Electrooxidation of Glycerol on Gold in Acidic Medium - American

Phone: +46 31 772 56 11. Abstract. Glycerol is a byproduct of biodiesel production and an abundant feedstock for syn- thesis of high-value chemicals. ...
0 downloads 5 Views 2MB Size
Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Electrooxidation of Glycerol on Gold in Acidic Medium: A Combined Experimental and DFT Study Mikael Valter, Michael Busch, Bjorn Wickman, Henrik Grönbeck, Jonas Baltrusaitis, and Anders Hellman J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b02685 • Publication Date (Web): 19 Apr 2018 Downloaded from http://pubs.acs.org on April 19, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Electrooxidation of Glycerol on Gold in Acidic Medium: a Combined Experimental and DFT Study Mikael Valter,† Michael Busch,‡ Björn Wickman,‡ Henrik Grönbeck,† Jonas Baltrusaitis,¶ and Anders Hellman∗,† †Department of Physics and Competence Centre for Catalysis, Chalmers University of Technology, SE-412 96 Göteborg, Sweden ‡Department of Physics, Chalmers University of Technology, SE-412 96 Göteborg, Sweden ¶Department of Chemical and Biomolecular Engineering, Lehigh University, B336 Iacocca Hall, 111 Research Drive, Bethlehem, PA 18015, USA E-mail: [email protected] Phone: +46 31 772 56 11

Abstract Glycerol is a byproduct of biodiesel production and an abundant feedstock for synthesis of high-value chemicals. A promising approach for valorization of glycerol is electrooxidation on gold. In this work, we investigate electrooxidation of glycerol on gold in acidic media using cyclic voltammetry and density functional theory calculations. Experimentally, we observe activity for electrooxidation above a potential of 0.5 V vs. the reversible hydrogen electrode (RHE). A Pourbaix diagram is calculated to evaluate the surface coverage under reaction conditions, indicating that the surface is free from adsorbates at the measured onset potential. Computationally, we find that the onset

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

potentials for partial dehydrogenation of glycerol to dihydroxyacetone, 2,3-dihydroxy2-propenal and glyceraldehyde are 0.39, 0.39, and 0.60 V vs. RHE, respectively, while complete dehydrogenation to carbon monoxide requires 0.50 V vs. RHE. Our theoretical and experimental findings are in agreement and show the possibility of using gold as a catalyst for production of hydrogen and other valuable chemicals from glycerol.

Introduction Glycerol is a byproduct originating from transesterification of vegetable oils and animal fat into biodiesel. 1 The global production of biodiesel during 2015 was 26 million t and is projected to increase to 35 million t by 2025, 2,3 corresponding to 3 and 4 million t of glycerol, 4 respectively. Glycerol can be used as a feedstock in biological conversion systems for production of valuable chemicals, such as acrolein, butanol, and syngas. 5 Another possibility to utilize glycerol is in conventional catalytic processes, 6 for instance, dehydrogenation 7 or electrooxidation 8–10 to produce H2 and derivatives such as dihydroxyacetone (DHA), glyceraldehyde or glyceric acid. H2 has many applications, such as direct fuel applications, 11,12 or upgrading biodiesel via hydrogenation of unsaturated fatty methyl esters, 13 while other derivatives are of use in cosmetic industry, 10 and production of bioplastics. 14 Glycerol electrooxidation has been studied experimentally on transition metals, including platinum and gold. 15–19 One one hand, gold has been shown to be an excellent electrocatalyst for glycerol oxidation in alkaline solution. 15,17 It displays an order of magnitude higher activity as compared to platinum, which has been attributed to the more anodic potential of gold oxidation. 17,20 On the other hand, gold is reported to have no activity in acidic solution, 15–17 owing to the lack of proton acceptors such as hydroxyl ions in solution (OH− ), as well as surface-bonded hydroxo adsorbates (*OH). 17,20 The presence of Brønsted bases shifts the equilibrium of

− + R−OH − )− −* − R−O + H

2

ACS Paragon Plus Environment

Page 2 of 21

Page 3 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

to the right, which promotes glycerol oxidation, as R−O – is more reactive than R−OH. 20 In this work, we show experimentally that glycerol electrooxidation on gold has a low, but detectable, activity in acidic solution. In particular, glycerol oxidation is detected on the negative going scan, which has previously been observed for glycerol oxidation on platinum and platinum-ruthenium alloys in acidic media. 17,21,22 The activity for glycerol oxidation on gold in acidic media raises the question of the underlying reaction mechanism without initial glycerol deprotonation in solution. In our computational approach, we model the surface as Au(111), since this facet has the lowest surface energy, 23 and thus should be abundant. Our computed Pourbaix diagram indicates that the reaction occurs without presence of hydroxo adsorbates. Using the theoretical normal hydrogen electrode 24 (NHE), we study the thermodynamics of glycerol dehydrogenation on bare Au(111) to find possible reaction paths and corresponding theoretical onset potentials.

