Hui
Luo‡
*ab,
Mianle
Xu‡
c,
Sihang
Liu
c,
Giulia
Tarantino
a,
Hanzhi
Ye
a,
Hossein
Yadegari
d,
Alain Y.
Li
a,
Ceri
Hammond
a,
Georg
Kastlunger
*c,
Ifan E. L.
Stephens
*d and
Maria-Magdalena
Titirici
*a
aDepartment of Chemical Engineering, Imperial College London, South Kensington Campus, London, SW7 2AZ, UK. E-mail: hui_luo@surrey.ac.uk; m.titirici@imperial.ac.uk
bSchool of Mechanical Engineering Sciences, University of Surrey, Stag Hill Campus, Guildford, GU2 7XH, UK
cCatalysis Theory Center, Department of Physics, Technical University of Denmark (DTU), 2800 Kgs, Lyngby, Denmark. E-mail: geokast@dtu.dk
dDepartment of Materials, Imperial College London, South Kensington Campus, London, SW7 2AZ, UK. E-mail: i.stephens@imperial.ac.uk
First published on 4th November 2024
Phasing out petrochemical-based thermoplastics with bio-plastics produced in an energy efficient and environmentally friendly way is of paramount interest. Among them, polylactic acid (PLA) is the flagship with its production accounting for 19% of the entire bioplastics industry. Glycerol electrolysis for producing the monomer lactic acid, while co-generating green H2, represents a promising approach to boost the production of PLA, yet the reaction selectivity has been a bottleneck. Here, we report a combined electrochemical and chemical route using a tandem Pt/C-γ-Al2O3 multicomponent catalyst which can achieve a glycerol-to-lactic acid selectivity of 61.3 ± 1.2%, among the highest performance reported so far. Combining an experimental and computational mechanistic analysis, we suggest that tuning the acidic sites on the catalyst surface is crucial for shifting the reaction towards the dehydration pathway, occurring via dihydroxyacetone intermediate. Within the tandem effect, Pt is the active site to electrochemically catalyze glycerol to dihydroxyacetone and glyceraldehyde, while the γ-Al2O3 provides the required acidic sites for catalyzing dihydroxyacetone to the pyruvaldehyde intermediate, which will then go through Cannizzaro rearrangement, catalyzed by the OH− ions to form lactic acid. This catalytic synergy improves the selectivity towards lactic acid by nearly two-fold. A selectivity descriptor (ΔGGLAD* − ΔGDHA*) from density functional theory calculations was identified, which could be used to screen other materials in further research. Our findings highlight the promise of tandem electrolysis in the development of strategies for selective electrochemical production of high-value commodity chemicals from low value (waste) precursors.
Broader contextUsing renewable energy to produce green hydrogen and chemicals is key for decarbonising the power generation, transport, and chemical industries. Proton exchange membrane water electrolysis is the state-of-the-art technology to generate green hydrogen. However, it has several drawbacks including (1) large thermodynamic potential and sluggish anode kinetics; (2) the use of scarce and expensive iridium as a catalyst; (3) the low economic value of oxygen produced at the anode. Replacing the water oxidation reaction with biomass substrates such as glycerol eliminates the bottlenecks associated with the water oxidation reaction, while enabling green hydrogen production at the cathode at a fraction of the cost. The key to achieve an overall beneficial economic and environmental impact is to selectively produce high-value product on the anode. Herein, we have designed platinum–metal oxide tandem catalyst that can convert glycerol to lactic acid (the precursor for flagship biopolymer poly-lactic acid) at 61.3 ± 1.2% selectivity, while co-produce green hydrogen with an electricity consumption per unit of H2 < 26 kW h kg−1 H2. Our findings highlight the promise of tandem electrolysis in the development of strategies for selective electrochemical production of high-value commodity chemicals from low value (waste) precursors. |
Polylactic acid (PLA) is a bio-based polyester and the most utilized degradable bioplastic4 with its production accounting for 18.9% of the entire bioplastics industry.5 Its monomer, lactic acid, is currently prepared via the enzymatic fermentation of sugars, under strict temperature (<313 K) and pH (5–7) conditions, followed by purification through subsequent esterification, distillation and hydrolysis (Fig. 1a).6,7 Although this approach is used by 90% of the lactic acid manufacturers, its low productivity (1–13 g L−1 h−1) and high production cost dictate the low annual growth rate for PLA. Perez-Ramirez and coworkers demonstrated an alternative approach where glycerol is first oxidised to dihydroxyacetone (DHA) and, in a separate batch, the DHA convert into lactic acid at 413 K and 25 bar.6
![]() | ||
