Download PDF
Article  |  Open Access  |  23 Jan 2025

Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation

Views: 149 |  Downloads: 46 |  Cited:  0
Energy Mater. 2025, 5, 500038.
10.20517/energymater.2024.91 |  © The Author(s) 2025.
Author Information
Article Notes
Cite This Article

Abstract

The rational design of Pd-based catalysts to enhance their applications in ethanol oxidation reaction (EOR) presents both exciting opportunities and significant challenges. Herein, a series of carbon-supported PdSn nanoparticle catalysts (PdSn/C-X, X = 0.1, 0.5, 1, 2) with tunable lattice strains were synthesized using a facile method at room temperature and applied to the EOR. Our findings demonstrate that the activity and stability of EOR can be modulated by manipulating the lattice strain in Pd-based catalysts. Remarkably, PdSn/C-1 exhibits an excellent mass current density of 8,452.3 mA/mgPd, which is higher than that of most Pd-based catalysts, along with great stability, maintaining a mass activity of 573.9 mA/mgPd after 5,000 s. By combining structural analysis, in situ spectral characterization, and theoretical calculation, we elucidate that the optimal tensile strain adjusted by Sn composition in PdSn/C optimizes the free energy of the key intermediate (*CH2CO) during EOR, thereby favoring the C1 pathway and enhancing catalytic activity. This study demonstrates that by controlling the composition, the lattice strain can be altered to improve catalytic performance of Pd-based catalysts in EOR.

Keywords

Pd-based catalysts, composition regulation, lattice strain, ethanol oxidation reaction, C1 pathway

INTRODUCTION

In recent years, direct ethanol fuel cells (DEFCs) have garnered significant attention due to their non-toxicity, high energy density (8 kWh/kg), and abundant ethanol sources[1-3]. However, the slow reaction kinetics and incomplete oxidation behavior of anodic ethanol oxidation reactions (EOR) hinder the commercialization of DEFCs[4,5]. EOR involves two pathways: complete oxidation (C1 pathway, involving 12 electrons transfer) and partial oxidation (C2 pathway, involving 2 or 4 electrons transfer)[6-10]. The C1 pathway offers a much higher energy density but is impeded by the high energy barrier associated with C-C bond cleavage and the slow desorption of CO-like species during oxidation[11]. Consequently, the C2 pathway predominates in practical applications, leading to poor DEFC efficiency[12]. Thus, developing high-efficiency catalysts that promote the C1 pathway by enhancing C-C bond cleavage and reducing CO poisoning during EOR is crucial for advancing DEFC technology.

To date, Rh- and Ir-based catalysts are known to favor the C2 pathway, but their scarcity, high costs, and low activity and selectivity for EOR limit their commercialization. Alternatively, Pd-based catalysts have shown promising performance due to their excellent catalytic activity[13,14]. However, Pd-based catalysts tend to favor the C2 pathway over the C1 pathway[15,16]. Surface strain engineering on Pd-based catalysts has emerged as an effective method to address this issue by adjusting the d-band properties[17-20], thus regulating the adsorption and desorption of key intermediates and shifting the balance toward the more efficient C1 pathway[8,21,22].

Research indicates that the tensile strain in Pd enhances EOR, while the compressive strain exerts the opposite effect[23]. A common strategy to create tensile strain involves preparing a core-shell structure, such as phosphorus-doped Ag@Pd catalysts, which show 19% selectivity for the C1 pathway due to the tensile strain from the differing lattice constants of Pd and Ag[11]. However, this strategy is limited by a complex multistep preparation process, making it difficult to simply regulate tensile strain. Additionally, Pd active sites are easily poisoned by CO-like intermediates, as they struggle to activate OH species, which are crucial for CO oxidation removal[11]. Previous studies have shown that doping Pd-based catalysts with oxyphilic metals enhances the removal of CO-like intermediates by facilitating the generation of OH species[24,25]. Thus, introducing oxyphilic atoms into Pd catalysts to regulate tensile strain and remove CO-like intermediates is a promising but underexplored approach.

Herein, we present our research on lattice strain-regulated activity of carbon-supported PdSn nanoparticles (NPs) catalysts for EOR, using a simple wet chemical method[26]. By adjusting the Sn content in PdSn NPs, we achieved various lattice strains in PdSn/C. PdSn/C-1, with an optimal tensile strain of 3.7%, demonstrated remarkable catalytic activity and enhanced stability compared to other Pd-based catalysts. The incorporation of oxyphilic Sn provides effective sites for the removal of CO-like species, improving the tolerance to CO-like intermediates. In situ attenuated total reflectance surface-enhanced infrared spectroscopy (ATR-SEIRAS) and density functional theory (DFT) calculations revealed that appropriate tensile strain in PdSn NPs significantly improves electrocatalytic activity by tuning the adsorption properties of key reaction intermediates (*CH2CO) for further C-C bond cleavage. Our work introduces a novel strategy to modulate tensile strain in Pd-based catalysts for EOR through composition regulation and offers a comprehensive mechanism for EOR enhancement.

EXPERIMENTAL

Materials and reagents

Palladium chloride (PdCl2) was obtained from Energy Chemistry. Stannous chloride (SnCl2, 99.9%). Sodium citrate (Na3C6H5O7·2H2O, 99.9%), ethanol (CH3CH2OH, 99.7%), sodium borohydride (NaBH4, 99.9%), and potassium hydroxide (KOH) were provided by Shanghai Sinopharm Chemical Reagent Co., Ltd. Activated carbon (Vulcan XC-72) was sourced from Shanghai Macklin Biochemical Co., Ltd. Nafion solution (5 wt% in propanol and water) was purchased from Alfa Aesar. All reagents were of analytical grade and used without further purification. H2O (18 MΩ) was purified using an ultra-pure purification system (HHitech, Eco-S15UF).