Method Cyclic voltammetry was performed on a polycrystalline gold wire in a solution of 0.1 M HClO4 (Sigma-Aldrich 70 %, ACS reagent grade, diluted with milli-Q water) and 0.5 M H2 SO4 (Sigma-Aldrich 95–98 %, ACS reagent grade, diluted with milli-Q water) and 0.5 M glycerol (Sigma-Aldrich, 99.5 % purity). An Ag/AgCl reference electrode (B3420+) from SI Analytics was used, and a graphite rod (Sigma-Aldrich 99.995 % purity) was used as counter electrode. The scan rate was 100 mV/s. Nitrogen was bubbled to purge atmospheric oxygen 20 minutes before and during the experiments. The experiments were conducted with an SP-300 potentiostat from BioLogic Instruments. Underpotential deposition of copper (UPD) was carried out in in 1 mM CuSO4 (SigmaAldrich, 99.995 % purity) and 0.5 M H2 SO4 . The potential was scanned down to 0.2 V vs. the reversible hydrogen electrode (RHE) in order to deposit one monolayer of Cu and the obtained charge was used to calculate the electrochemical area, as described by Rouya et

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

al. 25 Density functional calculations were performed using VASP, 26–28 with the optB86b-vdW exchange-correlation functional. 29–31 This functional was chosen as it gives accurate adsorption energies on gold. 32 The projector augmented wave method 33 was used to model the interaction between the valence electrons and the core. The Kohn-Sham orbitals were obtained using a plane wave basis set with 450 eV as cutoff energy and a Gaussian smearing of 0.05 eV was applied to the Fermi level discontinuity. Calculations on gas phase radicals were performed using spin polarization. The Au(111) surface was modelled as a 4-layer p(3 × 3) supercell, for the dehydrogenation reaction, and a 4-layer p(2 × 2) supercell, for the Pourbaix diagram. Solvent effects were ignored owing to the complexity and lack of fundamental understanding of the double layer. 34 This is a common approximation that has previously been used successfully to understand electrochemical reactions. 35,36 The periodic surface slabs were separated by a vacuum of 20 Å. The p(3 × 3) and p(2 × 2) supercells were sampled in a Monkhorst–Pack grid 37 with (6,6,1) and (8,8,1) k-points, respectively. The gas phase species were computed in a (20 × 20 × 20) Å cell using only the gamma point. The Quasi-Newton method was used for structural relaxations with total residual force of 0.02 eV/Å as convergence criterion. Vibrational modes were calculated by diagonalization of the partial Hessian matrix. Only adsorbates were considered and the forces were computed by means of the central difference approximation with a displacement of 0.05 Å. Gibbs free energy was calculated at standard temperature and pressure. The adsorbates were treated in the harmonic approximation with only vibrational degrees of freedom. As the pV -term was considered negligible, Gibbs free energy of the adsorbate was approximated as Helmholtz free energy, which was calculated at 298 K with vibrational entropy and zeropoint correction. Gibbs free energy was calculated for gas phase species by treating them as ideal gases. Electrochemical reactions were modelled using the theoretical NHE by Rossmeisl and

4

ACS Paragon Plus Environment

Page 4 of 21

Page 5 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Nørskov, which assumes coupled electron-proton transfer. 24 Gibbs free energy is calculated as

∆G = GC3 Hx O∗3 − G∗ − Gglycerol(g) +

8−x GH2 (g) − (8 − x)kB T ln 10 × pH + (8 − x)eUNHE (1) 2

where a star indicates an adsorption site, kB is the Boltzmann constant, x the number of hydrogen atoms in the species, e the elementary charge and UNHE the potential vs. NHE. Using RHE, the pH dependence is included in the potential term for coupled electron-proton transfer, leaving us with

∆G = GC3 Hx O∗3 − G∗ − Gglycerol(g) +

8−x GH2 (g) + (8 − x)eURHE 2

(2)

Adsorption energy for a species is defined as the energy of the adsorbed species relative to the species in gas phase.

