ACS Publications. Most Trusted. Most Cited. Most Read
Synergistic Bimetallic PdNi Nanoparticles: Enhancing Glycerol Electrooxidation While Preserving C3 Product Selectivity
My Activity

Figure 1Loading Img
  • Open Access
Article

Synergistic Bimetallic PdNi Nanoparticles: Enhancing Glycerol Electrooxidation While Preserving C3 Product Selectivity
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

ACS Applied Energy Materials

Cite this: ACS Appl. Energy Mater. 2024, 7, 5, 1802–1813
Click to copy citationCitation copied!
https://doi.org/10.1021/acsaem.3c02789
Published February 29, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Abstract

Click to copy section linkSection link copied!

Electrochemical conversion of glycerol offers a promising route to synthesize value-added glycerol oxidation products (GOPs) from an abundant biomass-based resource. While noble metals provide a low overpotential for the glycerol electrooxidation reaction (GEOR) and high selectivity toward three-carbon (C3) GOPs, their efficiency and cost can be improved by incorporating non-noble metals. Here, we introduce an effective strategy to enhance the performance of Pd nanoparticles for the GEOR by alloying them with Ni. The resulting PdNi nanoparticles show a significant increase in both specific activity (by almost 60%) and mass activity (by almost 35%) during the GEOR at 40 °C. Additionally, they exhibit higher resistance to deactivation compared to pure Pd. Analysis of the GOPs reveals that the addition of Ni into Pd does not compromise the selectivity, with glycerate remaining at around 60% of the product fraction and the other major product being lactate at around 30%. Density functional theory calculations confirm the reaction pathways and the basis for the higher activity of PdNi. This study demonstrates a significant increase in the GEOR catalytic performance while maintaining the selectivity for C3 GOPs, using a more cost-effective nanocatalyst.

This publication is licensed under

CC-BY 4.0 .
  • cc licence
  • by licence
Copyright © 2024 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

In recent years, biodiesel has emerged as a promising biofuel derived from vegetable oils, animal fat, and their residues. (1,2) The production of biodiesel generates glycerol (Gly) as a significant byproduct, accounting for approximately 10 wt % of the biodiesel produced. (3) The global growth of the biodiesel industry has resulted in a surplus of Gly, causing its market price to decline. (4,5) Consequently, there is growing interest in converting this excess Gly into commercially valuable chemicals. One promising approach for Gly valorization is through the Gly electrooxidation reaction (GEOR), which enables the simultaneous production of valuable chemicals on the anode from Gly and hydrogen on the cathode with lower energy requirements than conventional water electrolysis. (5)
Non-noble and noble metal-based bimetallic and trimetallic electrocatalysts have shown remarkable results for the GEOR by achieving higher activity than the monometallic counterparts. (6−10) Among these, Pd-based nanocatalysts exhibit great potential for the GEOR due to their outstanding performance in alkaline environments, similar cost and abundance to Pt, (11) and lower susceptibility to CO poisoning. (12) Although in the monometallic form, non-noble transition metals such as Ni require a significantly higher potential than Pd to electrochemically oxidize Gly; (13) the addition of such metals to noble metals can significantly enhance the catalytic activity and stability for the GEOR. (11,14−16) Further approaches to improve the catalytic performance of nanomaterials include tuning their morphology and increasing the surface area by introducing porosity or reducing the particle size. (17) The use of PdNi nanocatalysts extends beyond the GEOR to the electrooxidation of other biomass derivatives and organic molecules such as formic acid, (18,19) ethylene glycol, (20) carbohydrazide, (21) ethanol, (22−25) and methanol. (26−28) Aside from nanocatalysts, electrodeposited PdNi films have also been studied for the electrooxidation of lactic (29) and gluconic acids. (30)
The role of Ni in PdNi catalysts in enhancing the electrooxidation of alcohols has been mainly studied for ethanol (25,31,32) and methanol. (27) For ethanol oxidation, the presence of Ni significantly increased the current density compared to pure Pd when Ni served as the surface layer of sputtered PdNi catalysts, in contrast to when it was placed at the sublayer. (32) These findings were supported by density functional theory (DFT) calculations, which assessed the impact of varying the percentage of surface Ni on a PdNi catalyst with (111) facets. (23) These studies highlight the importance of the structural design of PdNi catalysts. For methanol oxidation, a combined experimental and DFT study revealed that increased Ni concentrations in PdNi nanocatalysts, and specifically Pd0.5Ni0.5 (at. %), led to higher selectivity toward CO2 formation, suppressed CO poisoning, and improved activity for methanol electrooxidation. (27)
For the GEOR, Pd0.5Ni0.5 (at. %) nanoparticles (NPs) have shown a higher current density than Pd NPs, (11) which aligns with the general consensus that Pd0.5Ni0.5 (at. %) represents the optimal Pd/Ni atomic ratio for Pd-rich catalysts. (14,15) Nevertheless, there is still much to explore regarding the formation of glycerol oxidation products (GOPs) using PdNi electrocatalysts. In alkaline media, Pd-based catalysts typically yield products such as glyceric acid/glycerate (GLA), tartronic acid/tartronate (TTA), oxalic acid/oxalate (OA), glycolic acid/glycolate (GA), and formic acid/formate (FA). (7) However, there are limited reports quantifying GOPs produced by PdNi catalysts. (33) FA, GA, and OA, formed through C–C bond cleavage, are believed to occur at high oxidation potentials, (34) but adsorbed CO has been observed at potentials below the onset of the GEOR on Pd and PdNi nanocatalysts, indicating C–C bond cleavage from a dissociative Gly adsorption step at low potentials. (11) Both scenarios are undesirable because the three carbon (C3) GOPs GLA, LA, and TTA are more valuable due to their uses in important industries such as biomedical (GLA and TTA) and pharmaceuticals (LA), (35) making the mitigation of C–C bond cleavage a primary aim in GEOR developments.
While the typical synthesis of PdNi nanocatalysts involves methods such as coprecipitation, sputtering, electrodeposition, and others, the alloying process can be challenging due to the need for high temperatures, particular additives and reductants, pH adjusters, and multistep synthetic procedures. (36−38) Herein, we report the catalytic performance for the GEOR in alkaline media of sub-10 nm Pd and PdNi NPs, obtained via a modified one-step metal reduction synthesis in organic medium (39) and characterized by transmission electron microscopy (TEM) and powder X-ray diffraction (PXRD). The Pd and PdNi NPs were deposited onto carbon paper electrodes (Pd/C and PdNi/C, respectively), and their electrocatalytic activity was studied by cyclic voltammetry (CV) and iR-corrected polarization curves (ICPCs). Galvanostatic electrolysis was employed to generate GOPs, and the product distribution was determined via high-performance liquid chromatography (HPLC), with 90% of the product fractions containing C3 GOPs. DFT calculations confirmed similar reaction pathways for GEOR on Pd(111) and PdNi(111). PdNi/C outperformed Pd/C in terms of both mass and specific activity while exhibiting minimal differences in the distribution of the GOPs. Our study demonstrates the promising performance of PdNi NPs in GEOR by maintaining selectivity for C3 products similar to that of Pd while enhancing the electrocatalytic activity and reducing the overall catalyst cost through a reduction in the required amount of Pd.

Experimental Section

Click to copy section linkSection link copied!

Materials and Reagents

Palladium(II) acetylacetonate (Pd(acac)2, Sigma-Aldrich, 99%), nickel(II) acetylacetonate (Ni(acac)2, Combi-Blocks Inc., 98%), oleylamine (OAm, Sigma-Aldrich, technical grade, 70%), borane tert-butylamine (BTB, Alfa Aesar, 97%), 1-octadecene (ODA, Sigma-Aldrich, for synthesis), acetone (Honeywell, ≥99.5%), and ethanol (absolute, VWR Chemicals, ≥99.7%) were used for the catalyst synthesis. For the GEOR experiments, the carbon fiber paper (H23, thickness 210 μm, Freudenberg) was cut into 1 cm2 pieces before hydrophilic treatment. Analytical-grade sodium hydroxide (NaOH, Merck) and Gly (Merck) were both purchased from VWR (Radnor, PA, USA). The ultrapure water (18 MΩ) used to prepare all solutions was obtained with a Millipore DirectQ3 purification system from Millipore (Burlington, MA, USA).

Electrochemical Instrumentation

A Princeton Applied Research PAR273A potentiostat/galvanostat from AMETEK (Minneapolis, MN, USA) was used for electrochemical measurements. To record ICPCs, a current interrupt method was used through the PAR273A in connection with a National Instruments (NI) cDAQ-9178 chassis and a NI9223 voltage digitalizer (Austin, TX, USA), as reported previously. (40) A Hg/HgO reference electrode (RE) (RE-A6P, BioLogic, 1.0 M NaOH) was used for all of the measurements.

Synthesis of Pd and PdNi Electrocatalysts

Pd and PdNi NPs were synthesized via the reduction of Pd(acac)2 and coreduction of Pd(acac)2 and Ni(acac)2, respectively, in an organic medium with OAm serving as a surfactant, BTB as a reducing agent, and ODE as a solvent (see Scheme 1).

Scheme 1

Scheme 1. Schematic Representation of the Pd and PdNi Nanoparticle Synthesis

Synthesis of Pd Nanoparticles

Pd NPs were synthesized using a modified previously reported procedure. (39) A metal-precursor mixture containing 150 mg of Pd(acac)2 (ca. 0.5 mmol) in 30 mL of OAm in a 50 mL round-bottomed flask equipped with a rubber septum was flushed with argon gas for 30 min to remove excess oxygen. After that, the mixture was heated to 90 °C under magnetic stirring (800 rpm) and an argon atmosphere until it became homogeneous. Next, a reductant-solvent mixture of 600 mg of BTB and 6 mL of OAm, preheated at 60 °C, was quickly injected into the metal-precursor mixture with a syringe. The resulting reaction mixture was allowed to react for 1 h under the same stirring rate and argon stream. The resulting solution was washed by centrifugation four times with an acetone/ethanol mixture (7:3) at 12,000 rpm for 15 min, followed by final redispersion in 1 mL of ethanol and vacuum drying at 60 °C for 24 h.

Synthesis of PdNi Nanoparticles

PdNi NPs were synthesized using a modified previously reported procedure. (39) A reductant-solvent mixture containing 400 mg of BTB, 6 mL of OAm, and 14 mL of ODE in a 50 mL round-bottomed flask equipped with a rubber septum was flushed with argon gas for 30 min to remove excess oxygen. After that, the mixture was heated to 90 °C under magnetic stirring (800 rpm) and an argon atmosphere until it became homogeneous. Next, a metal-precursor mixture of 38.08 mg of Pd(acac)2 (0.125 mmol) and 128.46 mg of Ni(acac)2 (0.300 mmol) in 6 mL of OAm, preheated at 60 °C, was quickly injected into the reductant-solvent mixture with a syringe. The resulting reaction mixture was allowed to react for 1 h under the same stirring rate and argon stream. Finally, the resulting solution was washed, redispersed, and dried similarly to that of the Pd NPs.
The excess Ni precursor introduced in solution is due to the tendency of Pd to be reduced much more readily than Ni. Therefore, to incorporate the desired amount of Ni into the PdNi bimetallic catalyst, a higher Ni content than Pd is required. Since the main goal was to study the performance of PdNi NPs with a Pd/Ni ratio of 1:1, the synthetic procedure was not further optimized to compensate for the losses. The optimization can therefore be a basis for future work and synthetic strategy improvement. Additionally, if the remaining Ni was simply discarded, it would decrease the cost-effectiveness and greenness of the catalyst synthesis process; however, the remaining Ni could easily be recovered as pure Ni through electrodeposition or another reduction process.

Characterization of Electrocatalysts

TEM imaging and energy-dispersive X-ray spectroscopy (EDS) were carried out on a JEOL JEM-2100F microscope operating at an acceleration voltage of 200 kV. TEM samples were prepared by depositing ∼20 μL of NP dispersion in ethanol on a 300-mesh ultrathin Lacey carbon-supported gold grid and allowing the solvent to evaporate. PXRD was performed on a Bruker D8 DISCOVER diffractometer using a Cu Kα X-ray source (λCuKα1 = 1.540598 Å and λCuKα2 = 1.544426 Å), equipped with an LYNXEYE XE-T detector. The diffractograms were collected within the 2θ range of 10–90° with a step size of 0.01° and scan rate of 0.375°/min.

Electrode Fabrication

The carbon paper working electrode (WE) served as the substrate for drop casting the catalyst ink, following pretreatment by ultrasonication in 3 M HCl, ultrapure water, and ethanol for 15, 10, and 5 min, respectively. It was then dried at 100 °C for 1 h to remove any contaminants and improve hydrophilicity. The catalyst ink was composed of 5 mg of the catalyst (either Pd or PdNi NPs), 5 mg of the Vulcan XC-72 carbon powder, 40 μL of 5% Nafion solution, 240 μL of isopropanol, and 720 μL of ultrapure water, resulting in a catalyst particle concentration of 5 mg mL–1. The catalyst ink was stirred overnight before being drop cast. The drop-casting process involved placing the carbon paper approximately 1.5 cm above a hot plate set at 60 °C. Subsequently, 50 μL of the catalyst ink was drop-cast onto one side of the carbon paper and allowed to dry for 10 min. After drying, the carbon paper was then flipped and an additional 50 μL of the catalyst ink was applied to the other side. After a 10 min drying period, the electrode was removed from the heat and left to dry in air overnight. In total, 0.5 mg of the catalysts were applied to the carbon paper, leading to a Pd mass of 0.5 mg for Pd/C and 0.299 mg of Pd and 0.201 mg of Ni for PdNi/C. The mass of Pd and Ni was determined by eqs S1 and S2, respectively.