Preparation of PdSn/C catalysts

PdCl2 (88 mg) was dissolved in 4.4 mL of water under magnetic stirring. Three drops of hydrochloric acid were added to adjust the acidity, and the mixture was left to stand for 2 h until complete dissolution was achieved. Then, 0.0948 g (0.5 mmol) of SnCl2 and 2.358 g (8 mmol) of sodium citrate were added. Sodium citrate was used as a complexing agent to coordinate with Pd and Sn ions for uniform dispersion. The solution was diluted with 200 mL of water, followed by the addition of 200 mg of XC-72 with sonication for 30 min. Next, 50 mL of 0.1 M NaBH4 was used as a reductant and injected into the mixture, followed by vigorous stirring at 25 °C for 8 h. The resulting suspension was centrifuged at 10,000 rpm for 5 min, and the supernatant was discarded. The sediment was washed with water and centrifuged three times. The collected product was dried in a vacuum oven at 60 °C for 5 h, resulting in the formation of PdSn/C-1. By varying the amounts of SnCl2 (0.0316, 0.0474, and 0.1896 g), different PdSn/C catalysts were obtained, labeled as PdSn/C-0.1, 0.5, and 2, respectively. Additionally, as a control material, Pd/C was synthesized using the same procedure without adding SnCl2.

Materials characterizations

Transmission electron microscopy (TEM, JEM-F200) and energy-dispersive X-ray spectroscopy (EDS, FEI Talos F200X) were employed to determine the morphology and chemical composition of the materials. X-ray powder diffraction (XRD) was performed using an X’Pert PRO MPD diffractometer with nickel-filtered Cu Kα radiation from Panalytical B.V. The Pd and Sn contents were measured using an inductively coupled plasma optical emission spectrometer (ICP-OES, Agilent 7700). The surface electronic states of the materials were identified using an X-ray photoelectron spectroscopy instrument (XPS, Thermo Scientific K-Alpha).

Electrochemical measurements

Electrochemical measurements were conducted using a three-electrode setup with an electrochemical workstation (AMETEK PARSTAT 4000A, Princeton). A glassy carbon electrode (GCE) with an internal area of 0.07068 cm2 served as the working electrode. Platinum foil (2 cm × 2 cm) and saturated calomel electrode (SCE) were used as the counter and reference electrodes, respectively. All experiments were conducted at room temperature, and all applied potentials were referenced to the reversible hydrogen electrode (RHE) based on Nernst equation (ERHE = ESCE + 0.244 + 0.0591 × pH), unless stated otherwise. For the preparation of ink solution, 1.4 mg of catalysts were dissolved in 1.4 mL of co-solvents (ethanol:water = 1:3, v/v) and sonicated for 30 min. Then, 28 μL of Nafion was added to the mixture, followed by additional sonication for 30 min. A working electrode with a loading amount of 200 μg/cm2 was prepared by applying 5 μL of the ink solution onto the GCE and drying it at room temperature.

Electrochemical measurements were conducted between 0.27 and 1.47 V at a scan rate of 50 mV/s in N2-saturated 1.0 M KOH or 1.0 M KOH + 1.0 M ethanol electrolyte. The electrochemically active surface area (ECSA) was calculated from the cyclic voltammetry (CV) curves in the PdO reduction peak region using ECSA = QPdO/(0.405 mC/cm2 × mPd), where QPdO is the charge obtained by integrating the reduction peak area from PdO to Pd, 0.405 mC/cm2 represents the charge required for the reduction of the monolayer PdO, and mPd is the mass of Pd on the working electrode (based on ICP-OES analysis, Supplementary Table 1). CO stripping experiments were conducted in N2-saturated 1.0 M KOH solution after bubbling CO for 20 min at a potential of 0.27 V. To remove dissolved CO, high-purity N2 was bubbled into the CO-saturated electrolyte for 20 min. CO stripping curves were recorded at a scan rate of 50 mV/s between 0.27 and 1.47 V.

In situ ATR-SEIRAS characterization

A three-electrode cell was employed for all SEIRAS tests, with a GCE, Ag/AgCl, and carbon rod as the working, reference, and counter electrodes, respectively. EOR measurements were conducted in an electrolyte of 1.0 M KOH + 1.0 M ethanol under N2-saturated solution. The background spectrum was collected at 0.5 V while the potential was swept from 0 to 1.3 V.

Theoretical calculation methods

The DFT calculations were performed using the Vienna Ab Initio Simulation Package (VASP 6.3.0)[27]. The Perdew-Burke-Ernzerhof (PBE) functional was employed within the generalized gradient approximation (GGA) framework to compute the electronic exchange-correlation energy[28]. The projector augmented wave (PAW) method was used to describe the ion-electron interactions[29]. The structure was fully optimized using a conjugate gradient optimization technique, with a cutoff energy of 450 eV for the plane-wave basis set. A 3 × 3 × 1 Monkhorst-Pack grid was employed during structural optimization. The maximal force on each atom was set to be less than 0.02 eV/Å, with an energy change convergence criterion of less than 1 × 10-5 eV. The DFT-D3 semi-empirical correction was applied using Grimme's scheme. The Gibbs free energy change (ΔG) was defined for each elementary step as ΔG = ΔE + ΔZPE - TΔS, where ΔE is the adsorption energy from DFT calculations, ΔZPE is the zero-point energy correction, T is the temperature, and ΔS is the entropy change.

RESULTS AND DISCUSSION

Material synthesis and characterization

We synthesized a series of PdSn NPs supported on activated carbon, labeled as PdSn/C-X (X = 0.1, 0.5, 1, 2, representing the Sn/Pd molar ratio in the precursor), using a straightforward co-reduction process at room temperature [Figure 1A]. TEM images [Figure 1B-F] reveal that the PdSn NPs (indicated by the white dotted circles) are uniformly distributed on the carbon support without aggregation. Additionally, statistical analysis of 100 well-distributed Pd and PdSn NPs indicates an average particle size ranging from 3.0 to 4.0 nm [Supplementary Figure 1]. EDS mappings [Supplementary Figure 2] show that the Pd and Sn elements are well distributed within the NPs, confirming the bimetallic structure of the PdSn/C-X catalysts. As measured by means of ICP-OES [Supplementary Table 1], the results confirm that the Sn contents in PdSn/C-0.1, 0.5, 1, and 2 are 1.17, 1.26, 2.11, and 3.59 wt%, respectively, verifying the tunable composition of the PdSn/C catalysts.

Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation

Figure 1. (A) Schematic illustration of preparing PdSn/C-X. TEM images of (B) Pd/C, (C) PdSn/C-0.1, (D) PdSn/C-0.5, (E) PdSn/C-1, and (F) PdSn/C-2.