Fig. 1 (a) Illustrations of three different technique routes to produce lactic acid; (b) Benchmark of literature reported values of electrochemical glycerol conversion to lactic acid. The liquid product distribution towards lactic acid is defined as: mole of lactic acid/mole of total liquid products detected ×100%. Other detectable liquid products include: glyceric acid, tartronic acid, glycolic acid, oxalic acid. (Source of references: Co-DPPE;8 Pt3Au7@Ag;9 Pt-CBAC;10 Pt-CC;11 NixBi1−x;12 Planar Au;13 Au NWs;14,15 Au/Ni(OH)2;14,15 AuPt (15% PtSurf);16 Au/CeO2−x;17 Pt/C-Zeolite;18 hp-PtAu/NF.19 Details in reaction conditions are summarized in Table S1, ESI.† To note the use of liquid product distribution towards lactic acid is for literature comparison. In the discussion below, more standard quantification methods with product yield and faradaic efficiency are used throughout this work. |
Glycerol electrolysis to lactic acid could potentially offer a more attractive, continuous route that avoids costs associated with the processing and purification of intermediates (Fig. 1a). Further, being an anodic reaction, this process can be coupled with water reduction, for simultaneous hydrogen production, or other reduction reactions, e.g. CO2 reduction, biomass-derived reduction reactions etc.11,20,21 A life-cycle assessment study demonstrated that electrocatalytic conversion of crude glycerol to lactic acid at 32% liquid product distribution and 100 kg h−1 production rate can result in a 57% reduction in global warming potential compared to the bio- and chemocatalytic processes, when combined with a low-carbon-intensity grid.22 Further emission reduction is possible when the product selectivity towards lactic acid is improved, pointing out the direction for further development. Reports of lactic acid electrosynthesis are limited, as listed in Fig. 1b.8 In most of the cases, the glycerol electrolysis has been performed in a static half-cell configuration rather than a continuous membrane-electrode-assembly (MEA) electrolysis system. However, to maximize the benefits of coupling anode and cathode reactions, the set up must be compatible within a single system with easy product separation and high mass transport, in which case MEA electrolysis cell operating in galvanostatic mode holds greater promise in transferring the research into up-scaling devices than conventional H-cells.23 Besides, to be industrial relevant, the process current density needs to be relatively high (i.e. >200 mA cm−2),24–26 and liquid product distribution towards lactic acid needs to exceed ∼60% to be compatible with the conventional fermentation method.27,28 Reports falling into this area are rare. To date, only Yan et al. reported 80% lactic acid liquid product distribution in their recent works with a two compartment flow cell, yet the method for calculating the glycerol conversion rate and lactic acid selectivity need to be further validated to confirm the high selectivity.14,15 Shi and co-workers also achieved 81% lactic acid liquid product distribution, albiet using a static half cell set up.17 Herein, based on the previous studies, we have chosen to perform our glycerol electrolysis towards lactic acid and H2 co-production in a MEA cell, with a reaction rate of 20 mA cm−2. Although the current density is a magnitude lower than practical industrial scale, the focus here is to achieve high lactic acid yield by designing the electrocatalyst structure and understanding the reaction mechanism. Engineering challenges towards scaling up will later be tackled with strategies such as constructing 3D electrode structure,29 selecting suitable anion-exchange membrane,30,31 investigating in different MEA configurations to minimise electrolyte crossover,32 and engineering the interface microenvironment,33–35 taking inspiration from the electrochemical CO2 reduction community.
Previous studies have shown that the glycerol to lactic acid transformation is a combination of electrochemical deprotonation – heterogeneous dehydration – homogeneous solution phase reaction process, with several intermediate steps involved, as illustrated in red (top) in Fig. 2.16,36,37 Pt-based electrocatalysts are the most active in catalyzing the glycerol deprotonation step towards dihydroxyacetone and glyceraldehyde.38–41 Density functional theory (DFT) calculations have already established the reaction pathways and the factors controlling activity and product selectivity within glycerol (electro-) oxidation on Pt(111) and other metal surfaces.42,43 Subsequently, in a non-electrochemical solution phase step, it is possible to dehydrate DHA to lactic acid.44,45 However, at oxidising potentials on Pt, the dehydration pathway towards lactic acid is less favoured than the competing electrochemical oxidation process towards glyceric acid (shown in blue (bottom) in Fig. 2).16,36,46 Therefore, it is essential to develop an effective strategy to promote the dehydration pathway suppressing the electrochemical oxidation pathway within electrochemical glycerol oxidation.