Calculations of ECSA

A three-electrode cell (Figure S1) was used to calculate the Pd-based ECSA of Pd/C and PdNi/C by measuring the charge transferred by the reduction of a monolayer of PdO. In this setup, both anodic and cathodic chambers consisted of an N2-purged 1.0 M NaOH, Gly-free solution at 25 °C with a Nafion (212) membrane dividing the cell. The Nafion membrane was used to mitigate the crossover of GOPs from the anode chamber into the cathode chamber to prevent the possibility of the Pt cathode reducing the formed GOPs. Therefore, the ions passing through the Nafion membrane are most likely Na+ ions. The catalytic electrode served as the WE, a Pt grid served as the counter electrode (CE) in a 1 M NaOH solution, and the RE was Hg/HgO (1 M NaOH). It has been previously established that a monolayer of PdO is formed in 1 M NaOH at approximately 1.25 V vs RHE, (41) also corroborated in acidic conditions. (42,43) Hence, CV between 0.265 and 1.250 V vs RHE at a sweep rate of 50 mV s–1 was initially applied for 20 cycles under solution stirring to stabilize the catalysts.
Then, the solution was held stagnant, and two cyclic voltammograms were recorded within the same potential window and sweep rate. The second cycle was used for ECSA calculation by determining the associated charge of the PdO reduction peak and using eq 1
ECSA=Q(C)S(C/cm2)=areaofPdOreductionpeak405μC/cm2
(1)
where Q is the charge passed during the reduction of PdO to Pd, and S is the characteristic charge density (405 μC cm–2) of the reduction of a monolayer of PdO to Pd. (43)
The ECSA measurements are indicative of the catalysts that were used for the GEOR measurements. Hence, the normalization of the current shown for the Pd and PdNi NPs by the ECSA for the GEOR measurements was done with catalysts that had already been through cycling to the point where the current response was stable.

GEOR Electrochemical Measurements

All GEOR experiments were undertaken in a divided three-electrode cell (Figure S1), with the anode and cathode compartments separated by using a Nafion 212 membrane (VWR). The Pd/C and PdNi/C electrocatalysts were the WEs, a Pt grid was the CE, and the RE was Hg/HgO (1 M NaOH). The WE potential was converted to the reversible hydrogen electrode (RHE) scale using eq 2
ERHE=EHg/HgORef+EHg/HgOM+(0.0591×pH)
(2)
where ERHE is the calculated potential of the electrode in V vs RHE, EHg/HgORef is 0.120 V vs SHE at 25 °C (determined from laboratory measurement), EHg/HgOM is the measured or controlled potential during the experiments when using Hg/HgO (1 M NaOH), and the pH is 14 as the supporting electrolyte was 1 M NaOH. The respective solutions studied consisted of 0.10 and 0.50 M Gly in 1.0 M NaOH, purged with nitrogen before experiments. The electrochemical measurements were conducted at 25 and 40 °C, with a stirring bar at 200 rpm. The RE was kept at ambient temperature using a Luggin capillary. The activity of the catalysts for GEOR was mainly determined via ICPCs, which were conducted from low to high current densities. Steady-state galvanostatic measurements were used for GOP analysis.

Product Analysis by HPLC

HPLC analyses were performed with an Agilent 1260 Infinity II isocratic pump, multisampler, and multicolumn thermostat with a 1290 Infinity II refractive index detector. The analytical columns, in series, included a Bio-Rad guard column with a standard cartridge holder and a Micro-Guard cation H+ cartridge (4.6 × 30 mm), Bio-Rad Aminex HPX-87H column (7.8 × 300 mm), and Shodex SUGAR SH1011 column (8 × 300 mm). All columns were kept at 30 °C. To achieve optimal peak separation and ensure accurate results in the analysis of GOPs, two mobile phases were used: 1 mM H2SO4 and 8 mM H2SO4 at a flow rate of 0.25 mL min–1. The analysis time for each sample was 85 min. All standards for GOPs were calibrated from 0.1 to 10 mM with linear regression R2 values greater than 0.99. Samples were taken in 100 μL aliquots and diluted with 100 μL of ultrapure water. Chromatograms for each standard analyzed can be seen in Figures S2 and S3.

First-Principles Calculations

The slab models of the catalytic surfaces (Figure S4) were constructed using the atomic simulation environment (ASE). (44) The calculations were carried out on a (4 × 4 × 4) slab with a vacuum of 20 Å to avoid interactions between periodic images (Figure S4). The structures were optimized to energy and atom force convergence criteria of 10–5 eV and 0.01 eV·Å–1, respectively. Spin-polarized calculations were carried out, and the most stable spin state for all systems is reported in this work. DFT calculations were carried out using the Vienna Ab initio Simulation Package (VASP). (45) The exchange and correlation functional was approximated as the Bayesian error estimation functional, with an additional nonlocal correction term (BEEF-vdW). (46) Potential determining steps (PDSs) were determined following our previous studies. (47,48) As observed in Figures S5–S10, the GEOR was simulated considering three elementary pathways: (1) proton removal, (2) hydrolysis, and (3) hydrogen rearrangement. The DFT total energies were corrected using zero-point energies (ZPEs) and thermal contributions. The harmonic approximation was employed for the vibrational frequency calculations for all of the adsorbates and molecules. Henry’s law was used to link the gas-phase pressure to the aqueous concentration using the NIST-JANAF tables. (49) To account for the electrode/electrolyte interface, implicit water solvation was used as implemented in VASP. (50) The electrochemical environment was modeled based on a computational hydrogen electrode. (51) Further information regarding the computational methodology is provided in the Supporting Information.

Results and Discussion

Click to copy section linkSection link copied!

Characterization of the Nanoparticles

Figure 1a,c shows the TEM images and size distribution histograms of the well-defined Pd and PdNi NPs, respectively. The NPs were monodispersed with a mean particle diameter of 5.9 ± 0.7 nm for Pd and 2.98 ± 0.22 nm for PdNi. The Pd/Ni atomic ratio was nearly 1:1 (Pd0.45Ni0.55) as determined by TEM–EDS. HRTEM images of single Pd and PdNi NPs are shown in Figure 1b,d, respectively. The d-spacings measured from the lattice fringes, and averaged over 10 individual particles, were 0.227 and 0.225 nm for Pd and PdNi NPs, respectively, which is close to the lattice spacing for Pd(111) (0.223 nm) and agrees well with d111 ≈ 0.22 nm from the corresponding X-ray diffractograms (Figure 1e). The PXRD pattern of Pd NPs shows characteristic diffraction peaks that can be assigned to the (111), (200), (220), and (311) planes, which matched with face-centered cubic Pd (JCPDS 04-015-9339). The PXRD pattern of PdNi NPs exhibits the same features but with peaks widening due to a smaller particle size and a slight shift by ∼0.5° toward the higher 2θ values. The small peaks in the PXRD diffractogram of the PdNi alloy at 74, 76, and 78° are artifacts and not diffraction peaks.

Figure 1

Figure 1. TEM and HRTEM images of (a,b) Pd NPs and (c,d) PdNi NPs. The insets of images (a,b) show particle size distributions of the corresponding catalysts. (e) PXRD patterns of Pd and PdNi NPs.

The crystallite sizes were estimated based on the (111) peaks using the Scherrer equation and were found to be 4.71 nm for Pd and 2.46 nm for PdNi NPs, which is slightly smaller than the TEM measurements indicating some polycrystallinity. (39) The characterization demonstrates that the synthesized Pd and PdNi NPs were predominantly composed of Pd(111) and PdNi(111) facets. Additionally, the lattice parameter of PdNi NPs, aPdNi = 3.88 Å, was smaller than that of Pd, aPd = 3.93 Å, indicating a lattice contraction due to successful alloying by incorporating Ni into the Pd lattice. (52−54)

Electrochemical Characterization

The ECSA of the Pd/C and PdNi/C catalysts was calculated through CV measurements in 1.0 M NaOH (Figure 2a). In the potential window of 0.5 to 1.0 V vs RHE, the cathodic peak is assigned to the reduction of PdO to Pd. The charge associated with this peak for Pd/C led to an ECSA of 68 ± 10 cm2, slightly more than double that of PdNi/C (29 ± 7 cm2). The total physical surface area for the PdNi NPs was estimated to be about 2.3 times larger than that of the Pd NPs by considering the mass loading, atomic composition, and morphology, assuming spherical NPs. Given these calculations, the ratio of the available Pd at the surface for Pd NPs to PdNi NPs should be approximately 1:1 using the atomic ratios determined by TEM–EDS. However, the ECSA of PdNi/C is less than half of that of Pd/C. There are two possible reasons that this might be the case.

Figure 2

Figure 2. (a) CV in 1.0 M NaOH for Pd/C and PdNi/C. (b) LSV in 1.0 M NaOH and 0.50 M Gly for Pd/C and PdNi/C along with the CV from (a). (c) LSV from (b) normalized by ECSA. (d) LSV from (b) normalized by the mass of Pd in the Pd/C and PdNi/C catalysts. All sweeps were conducted at 25 °C and a sweep rate of 50 mV s–1.

The first is that the PdNi NPs may have a PdNi alloy composition gradient from the core to the surface with a more Pd-rich core and a more Ni-enriched surface. The second is that the smaller PdNi NPs have a higher packing density due to them being half the size of the Pd NPs, as it has been reported that the smaller the NPs, the higher the packing density. (55) This phenomenon for NPs has been widely reported in the literature. (56−58) However, the PdNi particles in the present study are very small (2.5–3.5 nm), and it is unlikely that there is much segregation in composition when comparing the center and rim of the particles. Additionally, the PXRD pattern for the PdNi alloy in Figure 1e also shows symmetric peaks without shoulders. Therefore, the Pd/Ni ratios determined through TEM–EDS are likely representative of the surface composition, and the decreased ECSA may be a result of the decreased physical surface area due to the packing density of the PdNi NPs. Additionally, based on our previous work on PdNi electrocatalysts for the GEOR, (16,33) the chemical state of Pd and Ni in the PdNi/C catalyst is likely to be elemental Pd and predominantly Ni(OH)2, respectively.
To ascertain the electrochemical response of the GEOR for Pd/C and PdNi/C, linear sweep voltammetry (LSV) was recorded in 1.0 M NaOH and 0.50 M Gly (Figure 2b). The addition of Gly resulted in a significant increase in the anodic current, i.e., the GEOR, with an onset potential of approximately 0.6 V vs RHE for both catalysts. Although the LSV for Pd/C in Figure 2b reaches a higher current (mA) than PdNi/C at the most positive potential, when normalized by the ECSA of the active Pd sites and the mass of Pd in the catalysts, the PdNi/C has higher performance (specific activity: Pd/C = 1.3 mA cm–2; PdNi/C = 2.2 mA cm–2, mass activity: Pd/C = 170 mA mgPd–1; PdNi/C = 211 mA mgPd–1). These comparisons can be seen in Figure 2c,d for specific activity and mass activity, respectively. However, given the notable iR-drop observed in the LSV, the performance of both electrocatalysts was comparatively evaluated through ICPCs.
To determine the most suitable anolyte for the GEOR, ICPCs for PdNi/C were recorded using two different Gly concentrations, 0.10 and 0.50 M, in a supporting electrolyte of 1.0 M NaOH, see Figure 3. The ICPCs showed a similar trend; as the current is increased, the corresponding potential also increases up until a specific current level is reached, at which point it sharply spikes. This spike represents the deactivation of the catalyst for the GEOR, which occurs due to the oxidation of the active Pd sites. (16,59) The current and corresponding potentials preceding this spike are denoted as the critical current icr (or critical current density jcrECSA for specific activity, or critical current when normalized by mass jcrmass for mass activity) and the critical potential Ecr, respectively. For 0.50 M Gly solution, the icr, jcrECSA, and jcrmass are all approximately double compared to 0.10 M Gly, demonstrating that the higher Gly concentration results in enhanced overall performance for the GEOR. However, despite the fivefold increase in Gly concentration, this did not translate into a fivefold increase in activity. Notably, the PdNi/C catalyst also reached a higher Ecr in the solution with a higher Gly concentration, indicating that catalyst deactivation occurred at a higher potential. Since the 0.50 M Gly solution resulted in icr, jcrECSA, jcrmass, and Ecr values higher than those of the 0.10 M solution, the former was chosen to compare the performance of Pd/C and PdNi/C at 40 °C.

Figure 3

Figure 3. ICPCs for PdNi/C at 40 °C in a supporting electrolyte of 1.0 M NaOH with Gly concentrations of 0.10 M (blue triangles) and 0.50 M (red circles). (a) No normalization, i.e., current; (b) specific activity, i.e., normalized by ECSA; and (c) mass activity, i.e., normalized by the mass of Pd in the PdNi/C catalyst.