To gain further insight into the surface structure of PdSn/C-X, we employed aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and XRD. HAADF-STEM images at the atomic level [Figure 2A] show the lattice spacing of PdSn NPs, which were measured from the cyan line region covering ten planes. The lattice distances of the Pd/C and PdSn/C-0.1, 0.5, 1, and 2 catalysts were measured to be 2.16, 2.17, 2.19, 2.24, and 2.28 Å, respectively, corresponding to the (111) lattice plane of Pd[24]. Additionally, a lattice spacing of 2.91 Å was observed in PdSn/C-2, indicating the presence of the PdSn (111) phase in this sample[30]. It is worth emphasizing that the observed increase in the interplanar spacing indicates the presence of tensile strain resulting from the incorporation of Sn (which has a larger parameter compared to Pd) into the Pd lattice[31], as will be further supported by the XRD investigation. Furthermore, we defined the tensile strain as (dp-ds)/ds, where the dp represents the interplanar spacing of the doped sample, and ds indicates the interplanar spacing of the pure reference sample[32]. We therefore approximated the tensile strains of 0.46%, 1.39%, 3.7%, and 5.6% for the surfaces of PdSn/C-0.1, 0.5, 1, and 2 [Figure 2B], respectively, by comparing the lattice spacing of (111) plane in PdSn/C-X (dp) with that on Pd/C (ds). In addition, the powder XRD patterns of the PdSn/C catalysts are displayed in Figure 2C. All the catalysts exhibit a typical Pd face-centered cubic (fcc) crystal structure of Pd (PDF#46-1043), with peaks around 40.1°, 46.7°, and 68.1°, corresponding to the (111), (200), and (311) reflections, respectively[24]. The broad peak around 25° represents the (002) reflection of carbon from XC-72 carbon support. Notably, the enlarged image in Figure 2C shows a negative shift in the diffraction peak from 40.1° to 39.6° as the Sn contents increase, indicating lattice expansion due to the larger Sn lattice parameter compared to Pd[30,31,33,34]. This phenomenon is consistent with the HAADF-STEM results. Moreover, the Sn-related peaks in the XRD pattern of the PdSn/C-2 [Figure 2C] are clearly visible, while they are absent in the other samples. This observation can be attributed to the higher Sn content in PdSn/C-2 compared to the other samples, which allows the Sn-related peaks to be more readily detected by XRD. The phase equilibria of the Pd-Sn system are quite complex, with numerous intermediate phases[35]. The alloy formed with a Pd fcc structure can accommodate up to 17% of Sn in the Pd-rich region of the diagram[36,37], leading to the formation of intermetallic phases such as PdSn, Pd2Sn, PdSn2, or other mixtures. According to previous works[38], the prominent peaks at 30.8° and 32.3° are assigned to the PdSn (111) and (012) (PDF#04-0803), respectively, indicating the presence of the PdSn phase at high Sn doping in PdSn/C-2[30]. Additionally, the distinct peaks at 33.9°, 42.5°, and 51.8°, which correspond to the (101), (210), and (211) reflections of SnO2 (PDF#72-1147)[39], respectively, were observed on PdSn/C-2, suggesting the formation of SnO2 phase. As shown in Figure 2D, the Rietveld refinement of the PdSn/C-1 crystal structure (as an example) suggests that a Sn atom substitutes for a Pd atom in the Sn-doped Pd-based catalyst[40]. As expected, these findings indicate that the strain state of PdSn/C-X can be tuned by adjusting the Sn content.

Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation

Figure 2. (A) HAADF-STEM images, (B) Lattice stretching ratios, and (C) XRD patterns of Pd/C and PdSn/C-X. (D) XRD profiles and Rietveld refinement fit for PdSn/C-1.

Recent studies suggest that the highest occupied electronic states up to the Fermi level serve as electronic descriptors for electrocatalytic activity[41], as they influence the adsorption and desorption characteristics of reactants and intermediates. Therefore, we conducted density of states (DOS) simulations for the active Pd center using DFT (see Supplementary Material) to understand how strained structures affect surface adsorption behavior [Supplementary Figure 3A]. The Pd d-band center shows a strong correlation with tensile strain [Supplementary Figure 3B], and its value decreases toward the Fermi level as tensile strain increases. According to the d-band theory of metal-adsorbate interactions, this change in d-band electron density has significant catalytic implications[42]. A higher local average of the d-band electron energies (ɛd) near the Fermi level results in a stronger binding strength between the active site and adsorbates[43]. However, if ɛd is too low, the reactants struggle to adsorb, while a high ɛd causes strong adsorption of intermediates, leading to catalyst poisoning[44]. Therefore, an optimal ɛd level accelerates the absorption of reaction intermediates at the active site and promotes the subsequent oxidation phase during the EOR, thereby enhancing catalytic performance.

The surface chemical compositions of PdSn/C catalysts were further analyzed by XPS, using the C1s peak at 284.8 eV for calibration. The survey spectra [Supplementary Figure 4A] confirm the presence of Pd, Sn, and C. High-resolution spectra of Pd 3d for PdSn/C-0.1 to 2 [Supplementary Figure 4B] were deconvoluted into four peaks around 335.9, 337.4, 341.7, and 344.9 eV, corresponding to Pd(0)3d5/2, Pd(II)3d5/2, Pd(0)3d3/2, and Pd(II)3d3/2[11], respectively. The high-resolution spectra of Sn 3d [Supplementary Figure 4C] show four peaks around 485.6, 493.9, 487.5, and 495.8 eV, corresponding to Sn(0)3d5/2 Sn(IV)3d5/2, Sn(0)3d3/2, and Sn(IV)3d3/2, respectively. Notably, the Pd(0)3d5/2 peak shifts to lower binding energy with increasing Sn content, indicating significant electron transfer to Pd. In addition, the proportions of Pd(0) for PdSn/C-0.1, 0.5, 1, and 2 were calculated to be 67.3%, 43.1%, 40.9%, and 46.3%, respectively, based on the peak areas of Pd(0) in Supplementary Figure 4B. The higher proportions of Pd(0) at high Sn content (PdSn/C-2) owes to SnO2 formation, which hinders Pd oxidation and increases the Pd(0) proportion, demonstrating that Sn incorporation influences Pd valence states due to its oxyphilic properties[31]. Supplementary Figure 4D shows the high-resolution spectra of the O1s region on PdSn/C. The fitted peaks at 532.2, 533.7, and 535.7 eV can be attributed to lattice O from SnOx, OH- species, and O from adsorbed H2O[45,46], respectively. In particular, the proportions of OH- on PdSn/C-0.1, 0.2, 1, and 2 were calculated to be 16.9%, 26.2%, 36.9%, and 37.2%, demonstrating the increased Sn in PdSn/C promotes OH- species formation which enhances resistance to CO poisoning. These findings, combined with strain effects, indicate that Sn content modulates tensile strain and chemical states on the surfaces of PdSn/C-X catalysts, providing a basis for enhancing EOR.