![]() | ||
Fig. 2 Glycerol dual reaction pathway: dehydration vs. oxidation. Pt based electrocatalysts can effectively catalyze the glycerol → DHA (dihydroxyacetone)/GLAD (glyceraldehyde) step as well as the following oxidation steps.16,36,37 |
In thermal (i.e. non electrochemical) heterogeneous catalysis, acidic sites from certain metal oxides, such as TiO2,47 ZrO2,48 Al2O349 and zeolites50–52 can mediate the chemical transformation of DHA to pyruvaldehyde, by coordinating the carbonyl and hydroxyl groups, thus accelerating the keto–enol tautomerization and dehydration of the component.5 When combined with metal catalysts, some metal oxides such as TiO2 can also directly convert glycerol into lactic acid, with the metal providing the dehydrogenation sites to produce DHA, and the transformation to lactic acid taking place on TiO2 surface.53
Taking inspiration from the abovementioned strategies form heterogeneous catalysis, in this paper, we aim to combine electrochemical experiments and DFT simulations to discover tandem electrocatalysts consisting of Pt and metal oxides for improving lactic acid product yield. We first examined the lactic acid product yield % on Pt/C catalyst standalone and employ DFT calculations to identify the reaction limitations. Subsequently, we screened a series of metal oxide materials with different acidic site densities, and designed a multicomponent tandem catalyst system containing Pt sites and metal oxides, to identify the experimental and theoretical descriptor for improving the product yield towards lactic acid leveraging further product optimization attempts.
![]() | ||
Fig. 3 Electrolysis data showing the cell voltage change during 1 h experiment under 20 mA cm−2 applied current density. Inserts are the quantified products, defined as ![]() ![]() ![]() ![]() ![]() |
To rationalize the experimental observations on Pt above, we applied DFT to study the competition between the mixed electrochemical & chemical reaction path toward lactic acid and purely electrochemical (EC) pathways to the other products. We have simulated the Pt nanoparticle surfaces on the most stable (111) facet.54 As shown in the free energy diagram in Fig. 4, we find that on a Pt(111) surface at 0.5 V vs. RHE, the most endergonic step of lactic acid production (shown in red in Fig. 4) is the chemical transformation of surface bound 2-hydroxyacrylaldehyde to pyruvaldehyde. On the contrary, the competing electrochemical pathway is virtually exergonic throughout (shown in blue in Fig. 4). The potential response of the latter pathway leads to a takeover of the product via the electrochemical path (shown in blue in Fig. 2 and 4) at increasing overpotentials. However, close to 0 VRHE the limiting potential for both pathways is determined by glycerol electro-oxidation to dihydroxyacetone (*DHA) or glyceraldehyde (*GLAD), respectively, with the formation of *DHA being preferred over *GLAD. The free energy diagrams at different potentials are shown in Fig. S4 (ESI†). Besides, considering the pure electrochemical pathways, some elementary steps (outlined in in Fig. S5, ESI†) exhibit high limiting potentials (up to 0.85 VRHE), as shown in the free energy diagram in Fig. S6a (ESI†), which inhibit the further production of tartronic acid and C–C splitting products at r at the reaction conditions applied in this study. However, the electrochemical reaction proceeds smoothly at 0.5 V vs. RHE, as shown in Fig. S6b (ESI†). Considering the reaction path toward lactic acid (red line in Fig. 4) at 0.5 V vs. RHE, DHA as proton–electron transfer product is quite feasible and the following steps toward lactic acid are non-electrochemical which is potential independent. Therefore, the low selectivity towards lactic acid on pure Pt can be attributed to a combination of limiting electrochemical activity towards DHA and GLAD and the predominance of the electrochemical process towards glyceric acid, tartronic acid and C–C splitting products at elevated overpotentials. To increase lactic acid selectivity, engineering a catalytic system that promotes the dehydration route at moderate electrode potentials is thus required.
