/C commercial are shown in Fig. 8. The fuel cell tests were performed at 80 °C with humidified H2 and air gases and flow rates of 300 and 600 mL min−1, respectively, back pressure of 2 bar and 100% relative humidity. During the fuel cell measurements, both the anode/cathode inlet and outlet were heated at 100 °C to avoid water condensation and the liquid water excess after the outlets were removed by using a water trap. The Pt/C ref MEA was prepared with a Pt loading of 0.5 mgPt−1 cm−2 on the anode and cathode sides. Similarly, the Pt/C PLS MEA was also prepared using a theoretical Pt precursor loading of 0.5 mgPt−1 cm−2 on both electrodes, however taking into account 50% of conversion from the Pt precursor salt into Pt metallic nanoparticles after PLS synthesis, the final Pt loading from Pt/C PLS MEA was 0.25 mgPt−1 cm−2. The comparison between Pt/C PLS and Pt/C ref was made considering both values, the nominal Pt loading deposited by printing and the effective Pt catalyst loading obtained after the synthesis measured by ICP-MS, which take place in the HOR and ORR reactions. To evaluate the practical catalyst performance, polarization curves were recorded by measuring the output cell voltage from different applied current densities from 0.08 to 1 A cm−2 as illustrated in Fig. 8. According to these polarization curves, Pt/C PLS presented fuel cell performance close to reference MEA Pt/C (0.5 mgPt−1 cm−2). At 0.9 A cm−2 the maximum power densities obtained for Pt/C PLS and Pt/C ref were 0.42 and 0.45 W cm−2, respectively; the slightly higher power density for Pt/C ref of 7% is likely related to the higher catalyst loading with 0.5mgPt−1 cm−2 compared to Pt/C PLS that presents half of that (0.25 mgPt−1 cm−2). The power density increases linearly until around 0.9–1 A cm−2, where a plateau is clearly observed, before decreasing at higher current densities (>1 A cm−2) due to the higher resistance associated with the mass transport limitation.
When both polarization curves are corrected by the effective Pt loading present in the gas diffusion electrodes, 0.5 mgPt−1 cm−2 for Pt/C ref and 0.25 mgPt−1 cm−2 for Pt/C PLS considering a conversion rate of 50% from the Pt salt to Pt nanoparticles obtained after PLS synthesis, the Pt/C PLS catalyst presents better cell performance or Pt utilization per gram of catalyst, in particular at higher current density (0.9–1 A cm−2) with two times higher current density per of Pt compared to the commercial Pt/C (Fig. 9). Therefore, while the PLS process requires further optimization to enhance the conversion efficiency from metal precursors to metal catalysts, it shows its suitability for producing GDEs. This strategy allows for the controlled and uniform deposition of catalyst layers, with well-dispersed nanoparticles over the carbon support of the gas diffusion layer. This results in a significant improvement in Pt utilization efficiency per unit of generated power density. Thus, print-light-synthesis has several advantages such as the fast and low-cost process and scalability since inkjet and flash lamps can be used to print and flash on areas from mm2 to m2 catalysts and layers/gradients, etc., allowing a large-scale production of GDEs with applications in several energy conversion devices such as fuel cells, batteries, electrolyzers, etc.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta04837g |
This journal is © The Royal Society of Chemistry 2025 |