Catalytic performance of PdSn/C for EOR

The electrocatalytic performance of the as-prepared PdSn/C catalysts was thoroughly evaluated. Before EOR testing, the CV curves of PdSn/C and Pd/C catalysts were recorded in a N2-saturated 1.0 M KOH solution [Figure 3A]. The observed cathodic peak around 0.7 V during the backward scans is attributed to the reduction of the Pd oxide layer. Notably, the cathodic peaks of PdSn/C-0.1, 0.5, 1, and 2 are at 0.70, 0.69, 0.71, and 0.73 V (inset in Figure 3A), indicating a more positive peak potential compared to Pd/C (0.68 V). This shift implies the desorption of oxygenated species from their surfaces[15]. Based on the reduction peak area and the Pd content on the working electrode, the ECSAs of PdSn/C-0.1, 0.5, 1, 2, and Pd/C were calculated to be 38.6, 48.5, 104.3, 72.3, and 4.0 m2/gPd [Supplementary Figure 5A], respectively. The ECSA of PdSn/C initially grows and then decreases as the Sn content increases. The initial increase is attributed to the enhanced lattice expansion in PdSn/C-0.1 to 1, which activates additional Pd sites. However, the presence of the intermetallic PdSn phase reduces the number of active Pd sites[31]. The electrocatalytic activities of these catalysts for EOR were assessed through CVs recorded in 1.0 M KOH + 1.0 M ethanol solution. Figure 3B shows two typical oxidation peaks in the forward and backward scans, representing the oxidation of ethanol and carbonaceous intermediates, respectively. As shown in Figure 3C, to compare the activities, mass and specific activities were determined by normalizing the peak current densities of ethanol oxidation to the mass of Pd and the ECSAs, respectively. PdSn/C-1 exhibits significantly higher activity (8,452.3 mA/mgPd or 8.1 mA/cm2) than PdSn/C-0.1 (2,212.6 mA/mgPd or 3.7 mA/cm2), PdSn/C-0.5 (2,536.3 mA/mgPd or 5.7 mA/cm2), PdSn/C-2 (3,031.9 mA/mgPd or 4.1 mA/cm2), and Pd/C (246.8 mA/mgPd or 4.6 mA/cm2), outperforming most of Pd-based catalysts reported in previous studies [Supplementary Table 2].Figure 3C illustrates a clear “volcano” trend in activity with varying Sn/Pd ratios, reflecting the trend observed in ECSAs. This result indicates that incorporating Sn into Pd-based catalysts significantly improves the catalytic activity by altering the Sn/Pd ratio.

Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation

Figure 3. CV curves of Pd/C and PdSn/C-X in (A) 1.0 M KOH solution (inset: the magnified view of reduction peaks) and (B) 1.0 M KOH with 1.0 M ethanol solution. (C) Summarized mass activities (left) and specific activities (right) of Pd/C and PdSn/C-X. (D) i-t measurements at -0.24 V of Pd/C and PdSn/C-X in 1.0 M KOH + 1.0 M ethanol solution.

Furthermore, the catalytic kinetics were investigated using linear sweep voltammetry (LSV) and electrochemical impedance spectroscopy (EIS) techniques. As shown in Supplementary Figure 5B, PdSn/C-1 exhibits a significantly lower onset potential (0.63 V) compared to other catalysts (inset of Supplementary Figure 5B), indicating enhanced kinetics for ethanol oxidation. The Tafel slopes increase in the following order: PdSn/C-1 < PdSn/C-2 < PdSn/C-0.5 < PdSn/C-0.1 < Pd/C [Supplementary Figure 5C]. The smallest Tafel slope (175.02 mV dec-1) for PdSn/C-1 suggests the lowest overpotential and fastest reaction kinetics in EOR. Moreover, the EIS results [Supplementary Figure 5D] reveal that PdSn/C-1 has the smallest diameter of the Nyquist plots, demonstrating the lowest charge transfer resistance (Rct, Rct = 83 Ω; see details in Supplementary Table 3) and enhanced reaction kinetics for EOR due to the optimized Sn doping. To assess the long-term catalytic durability of PdSn/C catalysts, chronoamperometry (i-t) tests were conducted at 0.92 Vvs. RHE for 5,000 s in an N2-saturated 1.0 M KOH solution containing 1.0 M ethanol [Figure 3D]. PdSn/C-1 exhibits the slowest decay in current density, with a mass activity of 573.9 mA/mgPd at 5,000 s, outperforming the other catalysts and indicating excellent electrocatalytic durability. Additionally, the stability of PdSn/C-1 was evaluated through extended CV sweeps for 1,500 cycles [Supplementary Figure 6A]. The current density of the forward peaks from the CVs is depicted in Supplementary Figure 6B. Initially, the peak current density increased then decreased with the number of scans, consistent with previous studies[12,47]. Even after 1,500 cycles, the peak current density maintains a mass activity of 1,600 mA/mgPd, which is much higher than that of Pd/C, indicating good durability. The observed decrease in mass activity during the cycles may be attributed to the partial detachment of the catalyst from the GCE [Supplementary Figure 6C]. Further characterization of PdSn/C-1 was conducted after 1,500 cycles. As shown in Supplementary Figure 7, PdSn/C-1 exhibits a similar particle distribution, size, composition, and crystal structure compared to its state before EOR, confirming its structural stability during the EOR. Overall, the superior performance of PdSn/C-1 is attributed to Sn doping, which modulates the tensile strain and optimizes the adsorption energy of key intermediates, facilitating EOR.