The comparative ICPCs presented in Figure 4a reveal that the Pd/C catalyst achieves a higher icr of 199 mA compared to PdNi/C, which exhibits a icr of 159 mA. This higher icr is expected since Pd/C has a higher mass of Pd and a larger ECSA than PdNi/C, so the overall current is naturally higher, considering that the GEOR occurs on Pd sites. However, when the data are normalized by ECSA in Figure 4b, PdNi/C emerges as the superior catalyst by active sites for the GEOR. PdNi/C reaches a jcrECSA of 4.4 mA cm–2 compared to that of 2.8 mA cm–2 for Pd/C. This represents a 57% increase in specific activity by the PdNi/C catalyst. The significant increase in specific activity for PdNi/C was not equivalently repeated in terms of mass activity. However, the mass activity of PdNi/C shown in Figure 4c reaches a jcrmass of 531 mA mgPd–1 which is 34% higher than that of Pd/C at 397 mA mgPd–1. Given that the active surface of PdNi/C contains a smaller portion of available Pd than Pd/C, as discussed earlier, these results show that PdNi/C is inherently a more active catalyst for the GEOR, clearly demonstrating the positive role of Ni, at an approximate 1:1 Pd/Ni ratio. Furthermore, since the cost of Pd, at the time of writing, is nearly 2000 times higher than Ni, (60,61) by substituting approximately 40% of Pd, by mass, with Ni in the form of PdNi, a higher jcrmass can be reached at around two-thirds of the cost. Hence, the addition of Ni to Pd NPs enables a superior catalyst at a lower price. In Figure 4a–c, the Ecr is consistently higher for PdNi/C than for Pd/C in all three cases, indicating that Pd/C deactivates more readily at a potential lower than that of PdNi/C. This further emphasizes the advantage and superiority of the PdNi/C catalyst.

Figure 4

Figure 4. ICPCs for Pd/C (black squares) and PdNi/C (red circles) at 40 °C in a 1.0 M NaOH and 0.50 M Gly solution. (a) No normalization, i.e., current; (b) specific activity, i.e., normalized by ECSA; and (c) mass activity, i.e., normalized by the mass of Pd in the Pd/C and PdNi/C catalysts.

Since it is well known that for the GEOR, Ni oxidizes Gly at significantly higher potentials than Pd, the Ni present in the PdNi/C catalyst is likely only aiding in the activity of the Pd sites for the GEOR. Indeed, significant work has been done to show that the GEOR on Ni occurs as a result of the oxidation state change in Ni from Ni2+ to Ni3+ in the forms of Ni(OH)2 and NiOOH, respectively. (62) The oxidation state change occurs around 1.4 V vs RHE, and therefore studies on Ni for the GEOR regularly show the onset potential being around the mentioned voltage. Hence, the GEOR onset potential for Ni compared to Pd is around 700 mV higher. (16) Additionally, an increased onset potential and utilization of monometallic catalysts such as non-noble transition metals like Ni reportedly result in a significant C–C cleavage in highly alkaline conditions and, consequently, decreased yields of highly valued C3 GOPs are reported. (63) This is not the case from the HPLC GOP analysis presented in this study (vide infra).
The reason for the superior performance when noble metals are combined with non-noble metals is a topic of debate in the literature. The two main themes are the bifunctional mechanism and electronic effect. For the bifunctional mechanism, the addition of an oxophilic metal such as Ni to a noble metal catalyst enhances the catalytic activity by providing an increased amount of proximal hydroxide species adjacent to the GEOR active sites. (11,15,64) The enhanced activity is then a result of either the increased hydroxide species providing a more oxidative environment, participating in the GEOR or aiding the oxidative removal of inhibiting COad intermediates. For the electronic effect, the addition of non-noble transition metals to the lattice of noble metals such as Pt and Pd decreases the binding energy for inhibiting species such as CO (27) and OH, (65) and it can increase the number of available surface electrons (66) and lower glycerol adsorption energy; (66) the result being easier desorption of intermediates and, therefore, increased active sites during the GEOR as well as less C–C bond cleavage. From the perspective of this study, it is difficult to discern definitively which process explains the increased performance of PdNi/C over Pd/C, but both processes seem plausible. However, the study conducted on PtNi by Luo et al. (65) provides a detailed body of evidence pertaining to the role of Ni in enhancing the catalytic activity of Pt being a result of the electronic effect.
Note that in the present study, we have avoided kinetic analyses from the Tafel slope as calculating Tafel slopes from the ICPCs for the GEOR is problematic, as the reaction mechanism may vary along the polarization curve. Hence, the calculation of such slopes may be misleading and may not give true indications of a single reaction-limiting step.

Glycerol Oxidation Product Analysis

To analyze the GOPs generated by the PdNi/C and Pd/C catalysts, galvanostatic measurements were conducted in 0.50 M Gly in 1.0 M NaOH solutions to a final charge of 18C (Figures S11 and S12). Although long-term stability is an important aspect of electrocatalytic systems, the goal of this study was to synthesize the (111) crystal facet catalysts and directly compare their activity and product distribution experimentally and computationally close to the GEOR onset potential. That is why, for the experimental determination of the GOPs, a small volume of solution was used, and only 18C were passed such that a readable concentration of products could be determined in a short period of time. Further work on these catalysts in a larger volume cell setup, in which the catalysts can be used for extended periods to assess the stability, is important to be pursued. However, given that the NPs are somewhat spherical, it would be challenging to detect changes to the morphology or structure of such small particles (6 nm for Pd and 3 nm for PdNi).
The product analysis resulting from the galvanostatic measurements is presented as product fractions of the sum of the overall GOP concentrations in Table 1 and Figure 5 as a function of measured potential. A result of the experiments being performed galvanostatically is that the average potential during the experiments slightly differs between runs. Therefore, the measured potential for each experiment was corrected for the iR-drop in the electrochemical cell, and then the average potential, in addition to the standard deviation (SD), was calculated over the entire run. For repeat experiments, the same process was adopted, then a second average and SD were determined for the combined number of experiments. The average and SD of the WE potential are reported in Table 1, which shows that the potentials for the comparison of PdNi/C to Pd/C are not exactly the same but are within the SD. The potentials analyzed were chosen to be close to the GEOR onset because the catalysts become less selective, and C–C bond cleavage is more likely at higher potentials. The GOP concentrations determined by HPLC analysis were normalized by the charge passed and then converted to a fraction.
Table 1. Product Fractions of GOPs and Glycerol Conversion from Galvanostatic Experiments Using Pd/C and PdNi/C in 0.50 M Gly and 1.0 M NaOHa
electrocatalyst, V vs RHEproduct fraction average ± standard deviation (%)Gly conversion (%)
(number of experiments)GLALATTAFAGAOAAA 
PdNi/C, 0.983 ± 0.003 V (1)56273.89.90.92.302.4
PdNi/C, 0.846 ± 0.021 V (3)58 ± 429 ± 25.3 ± 0.61.1 ± 1.22.2 ± 1.73.6 ± 0.60.6 ± 0.41.4 ± 0.5
PdNi/C, 0.721 ± 0.015 V (2)60 ± 130 ± 16.6 ± 0.10.5 ± 0.50 ± 02.1 ± 0.40.6 ± 0.61.3 ± 0.6
PdNi/C, 0.688 ± 0.029 V (2)61 ± 430 ± 45.5 ± 2.01.8 ± 1.80 ± 01.2 ± 0.50.5 ± 0.51.6 ± 0.0
Pd/C, 0.832 ± 0.049 V (2)60 ± 024 ± 05.9 ± 0.15.6 ± 1.02.1 ± 2.11.8 ± 0.41.0 ± 1.01.6 ± 0.1
Pd/C, 0.734 ± 0.020 V (3)62 ± 226 ± 26.3 ± 0.52.0 ± 1.51.4 ± 0.61.4 ± 0.81.6 ± 1.21.9 ± 0.2
Pd/C, 0.672 ± 0.022 V (4)64 ± 224 ± 38.3 ± 1.30.6 ± 0.50.5 ± 0.42.8 ± 0.90.5 ± 0.42.3 ± 0.4
a

18C passed for each measurement.

Figure 5

Figure 5. Product fractions analyzed by HPLC for (a) PdNi/C and (b) Pd/C catalysts at varying potentials.

The product fractions listed in Table 1 show that an increased potential for both PdNi/C and Pd/C results in lower product fractions for GLA and LA. Overall, the two main products, GLA and LA, are formed in similar proportions as a function of potential for both PdNi/C and Pd/C. Additionally, to demonstrate C–C cleavage at higher potentials, one experiment was done at 0.983 V vs RHE for PdNi/C leading to a larger product fraction of FA. Figure 5 graphically illustrates the data presented in Table 1 and shows that, remarkably, there is very little difference in the product fraction distribution of the GOPs for Pd/C and PdNi/C with there being a maximum of 3 and 5% differences between the product fractions for GLA and LA, respectively, observed at similar potentials. The Gly conversion % shown in Table 1 is quite low due to the high concentration of Gly in solution and relatively short electrolysis times evaluated, which were done to minimize the chemical reactions that can occur in alkaline solution.
Figure 5 shows a decrease in the selectivity for GLA as the potential increases, for both PdNi/C and Pd/C. At the lowest examined potentials of 0.688 V vs RHE for PdNi/C and 0.672 V vs RHE for Pd/C, the product selectivity to GLA was the highest, with product fractions of 64 and 61%, respectively. This suggests that the closer the anodic potentials are to the GEOR onset potential and likewise to the PDS potential calculated by DFT (vide infra), the higher the selectivity toward the formation of GLA. At higher potentials, comparing PdNi/C and Pd/C at 0.721 V vs RHE and 0.734 V vs RHE, respectively, the GLA product fraction was 60 and 62% for PdNi/C and Pd/C, respectively. Therefore, under the current experimental conditions, the reaction pathway is not significantly altered by the addition of Ni to Pd. Hence, the PdNi/C demonstrates improved electrochemical performance for the GEOR compared to Pd/C, as discussed earlier, without compromising the selectivity for the predominant GLA product.
The second most abundant product was LA. Figure 5 demonstrates that PdNi/C had a slightly higher selectivity to lactate than Pd/C. However, the overall similar product fractions for the two catalysts further illustrate that the reaction pathways are not significantly altered by alloying Ni with Pd. At the highest potential studied for PdNi/C, 0.983 V vs RHE, the product fractions for GLA and LA were the lowest, and there was an increased selectivity to FA compared to lower potentials. This decrease in selectivity for C3 GOPs at higher potentials has also been reported previously. (67,68) The same increase in the FA product fraction at a higher potential also occurred for Pd/C at 0.832 V vs RHE when compared to lower potentials. Interestingly, there was a lower FA product fraction for PdNi/C at 0.846 V vs RHE than for Pd/C at 0.832 V vs RHE, suggesting a lower affinity for C–C bond cleavage.
In comparison to the present study, reports on Pt (69) and PtNi (65,70) in alkaline media have shown varied results as to the main product generated during the GEOR. On a Pt RDE, Melle et al. (69) showed that for supporting electrolytes of LiOH, NaOH, and KOH, LA and GLA are generated in approximately even amounts close to the GEOR onset potential, but for NaOH, LA formation dominates at higher potentials. For PtNi, Luo et al. (65) reported that close to the onset potential for the GEOR (0.5 V vs RHE), Pt/C and Pt1.2Ni1 supported on carbon paper (ratio of Pt/Ni in at. %) had similar selectivity with approximately 40% of the products assigned to GLA and the remaining 60% being TTA. The authors noted that only low quantities of LA were formed. Furthermore, at a higher potential (0.7 V vs RHE), the selectivity toward GLA dropped considerably for both Pt/C and Pt1.2Ni1 to around 20%, whereas in the present study, an increase of around 200 mV from the lowest studied potential shows no such decline in the product fraction of GLA. In contrast to Luo et al., (65) for Pt and Pt67Ni33 supported by graphene nanosheets under an applied potential close to the GEOR onset, Zhou et al. (70) reported GA to be the dominant product in both cases with around 40% selectivity. Additionally, the authors found GD and GLA present in both cases but with Pt having higher selectivity toward GD and Pt67Ni33 favoring GLA, whereas in the present study with PdNi, GD was not detected, and GA was a minor product.
In summary, the general trends from the product analysis indicate that PdNi/C maintains the selectivity for GLA and exhibits slightly higher selectivity for LA when compared to Pd/C. The selectivity toward GLA decreases only slightly at higher potentials for both Pd/C and PdNi/C. Despite the increased specific activity for the PdNi/C catalyst compared to Pd/C, it is remarkable that the product selectivity is consistently maintained at various potentials. The similar GEOR pathways for the two catalysts were confirmed through DFT calculations.

Density Functional Theory Calculations

Figure 6 shows the computed PDS for GEOR to form common liquid products. All reactions are labeled from 1 to 6. The oxidation of the secondary hydroxyl group on Gly can undergo either pathway 1, forming dihydroxyacetone (DHA) or hydroxypyruvic acid (HPA) as the major product (Gly–DHA–HPA); or pathway 6, forming DHA or LA as the major product (Gly–DHA–LA). The oxidation of the primary hydroxyl group on Gly follows pathway 2, forming glyceraldehyde (GD) or GLA as the major oxidation product (Gly–GD–GLA).

Figure 6

Figure 6. Representation of the calculated Gly electrooxidation pathway of Pd(111) and PdNi(111) catalysts in alkaline solution with PDS values (in eV) of each oxidation step. The oxidation pathway of each intermediate is shown using different colors as follows: (1) Gly–DHA (black), (1′) DHA–HPA (black), (2) Gly–GD (blue), (2′) GD–GLA (blue), (3) GLA–GA (red), (3′) GA–FA (red), (4) GLA–TTA (magenta), (4′) TTA–MLA (magenta), (5) GA–OA (gray), and (6) DHA–LA (green). Here, PDS represents the minimum potential required to overcome the energy barrier of the reactions illustrated in Figures S5–S10.