Results and discussion Cyclic voltammetry was carried out to investigate the activity of glycerol oxidation in acidic media, shown in Figure 1. In order to avoid oxygen evolution, which is reported to start at 1.9 V, 38 the potential was scanned between 0.1 V and 1.65 V vs. RHE. The surface oxidation occurs from 1.3–1.4 V on the anodic (positive going) scan, and the corresponding surface reduction can be seen clearly at 1.2–1.3 V on the cathodic (negative going) scan. All currents and charges below are defined as the difference of the data with and without glycerol. Glycerol oxidation in HClO4 (Figure 1a) takes place in three regions: On the anodic scan (i) before the surface reaction, starting at ∼0.5 V, (ii) after surface oxidation, and (iii) on the cathodic scan around 0.6 V. The maximum current in the metallic surface region is 0.08 mA/cm2 and in the oxidized surface region 0.34 mA/cm2 . The total oxidative charge over a cycle is 1.10 mC/cm2 . Based on these observations, we conclude that gold has an

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

a)

ii

iii

i

b)

ii

i iii

Figure 1: Cyclic voltammogram of a polycrystalline gold electrode in (a) 0.1 M HClO4 and (b) 0.1 M H2 SO4 with and without 0.5 M glycerol (blue solid and red dash-dot, respectively). Scan rate 100 mV/s. Regions of interest for glycerol oxidation are marked with i, ii, and iii, respectively.

6

ACS Paragon Plus Environment

Page 6 of 21

Page 7 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

measurable activity for glycerol oxidation in HClO4 . In contrast to HClO4 , no activity for glycerol oxidation is observed in 0.5 M H2 SO4 (Figure 1b) before surface oxidation on the anodic scan (i). After surface oxidation (ii), as well as on the negative going scan (iii), some activity is observed. The activity is lower in H2 SO4 compared to HClO4 ; the maximum current in the metallic surface region is 0.03 mA/cm2 and in the oxidized surface region 0.27 mA/cm2 . The total oxidative charge over a cycle was 0.16 mC/cm2 , implying that glycerol oxidation also has activity in H2 SO4 . A possible explanation for the different electrolyte activities is the competition between the electrolyte and glycerol. Both perchlorate and sulfate ions are expected to compete with the relatively weakly bound electrically neutral glycerol for free adsorption sites. In the case of perchorate, still sufficiently many free sites are available to facilitate the experimentally observed glycerol oxidation. In contrast, sulfate is known to bind stronger than perchlorate to the anode, 39,40 resulting in blocking of the surface in agreement with the lack of observed activity. Kwon et al. 17 and Beden et al. 15 investigated glycerol oxidation on gold in H2 SO4 and reported no activity. A possible explanation for this is that a lower glycerol concentration (0.1 M) was used in these studies. 15,17 Additionally, in Ref. [ 17 ], the glycerol oxidation on the cathodic scan would not have been detected since anodic linear sweep voltammetry was used. Our voltammograms are qualitatively similar to reported results in HClO4 and H2 SO4 on platinum and Pt/Ru, 16,21,22 with respect to the glycerol oxidation on the cathodic scan at 0.6 V vs. RHE (Figure 1, iii). Prior to this peak, we observe a small reduction pre-peak at 0.8 V, similar to reports from Kahyaoglu et al., 16 which might be of importance for the main peak. The underlying reason for the oxidation on the cathodic scan is not known. Possible explanations include removal of electrolyte anions from the surface, temporary formation of active sites owing to reduction of the surface, or access to active sites with higher reactivity, e.g. kinks, steps, and edges. However, these speculations are outside the scope of this study. Knowing that there is activity for glycerol oxidation on gold, we decided to study the

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

a ½ ML *OH

b ¼ ML *O

c ½ ML *O

d ¼ ML *O ½ ML *OH

Au(111) a or b

c or d

Figure 2: Surface Pourbaix diagram of Au(111). The surface is bare until 1.0 V vs. RHE, where (a) 1/2 ML *OH and (b) 1/4 ML *O are formed. At 1.5 V, (c) 1/2 ML *O and (d) a mixture of 1/4 ML *O and 1/2 ML *OH are formed.