![]() | ||
Fig. 4 DFT-calculated free energy diagrams of the glycerol oxidation pathways outlined in Fig. 2. on Pt(111) at 0.5 V vs. RHE. The blue line is the purely electrochemical (EC) path and the red line is the Mixed electrochemical and chemical (EC&C) path. The related structures of different pathways are shown at the top and bottom of figure that the main product of purely EC is outlined in blue and the lactic acid is outlined in red. (l): liquid phase, *: adsorption at surface. More details on DFT results are in ESI:† computational details. Color codes for atoms: grey – platinum, white – hydrogen, red – oxygen, dark grey – carbon. |
A series of metal oxides with different surface density of acidic sites were tested.55 We fabricated the Pt/C-MOx (MOx (metal oxide) = CeO2, TiO2 and γ-Al2O3) multicomponent electrodes by physically mixing commercial Pt/C (HiSPEC® 9100, Johnson & Matthey) and metal oxide nanopowders (Sigma Aldrich) and spray-coating on carbon paper support, as described previously for Pt/C. The mass ratio between Pt/C and metal oxides as well as the Pt loadings on all electrodes were kept equal as Pt/C, maintained at 1:
3 ratio and Pt loading of 0.1 mg cm−2, respectively. The resulting morphology was characterized using electron microscopy, energy-dispersive X-ray spectroscopy (EDS) mapping and XRD. As shown in the TEM and EDS mappings (Fig. 5 and Fig. S2, S3 (ESI†)), in all Pt/C-MOx multicomponent systems, the metal oxide nanoparticles are intimately mixed at the nanoscale, providing good contact for charge and mass transport. From the morphologies and XRD patterns (Fig. S3, ESI†), it can be seen that CeO2 and TiO2 have a more crystalline structure, while Al2O3 is more amorphous.56
![]() | ||
Fig. 5 (a) The morphology of Pt/C-Al2O3 multicomponent catalyst: (a) TEM image; (b)–(e) STEM-EDS (b: HAADF-STEM, c: multi-element color mix; d: Al map; e: Pt map). |
Electrolysis measurements applying Pt/C-MOx tandem catalysts were carried out under the same conditions as for the Pt/C. All the metal oxides and the carbon black features similar specific surface area (Fig. S9 and Table S2, ESI†), eliminating the effect from significant surface area differences.57Fig. 6a shows the product yield % for each product produced by Pt/C-MOx and Pt/C control. The results show that all Pt/C-MOx exhibit higher product yield towards lactic acid compared to the Pt/C control tested previously. In particular, for Pt/C-Al2O3, the lactic acid product yield % reached 61.3 ± 1.2%, nearly double that of the Pt/C standalone. Although the faradaic efficiency towards glyceric acid in Fig. S8 (ESI†) is still higher than lactic acid, as producing the latter needs only half the number of the electrons, it is obvious that after adding the metal oxides, the partial current density towards lactic acid has significantly improved. The category “Others” take into account all products that are not quantified (carbonate (which is the main component), formate, oxalate, etc.) through HPLC. Since carbonate formation accounts for 12 electrons but only one glycerol gets converted, the proportion represented in FE plot is substantially larger than in product yield. The extended stability of the glycerol electrolysis process was evaluated using the Pt/C-Al2O3 electrode. As shown in Fig. S10 (ESI†), the glycerol electrolysis system exhibits a stable cell voltage below 1 V during 72 h of continuous operation at 20 mA cm−2, corresponding to an electricity consumption per unit of H2 < 26 kW h kg−1 H2 (vs. ∼38 kW h kg−1 H2 for alkaline water electrolyser at similar current density),31 while the lactic acid product yield is maintained above 45%.
![]() | ||
Fig. 6 (a) The product yield % for Pt/C-MOx and Pt/C control. Values are averaged from at least three independent measurements. Current density: 20 mA cm−2. Electrolysis duration: 1 h. The corresponding electrolysis data and HPLC chromatograph compound calibration and assignment are shown in Fig. S7 (ESI†). The faradaic efficiency values of each product shown in Fig. S8 (ESI†); (b) experimental NH3-TPD profile on different surfaces: carbon black, TiO2, CeO2 and Al2O3; (c) the correlation between lactic acid product yield % and the acidic surface site density determined by NH3-TPD. (d) The difference in adsorption free energy between GLAD and DHA as a descriptor of experimental results. The DHA and GLAD adsorption structure on (e). (i), (iii) Pt(111) and on (f). (ii), (iv) γ-Al2O3(111), respectively. All relevant structures and energetics in our DFT calculations are provided in ESI,† Section 5, Fig. S12–S14. Color codes: red – oxygen, brown – carbon, silver– platinum, pink – hydrogen, blue – aluminum. |
To further validate whether the acidic sites on the metal oxide surfaces play a role, NH3 temperature programmed desorption (NH3-TPD) was used to determine the total acidic site concentration and relative strength of the metal oxide catalysts. The data profile is presented in Fig. 6b, and the acidity of all metal oxides was calculated by integrating the area under the profile curve normalised by the mass of the catalyst. As a benchmark, the surface acidity of pure carbon black was also measured. The values are summarized in Table S2 (ESI†). As Fig. 6c shows, we identified a linear relationship between the lactic acid product yield values and acidic site density of different metal oxides.