The Pd and PdNi surfaces require smaller PDS values (Pd: 0.66 eV, PdNi: 0.39 eV) in pathway 1 to form GLA and pathway 6 to form LA (Pd: 0.58 eV, PdNi: 0.36 eV) than following pathway 2 to form HPA (Pd: 0.66 eV, PdNi: 0.39 eV). Table 2 summarizes the PDS required to form the GOPs starting from the Gly molecule. The PDSs of GLA and LA are the lowest values, which justifies the higher product fraction of both species obtained by HPLC. For a close-packed Pd structure (Pd(111)), Valter et al. (71,72) found, using DFT calculations, that the first deprotonation step leading to the adsorption of Gly to Pd favors adsorption by the secondary carbon and that after two deprotonation steps, Gly was bound by both primary and secondary carbon. The authors noted that while this form of adsorption should give either GD or DHA, in experimental studies, the GD pathway was more commonly reported with the subsequent oxidation step, in alkaline conditions, forming GLA. Additionally, the authors found that the first adsorption/deprotonation step for Pt(111) mirrored that of Pd(111) but that the second adsorption step likely favored both primary carbons bound to Pt. This may be a reason for the difference in the product distribution observed in the present study compared to that of Pt and PtNi studies in alkaline media discussed earlier.
Table 2. Calculated PDS Values to Obtain the Liquid Products Are Given in Figure 6a
productsreactionCiPdPdNi
glycerateGly–GD–GLAC30.690.39
lactateGly–DHA–LAC30.580.36
tartronateGly–GD–GLA–TTAC30.660.46
formateGly–GD–GLA–GA–FAC10.690.46
glycolateGly–GD–GLA–GAC20.690.38
oxalateGly–GD–GLA–GA–OAC20.680.41
a

All PDS values are in eV.

The experimental product fraction of LA was considerably lower than that of GLA, which can be explained by two main reasons: (i) LA formation involves a series of chemical steps including C–O cleavage (Figure S10), which are not affected by the applied potential and (ii) the interconversion of GD and DHA leads to the LA route as the minor pathway. For (ii), DHA is unstable under alkaline conditions and can be converted into GD or pyruvaldehyde (PV), resulting in either GLA or LA formation, respectively. These conversions are more likely than HPA formation from DHA since the PDS for HPA formation is higher, (73−75) as shown in Figure 6 for reaction (1′). Although the formation of LA from PV is a known reaction, (76) and the DFT calculations show that PV and LA formation from DHA has a common intermediate (Figure S13), the formation of GD from DHA has been shown to have a higher rate constant than PV formation at elevated temperatures in alkaline conditions. (77) Therefore, it is likely that there is a mix of LA, PV, and GD chemically produced after any electrochemical DHA formation but with a preference for the interconversion of DHA to GD. Hence, GLA formation can be achieved electrochemically from GD in greater proportions than the chemical formation of LA mentioned in (i).
Table 2 presents data showing that, for all of the GOPs experimentally detected, PdNi has a lower PDS than Pd. This observation aligns with the experimental findings, indicating that PdNi/C catalysts exhibit superior electrocatalytic activity compared to Pd/C while maintaining selectivity toward valuable GOPs, such as GLA and LA.

Conclusions

Click to copy section linkSection link copied!

The addition of Ni to Pd to create PdNi NPs has proven to be an effective strategy for improving the performance of the GEOR under alkaline conditions. The novelty of the present study involves the successfully tuned synthesis of pure Pd and alloyed Pd0.45Ni0.55 (at. %) NPs, which are very close to the desired pure Pd(111) and PdNi(111) and resulted in nanocatalysts of a certain morphology and composition that were further investigated as promising materials for the GEOR supported by the DFT modeling. Catalyst fabrication procedures usually focus on the composition or morphology; meanwhile, we attempted to fix both aspects. PdNi/C exhibited significantly enhanced specific and mass activity compared to Pd/C, as analyzed by ICPCs. Generally, reports in the literature will show either specific activity or mass activity, but in our study, PdNi was the superior catalyst in both cases. Additionally, PdNi/C was able to operate at potentials higher than those of Pd/C before deactivation by PdO formation. The DFT calculations for the formation of all experimentally detected GOPs consistently indicated lower PDS values for PdNi than Pd, thereby supporting the electrochemical findings. Both PdNi/C and Pd/C displayed high selectivity toward GLA and LA, with product fractions of around 60 and 30%, respectively. At increased potentials, the selectivity for GLA and LA decreased but remained above 50 and 20%, respectively, with PdNi/C exhibiting slightly higher selectivity for LA than Pd/C. This study combines experimental and computational approaches to unambiguously establish that bimetallic PdNi NPs are promising electrocatalysts for the GEOR, significantly improving the activity compared to monometallic Pd while maintaining the selectivity for valuable GOPs. The number of studies including the quantitative analysis of GOPs using HPLC for PdNi catalysts, especially coupled with DFT calculations, is limited in the literature; hence, the findings of the present study provide new insights for the field.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsaem.3c02789.

  • Experimental methods used for the electrochemical measurements, HPLC standards, description of computational details, and experimental results of galvanostatic measurements (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Irina Terekhina - Department of Materials and Environmental Chemistry, Stockholm University, Stockholm SE-10691, SwedenOrcidhttps://orcid.org/0000-0002-4759-5895
    • Daniel Martín-Yerga - Department of Chemistry, Nanoscience Center, University of Jyväskylä, Jyväskylä FI-40014, FinlandOrcidhttps://orcid.org/0000-0002-9385-7577
    • Lars G. M. Pettersson - FYSIKUM, AlbaNova University Center, Stockholm University, Stockholm SE-106 91, Sweden
    • Mats Johnsson - Department of Materials and Environmental Chemistry, Stockholm University, Stockholm SE-10691, SwedenOrcidhttps://orcid.org/0000-0003-4319-1540
    • Ann Cornell - Department of Chemical Engineering, KTH Royal Institute of Technology, Stockholm SE-100 44, SwedenOrcidhttps://orcid.org/0000-0001-5816-2924
  • Author Contributions

    J.W., I.T., and E.C.d.S. contributed equally. The manuscript was written through contributions from all authors. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

The authors would like to acknowledge the Swedish Foundation for Strategic Research for funding this work through grant no. EM16-0010. The computations were enabled by resources provided by the National Academic Infrastructure for Supercomputing in Sweden (NAISS) and the Swedish National Infrastructure for Computing (SNIC) at the PDC and NSC centers partially funded by the Swedish Research Council through grant agreement nos. 2022-06725 and 2018-05973. D.M.-Y. thanks the Research Council of Finland for financial support (ref. 355569). The authors acknowledge that portions of this manuscript included the Ph.D. thesis written by J.W., entitled From Facets to Flow: The Electrooxidation of Glycerol on Pd-based catalysts (2023), for which J.W. holds the copyright (ISBN: 978-91-8040-752-6).

Abbreviations

Click to copy section linkSection link copied!

CV

cyclic voltammetry

DFT

density functional theory

DHA

dihydroxyacetone

ECSA

electrochemical surface area

FA

formic acid/formate

GA

glycolic acid/glycolate

GD

glyceraldehyde

GEOR

glycerol electrooxidation reaction

GLA

glyceric acid/glycerate

Gly

glycerol

GOPs

glycerol oxidation products

HPA

hydroxypyruvic acid

ICPCs

iR-corrected polarization curves

LA

lactic acid

MLA

mesoxalic acid

NPs

nanoparticles

OA

oxalic acid/oxalate

PDS

potential determining step

PXRD

powder X-ray diffraction

RHE

reversible hydrogen electrode

TEM

transmission electron microscopy

TTA

tartronic acid/tartronate

References

Click to copy section linkSection link copied!

This article references 77 other publications.