Finally, electrochemical CO stripping tests were performed on the PdSn/C catalysts to evaluate their capability for removing carbonaceous intermediates. CO stripping voltammograms were obtained by applying a voltage of 0.27 V to the working electrode for 20 min in a CO-saturated 1.0 M KOH electrolyte solution. As shown in Supplementary Figure 8, peaks at 0.88 and 1.08 V in all curves correspond to CO stripping, involving the oxidation of adsorbed CO and its intermediates according to the literature[48-50]. CO adsorbed on the catalysts was nearly completely removed after one scan between 0.27 and 1.47 V. The onset potentials of CO stripping for PdSn/C-0.1, 0.5, 1, and 2 (marked in Supplementary Figure 8) are downshifted by 0.12, 0.15, 0.23, 0.20 V, respectively, from that of Pd/C (at 0.78 V), indicating a reduced tendency for CO poisoning. In addition, among the catalysts with varying Pd/Sn ratios, the PdSn/C-1 catalyst exhibits the largest CO oxidation/removal peak area and the most negative onset potential for CO stripping, suggesting easy CO removal and an effective enhancement of CO tolerance with Sn incorporation[24,25,51]. Notably, appropriately doped Sn, as seen in PdSn/C-1, demonstrates superior CO tolerance, thereby enhancing the catalyst's performance.

Mechanistic investigation of PdSn/C-1 for EOR

To gain insights into how the metal composition of the catalysts affects reaction pathway selectivity during EOR, we conducted in situ ATR-SEIRAS studies using PdSn/C-1 and Pd/C catalysts in 1.0 M KOH solution containing 1.0 M ethanol. The spectra obtained during the positive potential sweep for the Pd/C catalyst are shown in Figure 4A. The appearance of the CO band at 1,800 cm-1 within the voltage range of 0 to 1.3 V indicates that the catalyst is capable of undergoing C-C bond cleavage, resulting in the predominance of bridge CO adsorption on the Pd sites[52-54]. This suggests the presence of intermediates via the C1 pathway[55]; however, the adsorbed CO may lead to significant poisoning, hindering the catalytic process. Thus, the C2 pathway dominates during catalytic processes on the Pd/C catalyst, as evidenced by the peaks around 1,420 cm-1, which are assigned to the stretching vibration of CH3COO-[56].

Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation

Figure 4. In situ ATR-SEIRAS spectra of (A) Pd/C and (B) PdSn/C-1 in N2-saturated 1 M KOH + 1 M C2H5OH solution.

However, the PdSn/C-1 catalyst exhibits no adsorbed CO on its surface [Figure 4B] due to the Sn introduction. Notably, a distinct CO2 adsorption peak emerges, and its intensity increases with potential. Peaks around 2,980 cm-1, assigned to the vibrational characteristics of C-Hx fragments from the methoxy group of ethanol remaining on the catalyst surface after C-C bond cleavage[57], are observed on PdSn/C-1 (right in Figure 4B) but not on the Pd/C catalyst (right in Figure 4A). These results indicate that ethanol electrooxidation on PdSn/C-1 primarily yields CO2 through the C1 pathway via C-C bond cleavage. Additionally, the strong vibration of H2O at 1,620 cm-1 on PdSn/C-1 (absent on Pd/C) implies the production of abundant OH- species on oxyphilic Sn for the oxidation removal of *CO[31,58], resulting in significant catalytic activity. Furthermore, a weak peak around 1,400 cm-1 was observed in Figure 4B, which can be deconvoluted into two peaks at 1,402 and 1,394 cm-1 [Supplementary Figure 9A] attributed to acetate and carbonate[55], respectively. Thus, despite the limitation of intensity, the C1 selectivity (η, in %) was approximately quantified using η = Ccarbonate/(Ccarbonate + Cacetate) = 6A1394/(6A1394 + 4A1402 × 2.2) × 100%[31], where A1394 and A1402 are the peak areas at 1,394 and 1,402 cm-1, 6 and 4 refer to the number of electrons transferred per one of carbonate and acetate, respectively, and 2.2 is the ratio of the absorbance coefficient per mole of carbonate and acetate. As shown in Supplementary Figure 9B, the η of PdSn/C-1 reached approximately 35.1%, indicating a significant C-C cleavage ability. In summary, the C1 pathway involves the dissociation and adsorption of ethanol, leading to the splitting of the C-C bond and generating hydrogen-rich C1 fragments that undergo complete oxidation without forming toxic intermediates[52].

The EOR mechanism is complex, generally involving C1 and C2 reaction pathways with various surface intermediates[59,60]. To elucidate the reaction mechanism of EOR, we conducted DFT simulations to evaluate the adsorption energy of the Pd (111) surface (as considered according to the XRD results) toward the intermediates during EOR [Supplementary Figure 10]. The free energy profiles of EOR leading to C2 (CH3COO-) and C1 (CO2) products on the surface of Pd are shown in Figure 5A. When comparing the C1 and C2 pathways, they share the same process before *CH3CO, but differ in whether the *CH3CO intermediate is dehydrogenated to *CH2CO or oxidized to *CH3COOH. The free energy differences between *CH3CO and *CH3COOH (ΔG1) and between *CH3CO and *CH2CO (ΔG2) are 0.24 and 0.44 eV, respectively, which are the highest in their respective pathways. Therefore, ΔG1 and ΔG2 not only decide the pathway but also determine the EOR rate. Furthermore, to investigate the effect of tensile strain on the reaction pathway and rate of the Pd-based catalysts, we calculated ΔG1 and ΔG2 at various tensile ratios and summarized their deviation (ΔG, ΔG = ΔG2 - ΔG1) in Figure 5B. ΔG reflects the EOR pathway, favoring the C1 pathway when ΔG < 0, and the C2 pathway when ΔG > 0. For Pd without strain, the C2 product predominates due to a lower ΔG1 compared to ΔG2. As the tensile strain increases from 0% to 10%, ΔG decreases from 0.24 to -0.18 eV and then increases to 0.06 eV, achieving the minimum at a tensile strain of 5%. Additionally, ΔG2 of Pd with a 5% tensile strain is the lowest (0.45 eV) via the C1 pathway compared to other tensile strains. These results indicate that the reasonable modulation of tensile strain not only regulates the EOR process via the C1 pathway by lowering the ΔG but also accelerates the reaction kinetics by decreasing ΔG2.

Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation

Figure 5. (A) Potential energy diagram for EOR via the C1 (green) and C2 (yellow) pathway on Pd (111). (B) The calculated free energy for ΔG1 and ΔG2 on various tensile ratios of Pd (111). (C) Schematic illustration for the EOR process on Pd/C (left) and PdSn/C-1 (right).

Structural analysis, electrochemical evolution, in situ ATR-SEIRAS studies, and DFT calculations collectively demonstrate that the incorporation of Sn into the Pd catalyst enhances the catalytic performance for EOR through two pivotal processes [Figure 5C]. First, Sn doping regulates the tensile strain of the Pd crystal, reducing the energy barrier between *CH3CO and *CH2CO (ΔG2) and promoting the C1 pathway. Notably, the optimal ΔG2 for improved selectivity and kinetics is achieved only with suitable tensile strain (PdSn/C-1). Second, and importantly, the incorporation of Sn provides the oxyphilic sites for OHads species, facilitating the oxidation removal of CO for enhanced anti-poisoning. These dual contributions underlie the exceptional catalytic performance of PdSn/C-1 in EOR.

CONCLUSIONS

In conclusion, we have developed an efficient co-reduction approach to synthesize PdSn/C catalysts, enabling tunable tensile strain through modulation of Sn content. The PdSn/C-1 catalyst has demonstrated superior catalytic efficacy in EOR, achieving a mass activity of 8,452.3 mA/mgPd, surpassing that of most Pd-based catalysts. Additionally, Sn incorporation has enhanced the tolerance to COads species through oxidation removal. By combining the results of ATR-SEIRAS and DFT calculations, we found that the appropriate Sn component in PdSn/C creates an optimal tensile strain. This strain lowers the reaction energy barrier for the formation of *CH2CO from *CH3CO intermediates, a key step in EOR, and favors the C1 pathway through C-C bond cleavage. The strategy of component-regulated strains to enhance catalytic performance in EOR holds promise for the future design of advanced electrocatalysts.

DECLARATIONS

Authors’ contributions

Prepared the PdSn/C and performed most of the electrochemical experiments: Tao, Y.

Prepared the PdSn/C and physicochemical characterizations: Cai, Y.

Conducted ATR-SEIRAS experiments: Ding, J.

Performed theoretical calculations: Li, F.

Analyzed the data and wrote and discussed the whole paper: Ma, T.

Supervised students, developed concepts, acquired funding, and revised the manuscript: Chen, R.; Ma, T.; Liang, F.

Availability of data and materials

The data are available upon request from the authors.

Financial support and sponsorship

This work was financially supported by the National Natural Science Foundation of China (Nos. 21871108, U21A20317) and the Program for Innovative Teams of Outstanding Young and Middle-Aged Researchers in the Higher Education Institutions of Hubei Province (No. T201702).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2025.

Supplementary Materials

REFERENCES

1. Nguyen, A. Q. K.; Pham, H. Q.; Huynh, S. T. M.; Huynh, T. T. Milestones of electrocatalyst development for direct alcohol fuel cells. Adv. Sustain. Syst. 2023, 7, 2300205.

2. Gul Sial MA, Ud Din MA, Wang X. Multimetallic nanosheets: synthesis and applications in fuel cells. Chem. Soc. Rev. 2018, 47, 6175-200.

3. Zhao, Y.; Setzler, B. P.; Wang, J.; et al. An efficient direct ammonia fuel cell for affordable carbon-neutral transportation. Joule 2019, 3, 2472-84.

4. Shekhawat, A.; Samanta, R.; Panigrahy, S.; Barman, S. Electrocatalytic oxidation of urea and ethanol on two-dimensional amorphous nickel oxide encapsulated on N-doped carbon nanosheets. ACS. Appl. Energy. Mater. 2023, 6, 3135-46.

5. Chen, T.; Xu, S.; Zhao, T.; et al. Accelerating ethanol complete electrooxidation via introducing ethylene as the precursor for the C-C bond splitting. Angew. Chem. Int. Ed. 2023, 62, e202308057.

6. Luo, S.; Zhang, L.; Liao, Y.; et al. A tensile-strained Pt-Rh single-atom alloy remarkably boosts ethanol oxidation. Adv. Mater. 2021, 33, e2008508.

7. Wang, Y.; Zheng, M.; Li, Y.; et al. p-d orbital hybridization induced by a monodispersed Ga site on a Pt3Mn nanocatalyst boosts ethanol electrooxidation. Angew. Chem. Int. Ed. 2022, 61, e202115735.

8. Monyoncho, E. A.; Steinmann, S. N.; Michel, C.; Baranova, E. A.; Woo, T. K.; Sautet, P. Ethanol electro-oxidation on palladium revisited using polarization modulation infrared reflection absorption spectroscopy (PM-IRRAS) and density functional theory (DFT): why is it difficult to break the C-C bond? ACS. Catal. 2016, 6, 4894-906.

9. Zhou, X.; Ma, Y.; Ge, Y.; et al. Preparation of Au@Pd core-shell nanorods with fcc-2H-fcc heterophase for highly efficient electrocatalytic alcohol oxidation. J. Am. Chem. Soc. 2022, 144, 547-55.

10. Luo, M.; Guo, S. Strain-controlled electrocatalysis on multimetallic nanomaterials. Nat. Rev. Mater. 2017, 2, 17059.

11. Yang, X.; Liang, Z.; Chen, S.; et al. A phosphorus-doped Ag@Pd catalyst for enhanced C-C bond cleavage during ethanol electrooxidation. Small 2020, 16, e2004727.

12. Huang, W.; Ma, X. Y.; Wang, H.; et al. Promoting effect of Ni(OH)2 on palladium nanocrystals leads to greatly improved operation durability for electrocatalytic ethanol oxidation in alkaline solution. Adv. Mater. 2017, 29, 1703057.