DFT calculations were performed to further elucidate why the addition of metal oxides enhances the product yield towards lactic acid by studying the adsorption of reaction intermediates on the metal oxides with respect to Pt.
We have modelled the most stable facets for fluorite-CeO2 and anatase-TiO2 following the XRD results shown in Fig. S3 (ESI†), which suggest predominance of the (111) and (101)-facets respectively. Given γ-Al2O3 amorphous’ nature, which did not even exhibit specific peak in XRD in the combined Pt/γ-Al2O3 system, we adopted the (111)-facet as a pragmatic compromise (see ESI,† Section 5c for more detail). The applied surface structures and the (meta-)stable molecules adsorption in our calculations are shown in Fig. S12 and S13 (ESI†).
As shown in Fig. S14 (ESI†), generally, TiO2 and CeO2 bind the relevant reaction intermediates slightly weaker than Pt. γ-Al2O3, on the other hand, shows strengthened binding throughout. This hints towards γ-Al2O3's enhanced activity, as it can retain DHA and GLAD, thus removing them from the solution phase and preventing further electrochemical oxidation on Pt. The enhanced lactic acid selectivity for the Pt/TiO2 and Pt/CeO2 systems, however, is not evident in this analysis.
We identified that the adsorption free energies of both DHA and GLAD correlate with NH3 adsorption, our probe to titrate the number of acidic sites (cf. Fig. S15, ESI†). Thus, merely increasing the acidic site activity does not fully explain the improved lactic acid yield. The relative adsorption strength of the two key intermediates, on the other hand, shows a clear trend reflecting the identified selectivity behaviour, with experimental lactic acid product yield % correlates linearly with (ΔGGLAD* − ΔGDHA*), the ability of the materials in adsorbing DHA, the precursor of lactic acid, over GLAD (Fig. 6d). We interpret this (ΔGGLAD* − ΔGDHA*). relative behaviour as the ability to preferentially remove DHA from the GLAD–DHA solution equilibrium, leading to further production of DHA in solution at the expense of GLAD. DHA can undergo further reaction on both Pt and the metal oxide surfaces. Thus, DHA is constantly removed from the solution, pushing the solution equilibrium towards producing it even at steady state.
Pure Pt preferably binds GLAD over DHA, while TiO2 and CeO2 adsorb both intermediates with comparable strength and γ-Al2O3 binds DHA stronger. To understand this selectivity descriptor in more detail, we highlight the binding configurations of DHA and GLAD on Pt and Al2O3 in Fig. 6e and f. GLAD tends to chemisorb via its central OH-group on both surfaces. DHA on the other hand is only physisorbed in a flat configuration on Pt, while being chemisorbed via both its terminal OH-groups on the acidic sites of Al2O3. An analogous chemisorbed binding configuration of DHA has been identified on the other studied metal oxide catalysts, as shown in Fig. S13 (ESI†).
Thus, we suggest that while both GLAD and DHA benefit from the acidity of the catalyst binding sites, the chemisorption of DHA on metal oxide catalysts compared to the mere physisorption on Pt is the primary cause in the increase in selectivity and product yield towards lactic acid and motivate the validation of the descriptor (ΔGGLAD* − ΔGDHA*) in future efforts towards lactic acid selectivity improvement.
As a final word of caution, we note that enhanced surface acidity does also lead to a competition of DHA adsorption with *OH formation on the active sites at reaction conditions. We show an evaluation of this deactivation mechanism in Fig. S16–S18 (ESI†). The active sites on both TiO2 and CeO2 have been identified to be *OH and *O free, while γ-Al2O3 exhibited a preferential coverage of 0.4 ML of *OH at reaction conditions. However, the binding of DHA exceeds the formation energy of *OH at the mild oxidative potential applied in this study, indicating its ability to displace *OH. Once, the potential would be raised further (to ∼0.75 V vs. RHE, cf. Fig. S18, ESI†), γ-Al2O3 might be deactivated as *OH's formation is stabilized by more oxidating potentials, in contrast to the adsorption of DHA. Thus, a balance in the surface acidity needs to be targeted which is high enough to exhibit preferential adsorption of DHA over GLAD, while not poisoning the surface with *OH at reaction conditions.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ey00236a |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2025 |