  1. 1
    Mahlia, T. M. I.; Syazmi, Z. A. H. S.; Mofijur, M.; Abas, A. P.; Bilad, M. R.; Ong, H. C.; Silitonga, A. S. Patent Landscape Review on Biodiesel Production: Technology Updates. Renewable Sustainable Energy Rev. 2020, 118, 109526,  DOI: 10.1016/j.rser.2019.109526
  2. 2
    Likhanov, V. A.; Lopatin, O. P. Research of High-Speed Diesel Engines of Small Dimension on Biofuel. J. Phys.: Conf. Ser. 2019, 1399 (5), 055016,  DOI: 10.1088/1742-6596/1399/5/055016
  3. 3
    Biodiesel: From Production to Combustion; Tabatabaei, M., Aghbashlo, M., Eds.; Biofuel and Biorefinery Technologies; Springer International Publishing: Cham, 2019; Vol. 8. DOI: 10.1007/978-3-030-00985-4 .
  4. 4
    Ciriminna, R.; Pina, C. D.; Rossi, M.; Pagliaro, M. Understanding the Glycerol Market. Eur. J. Lipid Sci. Technol. 2014, 116 (10), 14321439,  DOI: 10.1002/ejlt.201400229
  5. 5
    Coutanceau, C.; Baranton, S.; Kouamé, R. S. B. Selective Electrooxidation of Glycerol Into Value-Added Chemicals: A Short Overview. Front. Chem. 2019, 7, 100,  DOI: 10.3389/fchem.2019.00100
  6. 6
    Simões, M.; Baranton, S.; Coutanceau, C. Electrochemical Valorisation of Glycerol. ChemSusChem 2012, 5 (11), 21062124,  DOI: 10.1002/cssc.201200335
  7. 7
    Houache, M. S. E.; Hughes, K.; Baranova, E. A. Study on Catalyst Selection for Electrochemical Valorization of Glycerol. Sustainable Energy Fuels 2019, 3 (8), 18921915,  DOI: 10.1039/C9SE00108E
  8. 8
    Holade, Y.; Tuleushova, N.; Tingry, S.; Servat, K.; Napporn, T. W.; Guesmi, H.; Cornu, D.; Kokoh, K. B. Recent Advances in the Electrooxidation of Biomass-Based Organic Molecules for Energy, Chemicals and Hydrogen Production. Catal. Sci. Technol. 2020, 10 (10), 30713112,  DOI: 10.1039/C9CY02446H
  9. 9
    Wu, J.; Yang, X.; Gong, M. Recent Advances in Glycerol Valorization via Electrooxidation: Catalyst, Mechanism and Device. Chin. J. Catal. 2022, 43 (12), 29662986,  DOI: 10.1016/S1872-2067(22)64121-4
  10. 10
    Ge, R.; Li, J.; Duan, H. Recent Advances in Non-Noble Electrocatalysts for Oxidative Valorization of Biomass Derivatives. Sci. China Mater. 2022, 65 (12), 32733301,  DOI: 10.1007/s40843-022-2076-y
  11. 11
    Simões, M.; Baranton, S.; Coutanceau, C. Electro-Oxidation of Glycerol at Pd Based Nano-Catalysts for an Application in Alkaline Fuel Cells for Chemicals and Energy Cogeneration. Appl. Catal., B 2010, 93 (3–4), 354362,  DOI: 10.1016/j.apcatb.2009.10.008
  12. 12
    Chen, X.; Granda-Marulanda, L. P.; McCrum, I. T.; Koper, M. T. M. How Palladium Inhibits CO Poisoning during Electrocatalytic Formic Acid Oxidation and Carbon Dioxide Reduction. Nat. Commun. 2022, 13 (1), 38,  DOI: 10.1038/s41467-021-27793-5
  13. 13
    Oliveira, V. L.; Morais, C.; Servat, K.; Napporn, T. W.; Tremiliosi-Filho, G.; Kokoh, K. B. Glycerol Oxidation on Nickel Based Nanocatalysts in Alkaline Medium - Identification of the Reaction Products. J. Electroanal. Chem. 2013, 703, 5662,  DOI: 10.1016/j.jelechem.2013.05.021
  14. 14
    Holade, Y.; Morais, C.; Arrii-Clacens, S.; Servat, K.; Napporn, T. W.; Kokoh, K. B. New Preparation of PdNi/C and PdAg/C Nanocatalysts for Glycerol Electrooxidation in Alkaline Medium. Electrocatalysis 2013, 4 (3), 167178,  DOI: 10.1007/s12678-013-0138-1
  15. 15
    Holade, Y.; Morais, C.; Servat, K.; Napporn, T. W.; Kokoh, K. B. Toward the Electrochemical Valorization of Glycerol: Fourier Transform Infrared Spectroscopic and Chromatographic Studies. ACS Catal. 2013, 3 (10), 24032411,  DOI: 10.1021/cs400559d
  16. 16
    White, J.; Anil, A.; Martín-Yerga, D.; Salazar-Alvarez, G.; Henriksson, G.; Cornell, A. Electrodeposited PdNi on a Ni Rotating Disk Electrode Highly Active for Glycerol Electrooxidation in Alkaline Conditions. Electrochim. Acta 2022, 403, 139714,  DOI: 10.1016/j.electacta.2021.139714
  17. 17
    Koper, M. T. M. Structure Sensitivity and Nanoscale Effects in Electrocatalysis. Nanoscale 2011, 3 (5), 20542073,  DOI: 10.1039/c0nr00857e
  18. 18
    Li, R.; Wei, Z.; Huang, T.; Yu, A. Ultrasonic-Assisted Synthesis of Pd-Ni Alloy Catalysts Supported on Multi-Walled Carbon Nanotubes for Formic Acid Electrooxidation. Electrochim. Acta 2011, 56 (19), 68606865,  DOI: 10.1016/j.electacta.2011.05.097
  19. 19
    Du, C.; Chen, M.; Wang, W.; Yin, G. Nanoporous PdNi Alloy Nanowires As Highly Active Catalysts for the Electro-Oxidation of Formic Acid. ACS Appl. Mater. Interfaces 2011, 3 (2), 105109,  DOI: 10.1021/am100803d
  20. 20
    López-Coronel, A.; Torres-Pacheco, L. J.; Bañuelos, J. A.; Álvarez-López, A.; Guerra-Balcázar, M.; Álvarez-Contreras, L.; Arjona, N. Highly Active PdNi Bimetallic Nanocubes Electrocatalysts for the Ethylene Glycol Electro-Oxidation in Alkaline Medium. Appl. Surf. Sci. 2020, 530, 147210,  DOI: 10.1016/j.apsusc.2020.147210
  21. 21
    Guo, M.; Wang, H.; Cui, L.; Zhang, J.; Xiang, Y.; Lu, S. Nickel Promoted Palladium Nanoparticles for Electrocatalysis of Carbohydrazide Oxidation Reaction. Small 2019, 15 (28), 1900929,  DOI: 10.1002/smll.201900929
  22. 22
    Shen, S. Y.; Zhao, T. S.; Xu, J. B.; Li, Y. S. Synthesis of PdNi Catalysts for the Oxidation of Ethanol in Alkaline Direct Ethanol Fuel Cells. J. Power Sources 2010, 195 (4), 10011006,  DOI: 10.1016/j.jpowsour.2009.08.079
  23. 23
    Miao, B.; Wu, Z.-P.; Zhang, M.; Chen, Y.; Wang, L. Role of Ni in Bimetallic PdNi Catalysts for Ethanol Oxidation Reaction. J. Phys. Chem. C 2018, 122 (39), 2244822459,  DOI: 10.1021/acs.jpcc.8b05812
  24. 24
    Yang, H.; Wang, H.; Li, H.; Ji, S.; Davids, M. W.; Wang, R. Effect of Stabilizers on the Synthesis of Palladium-Nickel Nanoparticles Supported on Carbon for Ethanol Oxidation in Alkaline Medium. J. Power Sources 2014, 260, 1218,  DOI: 10.1016/j.jpowsour.2014.02.110
  25. 25
    Dutta, A.; Datta, J. Energy Efficient Role of Ni/NiO in PdNi Nano Catalyst Used in Alkaline DEFC. J. Mater. Chem. A 2014, 2 (9), 32373250,  DOI: 10.1039/c3ta12708g
  26. 26
    Zhao, Y.; Yang, X.; Tian, J.; Wang, F.; Zhan, L. Methanol Electro-Oxidation on Ni@Pd Core-Shell Nanoparticles Supported on Multi-Walled Carbon Nanotubes in Alkaline Media. Int. J. Hydrogen Energy 2010, 35 (8), 32493257,  DOI: 10.1016/j.ijhydene.2010.01.112
  27. 27
    Araujo, R. B.; Martín-Yerga, D.; Santos, E. C.; Cornell, A.; Pettersson, L. G. M. Elucidating the Role of Ni to Enhance the Methanol Oxidation Reaction on Pd Electrocatalysts. Electrochim. Acta 2020, 360, 136954,  DOI: 10.1016/j.electacta.2020.136954
  28. 28
    Liu, Z.; Zhang, X.; Hong, L. Physical and Electrochemical Characterizations of Nanostructured Pd/C and PdNi/C Catalysts for Methanol Oxidation. Electrochem. Commun. 2009, 11 (4), 925928,  DOI: 10.1016/j.elecom.2009.02.030
  29. 29
    Martín-Yerga, D.; Yu, X.; Terekhina, I.; Henriksson, G.; Cornell, A. In Situ Catalyst Reactivation for Enhancing Alcohol Electro-Oxidation and Coupled Hydrogen Generation. Chem. Commun. 2020, 56 (28), 40114014,  DOI: 10.1039/D0CC01321H
  30. 30
    Martín-Yerga, D.; White, J.; Henriksson, G.; Cornell, A. Structure-Reactivity Effects of Biomass-Based Hydroxyacids for Sustainable Electrochemical Hydrogen Production. ChemSusChem 2021, 14 (8), 19021912,  DOI: 10.1002/cssc.202100073
  31. 31
    Ipadeola, A. K.; Lisa Mathebula, N. Z.; Pagliaro, M. V.; Miller, H. A.; Vizza, F.; Davies, V.; Jia, Q.; Marken, F.; Ozoemena, K. I. Unmasking the Latent Passivating Roles of Ni(OH)2 on the Performance of Pd-Ni Electrocatalysts for Alkaline Ethanol Fuel Cells. ACS Appl. Energy Mater. 2020, 3 (9), 87868802,  DOI: 10.1021/acsaem.0c01314
  32. 32
    del Rosario, J. A. D.; Ocon, J. D.; Jeon, H.; Yi, Y.; Lee, J. K.; Lee, J. Enhancing Role of Nickel in the Nickel-Palladium Bilayer for Electrocatalytic Oxidation of Ethanol in Alkaline Media. J. Phys. Chem. C 2014, 118 (39), 2247322478,  DOI: 10.1021/jp411601c
  33. 33
    White, J.; Peters, L.; Martín-Yerga, D.; Terekhina, I.; Anil, A.; Lundberg, H.; Johnsson, M.; Salazar-Alvarez, G.; Henriksson, G.; Cornell, A. Glycerol Electrooxidation at Industrially Relevant Current Densities Using Electrodeposited PdNi/Nifoam Catalysts in Aerated Alkaline Media. J. Electrochem. Soc. 2023, 170 (8), 086504,  DOI: 10.1149/1945-7111/acee27
  34. 34
    Hiltrop, D.; Cychy, S.; Elumeeva, K.; Schuhmann, W.; Muhler, M. Spectroelectrochemical Studies on the Effect of Cations in the Alkaline Glycerol Oxidation Reaction over Carbon Nanotube-Supported Pd Nanoparticles. Beilstein J. Org. Chem. 2018, 14 (1), 14281435,  DOI: 10.3762/bjoc.14.120
  35. 35
    Rahim, S. A. N. M.; Lee, C. S.; Abnisa, F.; Aroua, M. K.; Daud, W. A. W.; Cognet, P.; Pérès, Y. A Review of Recent Developments on Kinetics Parameters for Glycerol Electrochemical Conversion - A by-Product of Biodiesel. Sci. Total Environ. 2020, 705, 135137,  DOI: 10.1016/j.scitotenv.2019.135137
  36. 36
    Rodrigues, T. S.; Zhao, M.; Yang, T.-H.; Gilroy, K. D.; da Silva, A. G. M.; Camargo, P. H. C.; Xia, Y. Synthesis of Colloidal Metal Nanocrystals: A Comprehensive Review on the Reductants. Chem.─Eur. J. 2018, 24 (64), 1694416963,  DOI: 10.1002/chem.201802194
  37. 37
    Poerwoprajitno, A. R.; Gloag, L.; Cheong, S.; Gooding, J. J.; Tilley, R. D. Synthesis of Low- and High-Index Faceted Metal (Pt, Pd, Ru, Ir, Rh) Nanoparticles for Improved Activity and Stability in Electrocatalysis. Nanoscale 2019, 11 (41), 1899519011,  DOI: 10.1039/C9NR05802H
  38. 38
    Gilroy, K. D.; Ruditskiy, A.; Peng, H.-C.; Qin, D.; Xia, Y. Bimetallic Nanocrystals: Syntheses, Properties, and Applications. Chem. Rev. 2016, 116 (18), 1041410472,  DOI: 10.1021/acs.chemrev.6b00211
  39. 39
    Göksu, H.; Ho, S. F.; Metin, O. ̈.; Korkmaz, K.; Mendoza Garcia, A.; Gültekin, M. S.; Sun, S. Tandem Dehydrogenation of Ammonia Borane and Hydrogenation of Nitro/Nitrile Compounds Catalyzed by Graphene-Supported NiPd Alloy Nanoparticles. ACS Catal. 2014, 4 (6), 17771782,  DOI: 10.1021/cs500167k
  40. 40
    Hummelgård, C.; Karlsson, R. K. B.; Bäckström, J.; Rahman, S. M. H.; Cornell, A.; Eriksson, S.; Olin, H. Physical and Electrochemical Properties of Cobalt Doped (Ti,Ru)O2 Electrode Coatings. Mater. Sci. Eng. B 2013, 178 (20), 15151522,  DOI: 10.1016/j.mseb.2013.08.018
  41. 41
    Bolzán, A. Phenomenological Aspects Related to the Electrochemical Behaviour of Smooth Palladium Electrodes in Alkaline Solutions. J. Electroanal. Chem. 1995, 380 (1–2), 127138,  DOI: 10.1016/0022-0728(94)03627-F
  42. 42
    Breiter, M. W. Dissolution and Adsorption of Hydrogen at Smooth Pd Wires of the Alpha Phase in Sulfuric Acid Solution. J. Electroanal. Chem. 1977, 81 (2), 275284,  DOI: 10.1016/S0022-0728(77)80023-5
  43. 43
    Chierchie, T.; Mayer, C.; Lorenz, W. J. Structural Changes of Surface Oxide Layers on Palladium. J. Electroanal. Chem. 1982, 135, 211220,  DOI: 10.1016/0368-1874(82)85121-6
  44. 44
    Hjorth Larsen, A.; Jørgen Mortensen, J.; Blomqvist, J.; Castelli, I. E.; Christensen, R.; Dułak, M.; Friis, J.; Groves, M. N.; Hammer, B.; Hargus, C.; Hermes, E. D.; Jennings, P. C.; Bjerre Jensen, P.; Kermode, J.; Kitchin, J. R.; Leonhard Kolsbjerg, E.; Kubal, J.; Kaasbjerg, K.; Lysgaard, S.; Bergmann Maronsson, J.; Maxson, T.; Olsen, T.; Pastewka, L.; Peterson, A.; Rostgaard, C.; Schiøtz, J.; Schütt, O.; Strange, M.; Thygesen, K. S.; Vegge, T.; Vilhelmsen, L.; Walter, M.; Zeng, Z.; Jacobsen, K. W. The Atomic Simulation Environment-a Python Library for Working with Atoms. J. Phys.: Condens. Matter 2017, 29 (27), 273002,  DOI: 10.1088/1361-648X/aa680e
  45. 45
    Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54 (16), 1116911186,  DOI: 10.1103/PhysRevB.54.11169
  46. 46
    Wellendorff, J.; Lundgaard, K. T.; Møgelhøj, A.; Petzold, V.; Landis, D. D.; Nørskov, J. K.; Bligaard, T.; Jacobsen, K. W. Density Functionals for Surface Science: Exchange-Correlation Model Development with Bayesian Error Estimation. Phys. Rev. B 2012, 85 (23), 235149,  DOI: 10.1103/PhysRevB.85.235149
  47. 47
    Anil, A.; White, J.; Campos dos Santos, E.; Terekhina, I.; Johnsson, M.; Pettersson, L. G. M.; Cornell, A.; Salazar-Alvarez, G. Effect of Pore Mesostructure on the Electrooxidation of Glycerol on Pt Mesoporous Catalysts. J. Mater. Chem. A 2023, 11 (31), 1657016577,  DOI: 10.1039/D3TA01738A
  48. 48
    dos Santos, E. C.; Araujo, R. B.; Valter, M.; Salazar-Alvarez, G.; Johnsson, M.; Bajdich, M.; Abild-Pedersen, F.; Pettersson, L. G. M. Efficient Screening of Bi-Metallic Electrocatalysts for Glycerol Valorization. Electrochim. Acta 2021, 398, 139283,  DOI: 10.1016/j.electacta.2021.139283
  49. 49
    T. C., Allison. NIST-JANAF Thermochemical Tables - SRD 13, 2013.  DOI: 10.18434/T42S31 .
  50. 50
    Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, T. A.; Hennig, R. G. Implicit Solvation Model for Density-Functional Study of Nanocrystal Surfaces and Reaction Pathways. J. Chem. Phys. 2014, 140 (8), 084106,  DOI: 10.1063/1.4865107
  51. 51
    Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J. R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108 (46), 1788617892,  DOI: 10.1021/jp047349j
  52. 52
    Sen, B.; Kuzu, S.; Demir, E.; Yıldırır, E.; Sen, F. Highly Efficient Catalytic Dehydrogenation of Dimethyl Ammonia Borane via Monodisperse Palladium-Nickel Alloy Nanoparticles Assembled on PEDOT. Int. J. Hydrogen Energy 2017, 42 (36), 2330723314,  DOI: 10.1016/j.ijhydene.2017.05.115
  53. 53
    Matin, M. A.; Jang, J.-H.; Kwon, Y.-U. PdM Nanoparticles (M = Ni, Co, Fe, Mn) with High Activity and Stability in Formic Acid Oxidation Synthesized by Sonochemical Reactions. J. Power Sources 2014, 262, 356363,  DOI: 10.1016/j.jpowsour.2014.03.109
  54. 54
    Khan, M. S.; Khattak, R.; Khan, A.; Chen, Q.; Nisar, J.; Iqbal, Z.; Rashid, A.; Kamran, A. W.; Zekker, I.; Zahoor, M.; Alzahrani, K. J.; Batiha, G. E.-S. Synthesis and Characterizations of PdNi Carbon Supported Nanomaterials: Studies of Electrocatalytic Activity for Oxygen Reduction in Alkaline Medium. Molecules 2021, 26 (11), 3440,  DOI: 10.3390/molecules26113440
  55. 55
    Wang, L.; Dong, K. J.; Wang, C. C.; Zou, R. P.; Zhou, Z. Y.; Yu, A. B. Computer Simulation of the Packing of Nanoparticles. Powder Technol. 2022, 401, 117317,  DOI: 10.1016/j.powtec.2022.117317
  56. 56
    Eldridge, M. D.; Madden, P. A.; Frenkel, D. Entropy-Driven Formation of a Superlattice in a Hard-Sphere Binary Mixture. Nature 1993, 365 (6441), 3537,  DOI: 10.1038/365035a0
  57. 57
    Wang, Z.; Schliehe, C.; Bian, K.; Dale, D.; Bassett, W. A.; Hanrath, T.; Klinke, C.; Weller, H. Correlating Superlattice Polymorphs to Internanoparticle Distance, Packing Density, and Surface Lattice in Assemblies of PbS Nanoparticles. Nano Lett. 2013, 13 (3), 13031311,  DOI: 10.1021/nl400084k
  58. 58
    Bouju, X.; Duguet, É.; Gauffre, F.; Henry, C. R.; Kahn, M. L.; Mélinon, P.; Ravaine, S. Nonisotropic Self-Assembly of Nanoparticles: From Compact Packing to Functional Aggregates. Adv. Mater. 2018, 30 (27), 1706558,  DOI: 10.1002/adma.201706558
  59. 59
    Wang, L.; Lavacchi, A.; Bellini, M.; D’Acapito, F.; Benedetto, F. D.; Innocenti, M.; Miller, H. A.; Montegrossi, G.; Zafferoni, C.; Vizza, F. Deactivation of Palladium Electrocatalysts for Alcohols Oxidation in Basic Electrolytes. Electrochim. Acta 2015, 177, 100106,  DOI: 10.1016/j.electacta.2015.02.026
  60. 60
    Daily Metal Price: Palladium Price (USD/Gram) Chart for the Last Week. https://www.dailymetalprice.com/metalpricecharts.php?c=pd&u=oz&d=240 (accessed 10 16, 2023).
  61. 61
    Daily Metal Price: Nickel Price (USD/Gram) Chart for the Last Week. https://www.dailymetalprice.com/metalpricecharts.php?c=ni&u=lb&d=0 (accessed 10 16, 2023).
  62. 62
    Oliveira, V. L.; Morais, C.; Servat, K.; Napporn, T. W.; Tremiliosi-Filho, G.; Kokoh, K. B. Glycerol Oxidation on Nickel Based Nanocatalysts in Alkaline Medium - Identification of the Reaction Products. J. Electroanal. Chem. 2013, 703, 56,  DOI: 10.1016/j.jelechem.2013.05.021
  63. 63
    Goetz, M. K.; Bender, M. T.; Choi, K.-S. Predictive Control of Selective Secondary Alcohol Oxidation of Glycerol on NiOOH. Nat. Commun. 2022, 13 (1), 5848,  DOI: 10.1038/s41467-022-33637-7
  64. 64
    Zhang, N.; Wang, J.; Zhang, W.; Zhao, Y.; Dong, Z.; Wu, Z.; Xu, G.-R.; Wang, L. Self-Supported PdNi Dendrite on Ni Foam for Improving Monohydric Alcohol and Polyhydric Alcohols Electrooxidation. Fuel 2022, 326, 125083,  DOI: 10.1016/j.fuel.2022.125083
  65. 65
    Luo, H.; Yukuhiro, V. Y.; Fernández, P. S.; Feng, J.; Thompson, P.; Rao, R. R.; Cai, R.; Favero, S.; Haigh, S. J.; Durrant, J. R.; Stephens, I. E. L.; Titirici, M.-M. Role of Ni in PtNi Bimetallic Electrocatalysts for Hydrogen and Value-Added Chemicals Coproduction via Glycerol Electrooxidation. ACS Catal. 2022, 12 (23), 1449214506,  DOI: 10.1021/acscatal.2c03907
  66. 66
    Mphahlele, N. E.; Ipadeola, A. K.; Haruna, A. B.; Mwonga, P. V.; Modibedi, R. M.; Palaniyandy, N.; Billing, C.; Ozoemena, K. I. Microwave-induced Defective PdFe/C Nano-electrocatalyst for Highly Efficient Alkaline Glycerol Oxidation Reactions. Electrochim. Acta 2022, 409, 139977,  DOI: 10.1016/j.electacta.2022.139977
  67. 67
    Houache, M. S. E.; Shubair, A.; Sandoval, M. G.; Safari, R.; Botton, G. A.; Jasen, P. V.; González, E. A.; Baranova, E. A. Influence of Pd and Au on Electrochemical Valorization of Glycerol over Ni-Rich Surfaces. J. Catal. 2021, 396, 113,  DOI: 10.1016/j.jcat.2021.02.008
  68. 68
    Wang, H.; Thia, L.; Li, N.; Ge, X.; Liu, Z.; Wang, X. Pd Nanoparticles on Carbon Nitride-Graphene for the Selective Electro-Oxidation of Glycerol in Alkaline Solution. ACS Catal. 2015, 5 (6), 31743180,  DOI: 10.1021/acscatal.5b00183
  69. 69
    Melle, G.; de Souza, M. B. C.; Santiago, P. V. B.; Corradini, P. G.; Mascaro, L. H.; Fernández, P. S.; Sitta, E. Glycerol Electro-Oxidation at Pt in Alkaline Media: Influence of Mass Transport and Cations. Electrochim. Acta 2021, 398, 139318,  DOI: 10.1016/j.electacta.2021.139318
  70. 70
    Zhou, Y.; Shen, Y.; Piao, J. Sustainable Conversion of Glycerol into Value-Added Chemicals by Selective Electro-Oxidation on Pt-Based Catalysts. ChemElectroChem 2018, 5 (13), 16361643,  DOI: 10.1002/celc.201800309
  71. 71
    Valter, M.; dos Santos, E. C.; Pettersson, L. G. M.; Hellman, A. Partial Electrooxidation of Glycerol on Close-Packed Transition Metal Surfaces: Insights from First-Principles Calculations. J. Phys. Chem. C 2020, 124 (33), 1790717915,  DOI: 10.1021/acs.jpcc.0c04002
  72. 72
    Valter, M.; Santos, E. C.; Pettersson, L. G. M.; Hellman, A. Selectivity of the First Two Glycerol Dehydrogenation Steps Determined Using Scaling Relationships. ACS Catal. 2021, 11 (6), 34873497,  DOI: 10.1021/acscatal.0c04186
  73. 73
    Yu, X.; dos Santos, E. C.; White, J.; Salazar-Alvarez, G.; Pettersson, L. G. M.; Cornell, A.; Johnsson, M. Electrocatalytic Glycerol Oxidation with Concurrent Hydrogen Evolution Utilizing an Efficient MoOx/Pt Catalyst. Small 2021, 17 (44), 2104288,  DOI: 10.1002/smll.202104288
  74. 74
    Yaylayan, V.; Harty-Majors, S.; Ismail, A. Investigation of Dl-Glyceraldehyde-Dihydroxyacetone Interconversion by FTIR Spectroscopy. Carbohydr. Res. 1999, 318 (1–4), 2025,  DOI: 10.1016/S0008-6215(99)00077-4
  75. 75
    Rasrendra, C. B.; Fachri, B. A.; Makertihartha, I. G. B. N.; Adisasmito, S.; Heeres, H. J. Catalytic Conversion of Dihydroxyacetone to Lactic Acid Using Metal Salts in Water. ChemSusChem 2011, 4 (6), 768777,  DOI: 10.1002/cssc.201000457
  76. 76
    Denis, W. On the Behaviour of Various Aldehydes, Ketones and Alcohols towards Oxidizing Agents. Am. Chem. J. 1907, 38, 561594
  77. 77
    Fedoroňko, M.; Königstein, J. Kinetics of Mutual Isomerization of Trioses and Their Dehydration to Methylglyoxal. Collect. Czech. Chem. Commun. 1969, 34, 38813894,  DOI: 10.1135/cccc19693881