13. Wang, L.; Liu, Z.; Zhang, S.; et al. In situ assembly of ultrafine AuPd nanowires as efficient electrocatalysts for ethanol electroxidation. Int. J. Hydrogen. Energy. 2021, 46, 8549-56.

14. Wang, J.; Zhang, G.; Liu, H.; et al. High-performance electrocatalytic reduction of CO2 to CO by ultrathin PdCu alloy nanosheets. Sep. Purif. Technol. 2023, 320, 124186.

15. Liu, D.; Zeng, Q.; Hu, C.; et al. Core-shell CuPd@NiPd nanoparticles: coupling lateral strain with electronic interaction toward high-efficiency electrocatalysis. ACS. Catal. 2022, 12, 9092-100.

16. Wang, H.; Zheng, H.; Ling, L.; et al. Pd metallene aerogels with single-atom W doping for selective ethanol oxidation. ACS. Nano. 2022, 16, 21266-74.

17. Zhang, Y.; Jang, H.; Ge, X.; et al. Single-atom Sn on tensile-strained ZnO nanosheets for highly efficient conversion of CO2 into formate. Adv. Energy. Mater. 2022, 12, 2202695.

18. Feng, C.; Lv, M.; Shao, J.; et al. Lattice strain engineering of Ni2P enables efficient catalytic hydrazine oxidation-assisted hydrogen production. Adv. Mater. 2023, 35, e2305598.

19. Cheng, W.; Zhao, X.; Su, H.; et al. Lattice-strained metal-organic-framework arrays for bifunctional oxygen electrocatalysis. Nat. Energy. 2019, 4, 115-22.

20. He, F.; Zheng, Q.; Yang, X.; et al. Spin-state modulation on metal-organic frameworks for electrocatalytic oxygen evolution. Adv. Mater. 2023, 35, e2304022.

21. Xu, L.; Fu, B.; Gao, F.; Ma, J. W.; Gao, H.; Guo, P. Strain engineering of face-centered cubic Pd-Pb nanosheets boosts electrocatalytic ethanol oxidation. ACS. Appl. Energy. Mater. 2023, 6, 2471-8.

22. Zhang, G.; Hui, C.; Yang, Z.; et al. Hydrogen-induced p-d orbital hybridization and tensile strain of PdGa single-atom alloy metallene boosts complete electrooxidation of ethanol. Appl. Catal. B. Environ. 2024, 342, 123377.

23. Han, S.; Ma, Y.; Yun, Q.; et al. The synergy of tensile strain and ligand effect in PtBi nanorings for boosting electrocatalytic alcohol oxidation. Adv. Funct. Mater. 2022, 32, 2208760.

24. You, J.; Zhang, Y.; Fan, G.; et al. Component dependent electrocatalytic activity of magnetic bimetallic Pd-Ni nanoparticles for formate oxidation. Mater. Lett. 2023, 343, 134387.

25. Liang, Y.; Ma, T.; Xiong, Y.; Qiu, L.; Yu, H.; Liang, F. Highly efficient blackberry-like trimetallic PdAuCu nanoparticles with optimized Pd content for ethanol electrooxidation. Nanoscale 2021, 13, 9960-70.

26. Zou, S.; Ji, Y.; Li, J.; et al. Novel leaflike Cu-O-Sn nanosheets as highly efficient catalysts for the Rochow reaction. J. Catal. 2016, 337, 1-13.

27. Hidalgo, M. C.; Murcia, J. J.; Navío, J. A.; Colón, G. Photodeposition of gold on titanium dioxide for photocatalytic phenol oxidation. Appl. Catal. A. General. 2011, 397, 112-20.

28. Maicu, M.; Hidalgo, M. C.; Colón, G.; Navío, J. A. Comparative study of the photodeposition of Pt, Au and Pd on pre-sulphated TiO2 for the photocatalytic decomposition of phenol. J. Photoch. Photobio. A. Chem. 2011, 217, 275-83.

29. Primo, A.; Corma, A.; García, H. Titania supported gold nanoparticles as photocatalyst. Phys. Chem. Chem. Phys. 2011, 13, 886-910.

30. Salomé, S.; Ferraria, A. M.; Botelho do Rego, A. M.; Alcaide, F.; Savadogo, O.; Rego, R. Enhanced activity and durability of novel activated carbon-supported PdSn heat-treated cathode catalyst for polymer electrolyte fuel cells. Electrochim. Acta. 2016, 192, 268-82.

31. Ye, N.; Sheng, W.; Zhang, R.; Yan, B.; Jiang, Z.; Fang, T. Interfacial electron engineering of PdSn-NbN/C for highly efficient cleavage of the C-C bonds in alkaline ethanol electrooxidation. Small 2024, 20, e2304990.

32. Yang, X.; Wang, Y.; Tong, X.; Yang, N. Strain engineering in electrocatalysts: fundamentals, progress, and perspectives. Adv. Energy. Mater. 2022, 12, 2102261.

33. Wang, W.; Kang, Y.; Yang, Y.; Liu, Y.; Chai, D.; Lei, Z. PdSn alloy supported on phenanthroline-functionalized carbon as highly active electrocatalysts for glycerol oxidation. Int. J. Hydrogen. Energy. 2016, 41, 1272-80.

34. Spinacé, E. V.; Linardi, M.; Neto, A. O. Co-catalytic effect of nickel in the electro-oxidation of ethanol on binary Pt-Sn electrocatalysts. Electrochem. Commun. 2005, 7, 365-9.

35. Okamoto, H. Pd-Sn (Palladium-Tin). J. Phase. Equilib. Diffus. 2012, 33, 253-4.

36. Dietrich, C.; Chen, S.; Uzunidis, G.; Hähsler, M.; Träutlein, Y.; Behrens, S. Bimetallic Pd/Sn-based nanoparticles and their catalytic properties in the semihydrogenation of diphenylacetylene. ChemistryOpen 2021, 10, 296-304.

37. Lanza, R.; Bersani, M.; Conte, L.; et al. Effect of crystalline phase and composition on the catalytic properties of PdSn bimetallic nanoparticles in the PROX reaction. J. Phys. Chem. C. 2014, 118, 25392-402.

38. Wang, K. W.; Kang, W. D.; Wei, Y. C.; et al. Promotion of PdCu/C catalysts for ethanol oxidation in alkaline solution by SnO2 modifier. ChemCatChem 2012, 4, 1154-61.

39. Wu, L.; Yan, H.; Li, X.; Wang, X. Characterization and photocatalytic properties of SnO2-TiO2 nanocomposites prepared through gaseous detonation method. Ceram. Int. 2017, 43, 1517-21.

40. Wei, Z.; Ding, J.; Duan, X.; et al. Enhancing selective electrochemical CO2 reduction by in situ constructing tensile-strained Cu catalysts. ACS. Catal. 2023, 13, 4711-8.

41. Li, T.; Zhang, L.; Zhang, L.; et al. Tailoring the chemisorption manner of Fe d-band center with La2O3 for enhanced oxygen reduction in anion exchange membrane fuel cells. Adv. Funct. Mater. 2024, 34, 2309886.

42. Nørskov, J. K.; Abild-Pedersen, F.; Studt, F.; Bligaard, T. Density functional theory in surface chemistry and catalysis. Proc. Natl. Acad. Sci. USA. 2011, 108, 937-43.

43. Zhao, C.; Wang, J.; Gao, Y.; et al. D-orbital manipulated Ru nanoclusters for high-efficiency overall water splitting at industrial-level current densities. Adv. Funct. Mater. 2024, 34, 2307917.

44. Zamora, Z. J. A.; Stevens, M. B.; Gunasooriya, G. T. K. K.; et al. Tuning the electronic structure of Ag-Pd alloys to enhance performance for alkaline oxygen reduction. Nat. Commun. 2021, 12, 620.

45. Gao, Q.; Mou, T.; Liu, S.; et al. Monodisperse PdSn/SnOx core/shell nanoparticles with superior electrocatalytic ethanol oxidation performance. J. Mater. Chem. A. 2020, 8, 20931-8.

46. Pu, Y.; Celorrio, V.; Stockmann, J. M.; et al. Surface galvanic formation of Co-OH on Birnessite and its catalytic activity for the oxygen evolution reaction. J. Catal. 2021, 396, 304-14.

47. Ahmed, M. S.; Jeon, S. Highly active graphene-supported NixPd100-x binary alloyed catalysts for electro-oxidation of ethanol in an alkaline media. ACS. Catal. 2014, 4, 1830-7.

48. Guo, R. H.; Liu, C. F.; Wei, T. C.; Hu, C. C. Electrochemical behavior of CO2 reduction on palladium nanoparticles: dependence of adsorbed CO on electrode potential. Electrochem. Commun. 2017, 80, 24-8.

49. Gao, D.; Zhou, H.; Cai, F.; et al. Switchable CO2 electroreduction via engineering active phases of Pd nanoparticles. Nano. Res. 2017, 10, 2181-91.

50. Torralba, E.; Blanchard, N.; Cachet-Vivier, C.; Muller-Bouvet, D.; González, J.; Bastide, S. Electrochemical study of carbon dioxide reduction at copper-palladium nanoparticles: influence of the bimetallic composition in the CO poisoning tolerance. Electrochim. Acta. 2020, 354, 136739.

51. Qiu, L.; Tao, Y.; Ma, T.; Liang, F. Synergistic effect of trimetallic PdCuIn nanoparticles in ethanol and formate oxidation reaction for remarkable catalytic performance. Energy. Technol. 2024, 12, 2301618.

52. Zhang, Y.; Liu, X.; Liu, T.; et al. Rhombohedral Pd-Sb nanoplates with Pd-terminated surface: an efficient bifunctional fuel-cell catalyst. Adv. Mater. 2022, 34, e2202333.

53. Demarconnay, L.; Brimaud, S.; Coutanceau, C.; Léger, J. Ethylene glycol electrooxidation in alkaline medium at multi-metallic Pt based catalysts. J. Electroanal. Chem. 2007, 601, 169-80.

54. Han, C.; Lyu, Y.; Wang, S.; et al. Highly utilized active sites on Pt@Cu/C for ethanol electrocatalytic oxidation in alkali metal hydroxide solutions. Adv. Funct. Mater. 2023, 33, 2305436.

55. Liang, Z.; Song, L.; Deng, S.; et al. Direct 12-electron oxidation of ethanol on a ternary Au(core)-PtIr(Shell) electrocatalyst. J. Am. Chem. Soc. 2019, 141, 9629-36.

56. Zhang, J.; Ye, J.; Fan, Q.; et al. Cyclic penta-twinned rhodium nanobranches as superior catalysts for ethanol electro-oxidation. J. Am. Chem. Soc. 2018, 140, 11232-40.

57. Torrero, J.; Montiel, M.; Peña, M. A.; Ocón, P.; Rojas, S. Insights on the electrooxidation of ethanol with Pd-based catalysts in alkaline electrolyte. Int. J. Hydrogen. Energy. 2019, 44, 31995-2002.

58. Farsadrooh, M.; Torrero, J.; Pascual, L.; Peña, M. A.; Retuerto, M.; Rojas, S. Two-dimensional Pd-nanosheets as efficient electrocatalysts for ethanol electrooxidation. Evidences of the C-C scission at low potentials. Appl. Catal. B. Environ. 2018, 237, 866-75.

59. Wu, R.; Wang, L. A density functional theory study on the mechanism of complete ethanol oxidation on Ir(100): surface diffusion-controlled C-C bond cleavage. J. Phys. Chem. C. 2020, 124, 26953-64.

60. Wang, Y.; Zou, S.; Cai, W. B. Recent Advances on electro-oxidation of ethanol on Pt- and Pd-based catalysts: from reaction mechanisms to catalytic materials. Catalysts 2015, 5, 1507-34.

Cite This Article

Article
Open Access
Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation
Yin Cai, ... Feng LiangFeng Liang

How to Cite

Cai, Y.; Tao, Y.; Ding, J.; Li, F.; Chen, R.; Ma, T.; Liang, F. Composition-regulated lattice strain of PdSn/C for boosting C1 pathway in ethanol electrooxidation. Energy Mater. 2025, 5, 500038. http://dx.doi.org/10.20517/energymater.2024.91

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

© The Author(s) 2025. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
149
Downloads
46
Citations
0
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Energy Materials
ISSN 2770-5900 (Online)
Follow Us

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/