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 4 publications.

  1. Irina Terekhina, Mats Johnsson. Improving Glycerol Electrooxidation Performance on Nanocubic PtCo Catalysts. ACS Applied Materials & Interfaces 2024, 16 (42) , 56987-56996. https://doi.org/10.1021/acsami.4c10219
  2. De‐Xin Tan, Yan‐li Wang, Xiao‐Mei Ning, Xiao‐Li Luo, Chun‐Ru Chen, Qiao Xiao, Jian‐Hong Chen. One‐dimensional Branched PdAg Nanoalloy Rich in Defects for Boosting Glycerol and Ethanol Oxidation. Chemistry – A European Journal 2025, 31 (15) https://doi.org/10.1002/chem.202404172
  3. Junhao Wu, Xiao Zhang, Sijia Ren, Xinhui Lu, Jiaxin Yang, Kui Li. Employing a MoO 2 @NiO heterojunction as a highly selective and efficient electrochemical ethanol-to-acetaldehyde conversion catalyst. CrystEngComm 2024, 26 (47) , 6701-6706. https://doi.org/10.1039/D4CE01039F
  4. Yang Zhang, Lin Wang, Shengmin Pan, Lin Zhou, Man Zhang, Yaoyue Yang, Wenbin Cai. Improving the Electrochemical Glycerol-to-Glycerate Conversion at Pd Sites via the Interfacial Hydroxyl Immigrated from Ni Sites. Molecules 2024, 29 (16) , 3890. https://doi.org/10.3390/molecules29163890

ACS Applied Energy Materials

Cite this: ACS Appl. Energy Mater. 2024, 7, 5, 1802–1813
Click to copy citationCitation copied!
https://doi.org/10.1021/acsaem.3c02789
Published February 29, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY 4.0 .

Article Views

1660

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Scheme 1

    Scheme 1. Schematic Representation of the Pd and PdNi Nanoparticle Synthesis

    Figure 1

    Figure 1. TEM and HRTEM images of (a,b) Pd NPs and (c,d) PdNi NPs. The insets of images (a,b) show particle size distributions of the corresponding catalysts. (e) PXRD patterns of Pd and PdNi NPs.

    Figure 2

    Figure 2. (a) CV in 1.0 M NaOH for Pd/C and PdNi/C. (b) LSV in 1.0 M NaOH and 0.50 M Gly for Pd/C and PdNi/C along with the CV from (a). (c) LSV from (b) normalized by ECSA. (d) LSV from (b) normalized by the mass of Pd in the Pd/C and PdNi/C catalysts. All sweeps were conducted at 25 °C and a sweep rate of 50 mV s–1.

    Figure 3

    Figure 3. ICPCs for PdNi/C at 40 °C in a supporting electrolyte of 1.0 M NaOH with Gly concentrations of 0.10 M (blue triangles) and 0.50 M (red circles). (a) No normalization, i.e., current; (b) specific activity, i.e., normalized by ECSA; and (c) mass activity, i.e., normalized by the mass of Pd in the PdNi/C catalyst.

    Figure 4

    Figure 4. ICPCs for Pd/C (black squares) and PdNi/C (red circles) at 40 °C in a 1.0 M NaOH and 0.50 M Gly solution. (a) No normalization, i.e., current; (b) specific activity, i.e., normalized by ECSA; and (c) mass activity, i.e., normalized by the mass of Pd in the Pd/C and PdNi/C catalysts.

    Figure 5

    Figure 5. Product fractions analyzed by HPLC for (a) PdNi/C and (b) Pd/C catalysts at varying potentials.

    Figure 6

    Figure 6. Representation of the calculated Gly electrooxidation pathway of Pd(111) and PdNi(111) catalysts in alkaline solution with PDS values (in eV) of each oxidation step. The oxidation pathway of each intermediate is shown using different colors as follows: (1) Gly–DHA (black), (1′) DHA–HPA (black), (2) Gly–GD (blue), (2′) GD–GLA (blue), (3) GLA–GA (red), (3′) GA–FA (red), (4) GLA–TTA (magenta), (4′) TTA–MLA (magenta), (5) GA–OA (gray), and (6) DHA–LA (green). Here, PDS represents the minimum potential required to overcome the energy barrier of the reactions illustrated in Figures S5–S10.

  • References


    This article references 77 other publications.

    1. 1
      Mahlia, T. M. I.; Syazmi, Z. A. H. S.; Mofijur, M.; Abas, A. P.; Bilad, M. R.; Ong, H. C.; Silitonga, A. S. Patent Landscape Review on Biodiesel Production: Technology Updates. Renewable Sustainable Energy Rev. 2020, 118, 109526,  DOI: 10.1016/j.rser.2019.109526
    2. 2
      Likhanov, V. A.; Lopatin, O. P. Research of High-Speed Diesel Engines of Small Dimension on Biofuel. J. Phys.: Conf. Ser. 2019, 1399 (5), 055016,  DOI: 10.1088/1742-6596/1399/5/055016
    3. 3
      Biodiesel: From Production to Combustion; Tabatabaei, M., Aghbashlo, M., Eds.; Biofuel and Biorefinery Technologies; Springer International Publishing: Cham, 2019; Vol. 8. DOI: 10.1007/978-3-030-00985-4 .
    4. 4
      Ciriminna, R.; Pina, C. D.; Rossi, M.; Pagliaro, M. Understanding the Glycerol Market. Eur. J. Lipid Sci. Technol. 2014, 116 (10), 14321439,  DOI: 10.1002/ejlt.201400229
    5. 5
      Coutanceau, C.; Baranton, S.; Kouamé, R. S. B. Selective Electrooxidation of Glycerol Into Value-Added Chemicals: A Short Overview. Front. Chem. 2019, 7, 100,  DOI: 10.3389/fchem.2019.00100
    6. 6
      Simões, M.; Baranton, S.; Coutanceau, C. Electrochemical Valorisation of Glycerol. ChemSusChem 2012, 5 (11), 21062124,  DOI: 10.1002/cssc.201200335
    7. 7
      Houache, M. S. E.; Hughes, K.; Baranova, E. A. Study on Catalyst Selection for Electrochemical Valorization of Glycerol. Sustainable Energy Fuels 2019, 3 (8), 18921915,  DOI: 10.1039/C9SE00108E
    8. 8
      Holade, Y.; Tuleushova, N.; Tingry, S.; Servat, K.; Napporn, T. W.; Guesmi, H.; Cornu, D.; Kokoh, K. B. Recent Advances in the Electrooxidation of Biomass-Based Organic Molecules for Energy, Chemicals and Hydrogen Production. Catal. Sci. Technol. 2020, 10 (10), 30713112,  DOI: 10.1039/C9CY02446H
    9. 9
      Wu, J.; Yang, X.; Gong, M. Recent Advances in Glycerol Valorization via Electrooxidation: Catalyst, Mechanism and Device. Chin. J. Catal. 2022, 43 (12), 29662986,  DOI: 10.1016/S1872-2067(22)64121-4
    10. 10
      Ge, R.; Li, J.; Duan, H. Recent Advances in Non-Noble Electrocatalysts for Oxidative Valorization of Biomass Derivatives. Sci. China Mater. 2022, 65 (12), 32733301,  DOI: 10.1007/s40843-022-2076-y
    11. 11
      Simões, M.; Baranton, S.; Coutanceau, C. Electro-Oxidation of Glycerol at Pd Based Nano-Catalysts for an Application in Alkaline Fuel Cells for Chemicals and Energy Cogeneration. Appl. Catal., B 2010, 93 (3–4), 354362,  DOI: 10.1016/j.apcatb.2009.10.008
    12. 12
      Chen, X.; Granda-Marulanda, L. P.; McCrum, I. T.; Koper, M. T. M. How Palladium Inhibits CO Poisoning during Electrocatalytic Formic Acid Oxidation and Carbon Dioxide Reduction. Nat. Commun. 2022, 13 (1), 38,  DOI: 10.1038/s41467-021-27793-5
    13. 13
      Oliveira, V. L.; Morais, C.; Servat, K.; Napporn, T. W.; Tremiliosi-Filho, G.; Kokoh, K. B. Glycerol Oxidation on Nickel Based Nanocatalysts in Alkaline Medium - Identification of the Reaction Products. J. Electroanal. Chem. 2013, 703, 5662,  DOI: 10.1016/j.jelechem.2013.05.021
    14. 14
      Holade, Y.; Morais, C.; Arrii-Clacens, S.; Servat, K.; Napporn, T. W.; Kokoh, K. B. New Preparation of PdNi/C and PdAg/C Nanocatalysts for Glycerol Electrooxidation in Alkaline Medium. Electrocatalysis 2013, 4 (3), 167178,  DOI: 10.1007/s12678-013-0138-1
    15. 15
      Holade, Y.; Morais, C.; Servat, K.; Napporn, T. W.; Kokoh, K. B. Toward the Electrochemical Valorization of Glycerol: Fourier Transform Infrared Spectroscopic and Chromatographic Studies. ACS Catal. 2013, 3 (10), 24032411,  DOI: 10.1021/cs400559d
    16. 16
      White, J.; Anil, A.; Martín-Yerga, D.; Salazar-Alvarez, G.; Henriksson, G.; Cornell, A. Electrodeposited PdNi on a Ni Rotating Disk Electrode Highly Active for Glycerol Electrooxidation in Alkaline Conditions. Electrochim. Acta 2022, 403, 139714,  DOI: 10.1016/j.electacta.2021.139714
    17. 17
      Koper, M. T. M. Structure Sensitivity and Nanoscale Effects in Electrocatalysis. Nanoscale 2011, 3 (5), 20542073,  DOI: 10.1039/c0nr00857e
    18. 18
      Li, R.; Wei, Z.; Huang, T.; Yu, A. Ultrasonic-Assisted Synthesis of Pd-Ni Alloy Catalysts Supported on Multi-Walled Carbon Nanotubes for Formic Acid Electrooxidation. Electrochim. Acta 2011, 56 (19), 68606865,  DOI: 10.1016/j.electacta.2011.05.097
    19. 19
      Du, C.; Chen, M.; Wang, W.; Yin, G. Nanoporous PdNi Alloy Nanowires As Highly Active Catalysts for the Electro-Oxidation of Formic Acid. ACS Appl. Mater. Interfaces 2011, 3 (2), 105109,  DOI: 10.1021/am100803d
    20. 20
      López-Coronel, A.; Torres-Pacheco, L. J.; Bañuelos, J. A.; Álvarez-López, A.; Guerra-Balcázar, M.; Álvarez-Contreras, L.; Arjona, N. Highly Active PdNi Bimetallic Nanocubes Electrocatalysts for the Ethylene Glycol Electro-Oxidation in Alkaline Medium. Appl. Surf. Sci. 2020, 530, 147210,  DOI: 10.1016/j.apsusc.2020.147210
    21. 21
      Guo, M.; Wang, H.; Cui, L.; Zhang, J.; Xiang, Y.; Lu, S. Nickel Promoted Palladium Nanoparticles for Electrocatalysis of Carbohydrazide Oxidation Reaction. Small 2019, 15 (28), 1900929,  DOI: 10.1002/smll.201900929
    22. 22
      Shen, S. Y.; Zhao, T. S.; Xu, J. B.; Li, Y. S. Synthesis of PdNi Catalysts for the Oxidation of Ethanol in Alkaline Direct Ethanol Fuel Cells. J. Power Sources 2010, 195 (4), 10011006,  DOI: 10.1016/j.jpowsour.2009.08.079
    23. 23
      Miao, B.; Wu, Z.-P.; Zhang, M.; Chen, Y.; Wang, L. Role of Ni in Bimetallic PdNi Catalysts for Ethanol Oxidation Reaction. J. Phys. Chem. C 2018, 122 (39), 2244822459,  DOI: 10.1021/acs.jpcc.8b05812
    24. 24
      Yang, H.; Wang, H.; Li, H.; Ji, S.; Davids, M. W.; Wang, R. Effect of Stabilizers on the Synthesis of Palladium-Nickel Nanoparticles Supported on Carbon for Ethanol Oxidation in Alkaline Medium. J. Power Sources 2014, 260, 1218,  DOI: 10.1016/j.jpowsour.2014.02.110
    25. 25
      Dutta, A.; Datta, J. Energy Efficient Role of Ni/NiO in PdNi Nano Catalyst Used in Alkaline DEFC. J. Mater. Chem. A 2014, 2 (9), 32373250,  DOI: 10.1039/c3ta12708g
    26. 26
      Zhao, Y.; Yang, X.; Tian, J.; Wang, F.; Zhan, L. Methanol Electro-Oxidation on Ni@Pd Core-Shell Nanoparticles Supported on Multi-Walled Carbon Nanotubes in Alkaline Media. Int. J. Hydrogen Energy 2010, 35 (8), 32493257,  DOI: 10.1016/j.ijhydene.2010.01.112
    27. 27
      Araujo, R. B.; Martín-Yerga, D.; Santos, E. C.; Cornell, A.; Pettersson, L. G. M. Elucidating the Role of Ni to Enhance the Methanol Oxidation Reaction on Pd Electrocatalysts. Electrochim. Acta 2020, 360, 136954,  DOI: 10.1016/j.electacta.2020.136954
    28. 28
      Liu, Z.; Zhang, X.; Hong, L. Physical and Electrochemical Characterizations of Nanostructured Pd/C and PdNi/C Catalysts for Methanol Oxidation. Electrochem. Commun. 2009, 11 (4), 925928,  DOI: 10.1016/j.elecom.2009.02.030
    29. 29
      Martín-Yerga, D.; Yu, X.; Terekhina, I.; Henriksson, G.; Cornell, A. In Situ Catalyst Reactivation for Enhancing Alcohol Electro-Oxidation and Coupled Hydrogen Generation. Chem. Commun. 2020, 56 (28), 40114014,  DOI: 10.1039/D0CC01321H
    30. 30
      Martín-Yerga, D.; White, J.; Henriksson, G.; Cornell, A. Structure-Reactivity Effects of Biomass-Based Hydroxyacids for Sustainable Electrochemical Hydrogen Production. ChemSusChem 2021, 14 (8), 19021912,  DOI: 10.1002/cssc.202100073
    31. 31
      Ipadeola, A. K.; Lisa Mathebula, N. Z.; Pagliaro, M. V.; Miller, H. A.; Vizza, F.; Davies, V.; Jia, Q.; Marken, F.; Ozoemena, K. I. Unmasking the Latent Passivating Roles of Ni(OH)2 on the Performance of Pd-Ni Electrocatalysts for Alkaline Ethanol Fuel Cells. ACS Appl. Energy Mater. 2020, 3 (9), 87868802,  DOI: 10.1021/acsaem.0c01314
    32. 32
      del Rosario, J. A. D.; Ocon, J. D.; Jeon, H.; Yi, Y.; Lee, J. K.; Lee, J. Enhancing Role of Nickel in the Nickel-Palladium Bilayer for Electrocatalytic Oxidation of Ethanol in Alkaline Media. J. Phys. Chem. C 2014, 118 (39), 2247322478,  DOI: 10.1021/jp411601c
    33. 33
      White, J.; Peters, L.; Martín-Yerga, D.; Terekhina, I.; Anil, A.; Lundberg, H.; Johnsson, M.; Salazar-Alvarez, G.; Henriksson, G.; Cornell, A. Glycerol Electrooxidation at Industrially Relevant Current Densities Using Electrodeposited PdNi/Nifoam Catalysts in Aerated Alkaline Media. J. Electrochem. Soc. 2023, 170 (8), 086504,  DOI: 10.1149/1945-7111/acee27
    34. 34
      Hiltrop, D.; Cychy, S.; Elumeeva, K.; Schuhmann, W.; Muhler, M. Spectroelectrochemical Studies on the Effect of Cations in the Alkaline Glycerol Oxidation Reaction over Carbon Nanotube-Supported Pd Nanoparticles. Beilstein J. Org. Chem. 2018, 14 (1), 14281435,  DOI: 10.3762/bjoc.14.120
    35. 35
      Rahim, S. A. N. M.; Lee, C. S.; Abnisa, F.; Aroua, M. K.; Daud, W. A. W.; Cognet, P.; Pérès, Y. A Review of Recent Developments on Kinetics Parameters for Glycerol Electrochemical Conversion - A by-Product of Biodiesel. Sci. Total Environ. 2020, 705, 135137,  DOI: 10.1016/j.scitotenv.2019.135137
    36. 36
      Rodrigues, T. S.; Zhao, M.; Yang, T.-H.; Gilroy, K. D.; da Silva, A. G. M.; Camargo, P. H. C.; Xia, Y. Synthesis of Colloidal Metal Nanocrystals: A Comprehensive Review on the Reductants. Chem.─Eur. J. 2018, 24 (64), 1694416963,  DOI: 10.1002/chem.201802194
    37. 37
      Poerwoprajitno, A. R.; Gloag, L.; Cheong, S.; Gooding, J. J.; Tilley, R. D. Synthesis of Low- and High-Index Faceted Metal (Pt, Pd, Ru, Ir, Rh) Nanoparticles for Improved Activity and Stability in Electrocatalysis. Nanoscale 2019, 11 (41), 1899519011,  DOI: 10.1039/C9NR05802H
    38. 38
      Gilroy, K. D.; Ruditskiy, A.; Peng, H.-C.; Qin, D.; Xia, Y. Bimetallic Nanocrystals: Syntheses, Properties, and Applications. Chem. Rev. 2016, 116 (18), 1041410472,  DOI: 10.1021/acs.chemrev.6b00211
    39. 39
      Göksu, H.; Ho, S. F.; Metin, O. ̈.; Korkmaz, K.; Mendoza Garcia, A.; Gültekin, M. S.; Sun, S. Tandem Dehydrogenation of Ammonia Borane and Hydrogenation of Nitro/Nitrile Compounds Catalyzed by Graphene-Supported NiPd Alloy Nanoparticles. ACS Catal. 2014, 4 (6), 17771782,  DOI: 10.1021/cs500167k
    40. 40
      Hummelgård, C.; Karlsson, R. K. B.; Bäckström, J.; Rahman, S. M. H.; Cornell, A.; Eriksson, S.; Olin, H. Physical and Electrochemical Properties of Cobalt Doped (Ti,Ru)O2 Electrode Coatings. Mater. Sci. Eng. B 2013, 178 (20), 15151522,  DOI: 10.1016/j.mseb.2013.08.018
    41. 41
      Bolzán, A. Phenomenological Aspects Related to the Electrochemical Behaviour of Smooth Palladium Electrodes in Alkaline Solutions. J. Electroanal. Chem. 1995, 380 (1–2), 127138,  DOI: 10.1016/0022-0728(94)03627-F
    42. 42
      Breiter, M. W. Dissolution and Adsorption of Hydrogen at Smooth Pd Wires of the Alpha Phase in Sulfuric Acid Solution. J. Electroanal. Chem. 1977, 81 (2), 275284,  DOI: 10.1016/S0022-0728(77)80023-5
    43. 43
      Chierchie, T.; Mayer, C.; Lorenz, W. J. Structural Changes of Surface Oxide Layers on Palladium. J. Electroanal. Chem. 1982, 135, 211220,  DOI: 10.1016/0368-1874(82)85121-6
    44. 44
      Hjorth Larsen, A.; Jørgen Mortensen, J.; Blomqvist, J.; Castelli, I. E.; Christensen, R.; Dułak, M.; Friis, J.; Groves, M. N.; Hammer, B.; Hargus, C.; Hermes, E. D.; Jennings, P. C.; Bjerre Jensen, P.; Kermode, J.; Kitchin, J. R.; Leonhard Kolsbjerg, E.; Kubal, J.; Kaasbjerg, K.; Lysgaard, S.; Bergmann Maronsson, J.; Maxson, T.; Olsen, T.; Pastewka, L.; Peterson, A.; Rostgaard, C.; Schiøtz, J.; Schütt, O.; Strange, M.; Thygesen, K. S.; Vegge, T.; Vilhelmsen, L.; Walter, M.; Zeng, Z.; Jacobsen, K. W. The Atomic Simulation Environment-a Python Library for Working with Atoms. J. Phys.: Condens. Matter 2017, 29 (27), 273002,  DOI: 10.1088/1361-648X/aa680e
    45. 45
      Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54 (16), 1116911186,  DOI: 10.1103/PhysRevB.54.11169
    46. 46
      Wellendorff, J.; Lundgaard, K. T.; Møgelhøj, A.; Petzold, V.; Landis, D. D.; Nørskov, J. K.; Bligaard, T.; Jacobsen, K. W. Density Functionals for Surface Science: Exchange-Correlation Model Development with Bayesian Error Estimation. Phys. Rev. B 2012, 85 (23), 235149,  DOI: 10.1103/PhysRevB.85.235149
    47. 47
      Anil, A.; White, J.; Campos dos Santos, E.; Terekhina, I.; Johnsson, M.; Pettersson, L. G. M.; Cornell, A.; Salazar-Alvarez, G. Effect of Pore Mesostructure on the Electrooxidation of Glycerol on Pt Mesoporous Catalysts. J. Mater. Chem. A 2023, 11 (31), 1657016577,  DOI: 10.1039/D3TA01738A
    48. 48
      dos Santos, E. C.; Araujo, R. B.; Valter, M.; Salazar-Alvarez, G.; Johnsson, M.; Bajdich, M.; Abild-Pedersen, F.; Pettersson, L. G. M. Efficient Screening of Bi-Metallic Electrocatalysts for Glycerol Valorization. Electrochim. Acta 2021, 398, 139283,  DOI: 10.1016/j.electacta.2021.139283
    49. 49
      T. C., Allison. NIST-JANAF Thermochemical Tables - SRD 13, 2013.  DOI: 10.18434/T42S31 .
    50. 50
      Mathew, K.; Sundararaman, R.; Letchworth-Weaver, K.; Arias, T. A.; Hennig, R. G. Implicit Solvation Model for Density-Functional Study of Nanocrystal Surfaces and Reaction Pathways. J. Chem. Phys. 2014, 140 (8), 084106,  DOI: 10.1063/1.4865107
    51. 51
      Nørskov, J. K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J. R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108 (46), 1788617892,  DOI: 10.1021/jp047349j
    52. 52
      Sen, B.; Kuzu, S.; Demir, E.; Yıldırır, E.; Sen, F. Highly Efficient Catalytic Dehydrogenation of Dimethyl Ammonia Borane via Monodisperse Palladium-Nickel Alloy Nanoparticles Assembled on PEDOT. Int. J. Hydrogen Energy 2017, 42 (36), 2330723314,  DOI: 10.1016/j.ijhydene.2017.05.115
    53. 53
      Matin, M. A.; Jang, J.-H.; Kwon, Y.-U. PdM Nanoparticles (M = Ni, Co, Fe, Mn) with High Activity and Stability in Formic Acid Oxidation Synthesized by Sonochemical Reactions. J. Power Sources 2014, 262, 356363,  DOI: 10.1016/j.jpowsour.2014.03.109
    54. 54
      Khan, M. S.; Khattak, R.; Khan, A.; Chen, Q.; Nisar, J.; Iqbal, Z.; Rashid, A.; Kamran, A. W.; Zekker, I.; Zahoor, M.; Alzahrani, K. J.; Batiha, G. E.-S. Synthesis and Characterizations of PdNi Carbon Supported Nanomaterials: Studies of Electrocatalytic Activity for Oxygen Reduction in Alkaline Medium. Molecules 2021, 26 (11), 3440,  DOI: 10.3390/molecules26113440
    55. 55
      Wang, L.; Dong, K. J.; Wang, C. C.; Zou, R. P.; Zhou, Z. Y.; Yu, A. B. Computer Simulation of the Packing of Nanoparticles. Powder Technol. 2022, 401, 117317,  DOI: 10.1016/j.powtec.2022.117317
    56. 56
      Eldridge, M. D.; Madden, P. A.; Frenkel, D. Entropy-Driven Formation of a Superlattice in a Hard-Sphere Binary Mixture. Nature 1993, 365 (6441), 3537,  DOI: 10.1038/365035a0
    57. 57
      Wang, Z.; Schliehe, C.; Bian, K.; Dale, D.; Bassett, W. A.; Hanrath, T.; Klinke, C.; Weller, H. Correlating Superlattice Polymorphs to Internanoparticle Distance, Packing Density, and Surface Lattice in Assemblies of PbS Nanoparticles. Nano Lett. 2013, 13 (3), 13031311,  DOI: 10.1021/nl400084k
    58. 58
      Bouju, X.; Duguet, É.; Gauffre, F.; Henry, C. R.; Kahn, M. L.; Mélinon, P.; Ravaine, S. Nonisotropic Self-Assembly of Nanoparticles: From Compact Packing to Functional Aggregates. Adv. Mater. 2018, 30 (27), 1706558,  DOI: 10.1002/adma.201706558
    59. 59
      Wang, L.; Lavacchi, A.; Bellini, M.; D’Acapito, F.; Benedetto, F. D.; Innocenti, M.; Miller, H. A.; Montegrossi, G.; Zafferoni, C.; Vizza, F. Deactivation of Palladium Electrocatalysts for Alcohols Oxidation in Basic Electrolytes. Electrochim. Acta 2015, 177, 100106,  DOI: 10.1016/j.electacta.2015.02.026
    60. 60
      Daily Metal Price: Palladium Price (USD/Gram) Chart for the Last Week. https://www.dailymetalprice.com/metalpricecharts.php?c=pd&u=oz&d=240 (accessed 10 16, 2023).
    61. 61
      Daily Metal Price: Nickel Price (USD/Gram) Chart for the Last Week. https://www.dailymetalprice.com/metalpricecharts.php?c=ni&u=lb&d=0 (accessed 10 16, 2023).
    62. 62
      Oliveira, V. L.; Morais, C.; Servat, K.; Napporn, T. W.; Tremiliosi-Filho, G.; Kokoh, K. B. Glycerol Oxidation on Nickel Based Nanocatalysts in Alkaline Medium - Identification of the Reaction Products. J. Electroanal. Chem. 2013, 703, 56,  DOI: 10.1016/j.jelechem.2013.05.021
    63. 63
      Goetz, M. K.; Bender, M. T.; Choi, K.-S. Predictive Control of Selective Secondary Alcohol Oxidation of Glycerol on NiOOH. Nat. Commun. 2022, 13 (1), 5848,  DOI: 10.1038/s41467-022-33637-7
    64. 64
      Zhang, N.; Wang, J.; Zhang, W.; Zhao, Y.; Dong, Z.; Wu, Z.; Xu, G.-R.; Wang, L. Self-Supported PdNi Dendrite on Ni Foam for Improving Monohydric Alcohol and Polyhydric Alcohols Electrooxidation. Fuel 2022, 326, 125083,  DOI: 10.1016/j.fuel.2022.125083
    65. 65
      Luo, H.; Yukuhiro, V. Y.; Fernández, P. S.; Feng, J.; Thompson, P.; Rao, R. R.; Cai, R.; Favero, S.; Haigh, S. J.; Durrant, J. R.; Stephens, I. E. L.; Titirici, M.-M. Role of Ni in PtNi Bimetallic Electrocatalysts for Hydrogen and Value-Added Chemicals Coproduction via Glycerol Electrooxidation. ACS Catal. 2022, 12 (23), 1449214506,  DOI: 10.1021/acscatal.2c03907
    66. 66
      Mphahlele, N. E.; Ipadeola, A. K.; Haruna, A. B.; Mwonga, P. V.; Modibedi, R. M.; Palaniyandy, N.; Billing, C.; Ozoemena, K. I. Microwave-induced Defective PdFe/C Nano-electrocatalyst for Highly Efficient Alkaline Glycerol Oxidation Reactions. Electrochim. Acta 2022, 409, 139977,  DOI: 10.1016/j.electacta.2022.139977
    67. 67
      Houache, M. S. E.; Shubair, A.; Sandoval, M. G.; Safari, R.; Botton, G. A.; Jasen, P. V.; González, E. A.; Baranova, E. A. Influence of Pd and Au on Electrochemical Valorization of Glycerol over Ni-Rich Surfaces. J. Catal. 2021, 396, 113,  DOI: 10.1016/j.jcat.2021.02.008
    68. 68
      Wang, H.; Thia, L.; Li, N.; Ge, X.; Liu, Z.; Wang, X. Pd Nanoparticles on Carbon Nitride-Graphene for the Selective Electro-Oxidation of Glycerol in Alkaline Solution. ACS Catal. 2015, 5 (6), 31743180,  DOI: 10.1021/acscatal.5b00183
    69. 69
      Melle, G.; de Souza, M. B. C.; Santiago, P. V. B.; Corradini, P. G.; Mascaro, L. H.; Fernández, P. S.; Sitta, E. Glycerol Electro-Oxidation at Pt in Alkaline Media: Influence of Mass Transport and Cations. Electrochim. Acta 2021, 398, 139318,  DOI: 10.1016/j.electacta.2021.139318
    70. 70
      Zhou, Y.; Shen, Y.; Piao, J. Sustainable Conversion of Glycerol into Value-Added Chemicals by Selective Electro-Oxidation on Pt-Based Catalysts. ChemElectroChem 2018, 5 (13), 16361643,  DOI: 10.1002/celc.201800309
    71. 71
      Valter, M.; dos Santos, E. C.; Pettersson, L. G. M.; Hellman, A. Partial Electrooxidation of Glycerol on Close-Packed Transition Metal Surfaces: Insights from First-Principles Calculations. J. Phys. Chem. C 2020, 124 (33), 1790717915,  DOI: 10.1021/acs.jpcc.0c04002
    72. 72
      Valter, M.; Santos, E. C.; Pettersson, L. G. M.; Hellman, A. Selectivity of the First Two Glycerol Dehydrogenation Steps Determined Using Scaling Relationships. ACS Catal. 2021, 11 (6), 34873497,  DOI: 10.1021/acscatal.0c04186
    73. 73
      Yu, X.; dos Santos, E. C.; White, J.; Salazar-Alvarez, G.; Pettersson, L. G. M.; Cornell, A.; Johnsson, M. Electrocatalytic Glycerol Oxidation with Concurrent Hydrogen Evolution Utilizing an Efficient MoOx/Pt Catalyst. Small 2021, 17 (44), 2104288,  DOI: 10.1002/smll.202104288
    74. 74
      Yaylayan, V.; Harty-Majors, S.; Ismail, A. Investigation of Dl-Glyceraldehyde-Dihydroxyacetone Interconversion by FTIR Spectroscopy. Carbohydr. Res. 1999, 318 (1–4), 2025,  DOI: 10.1016/S0008-6215(99)00077-4
    75. 75
      Rasrendra, C. B.; Fachri, B. A.; Makertihartha, I. G. B. N.; Adisasmito, S.; Heeres, H. J. Catalytic Conversion of Dihydroxyacetone to Lactic Acid Using Metal Salts in Water. ChemSusChem 2011, 4 (6), 768777,  DOI: 10.1002/cssc.201000457
    76. 76
      Denis, W. On the Behaviour of Various Aldehydes, Ketones and Alcohols towards Oxidizing Agents. Am. Chem. J. 1907, 38, 561594
    77. 77
      Fedoroňko, M.; Königstein, J. Kinetics of Mutual Isomerization of Trioses and Their Dehydration to Methylglyoxal. Collect. Czech. Chem. Commun. 1969, 34, 38813894,  DOI: 10.1135/cccc19693881
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsaem.3c02789.

    • Experimental methods used for the electrochemical measurements, HPLC standards, description of computational details, and experimental results of galvanostatic measurements (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.