8

ACS Paragon Plus Environment

Page 8 of 21

Page 9 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

mechanism computationally. Choosing Au(111) as model surface, we constructed a surface Pourbaix diagram 41 (Figure 2) to evaluate the surface coverage of adsorbed hydroxo (*OH) and oxo (*O) at reaction conditions. The configuration space was mapped by calculating 1/4, 1/2, 3/4, and 1 monolayer (ML) of *OH and *O at fcc, bcc, bridge and top sites. The surface starts being covered by *OH and *O at 1.0 V vs. RHE. In this region, (a) 1/2 ML *OH (bridge) and (b) 1/4 ML *O (fcc) differ energetically by only 0.002 eV/Å2 and are thus expected to coexist. At 1.5 V, (c) 1/2 ML *O (fcc) and (d) a mixture of 1/4 ML *O (fcc) and 1/2 ML *OH (top, bridge) are formed with an energy difference of 0.003 eV/Å2 . The other configurations are less stable and are therefore not included in the diagram. The initial gold oxidation at 1.0 V is in agreement with previous computational results. 38 At higher potentials, thin bulk hydroxides and oxides are reported to form. 38 According to our experiments as well as in literature, gold oxidation starts at 1.2–1.3 V on the anodic scan, while reduction ends at 1.0–1.1 V on the cathodic scan. 38,42,43 This is in good agreement with our theoretical results. It has been reported that adsorbed CO enhances co-adsorption of OH groups at as low potentials as 0.6 V vs. RHE 44 and that it enhances electrooxidation of methanol. 45 Formation of CO from glycerol oxidation could thus potentially have an autocatalytic effect, enhancing reaction kinetics. However, such an effect lies outside the scope of the present study. We conclude that the Au(111) surface is water covered below 1.0 V vs. RHE. Neglecting solvent effects, glycerol oxidation is, therefore, modeled assuming a bare Au(111) surface. There is a large number of possible ways of oxidizing glycerol. As the Pourbaix diagram shows that there is no oxygen available for glycerol oxidation, we restrict ourselves to study thermodynamics of electrochemical dehydrogenation, with the addition of allowing concerted C-C bond breaking. As the number of possible intermediates still remain large, glycerol dehydrogenation paths were explored through the following scheme: From glycerol, all possible ways of single deprotonation were tested. Then, we continue with the two most thermodynamically favorable steps by deprotonating them in turn, and so on. Based on this,

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry

a)

H+ + e-

7b OH

O O

O

PDS 0.50 V

Au

O

H+ + e-

Au

Au

Au

O CO + Au

O

H+ + e-

OH

CO Au

3 CO

Most favorable route

OH HO

HO

HO

HO

3a 2a HO

H+ + e-

Au

O

OH

Dissolution O

O

HO

4b

Au Au

Au

OH HO Au Au

H+ + eHO

OH

H+ + e-

OH

OH HO

HO

O

OH Au Au

Au H+ + e-

1ab O

H+ + e-

PDS 0.62 V H+ + e-

Au

2c OH

H+ + e-

Au Au

O

OH

1c

Au

OH

HO

HO PDS 0.60 V

5b

OH

OH

0 OH

4a

O

H+ + e-

Glyceraldehyde route H+ + e-

H+ + e-

Second most favorable route

O Au Au

OH

O Au Au

6b

HO

8

5a O

OH

O

Au Au

3x

HO

H+ + e-

C•

7a

6a

HO

HO

2b

Dissolution

O

Au Au OH

OH

3b

H+ + e-

Au

O HO

OH

b) Gibbs Free Energy (eV)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 21

1.5

PDS 0.50

7b 6a

0.03 0.27

1.0

6b

Most favorable route Second most favorable route Glyceraldehyde route

PDS 0.60

0.5 OH HO

-0.18

2b

0.39 -0.21

5a

3 CO

8 0.03

0.40

0.40

3b

-0.66

-0.31

0.0

0.08

-0.31

5b 1c 1ab

OH

4a

-0.40

7a

0.38

3a 0.42

-0.26

4b

PDS 0.62

2a

0 2c

-0.5

Glycerol

0

- (H+, e-) - 2(H+, e-) - 3(H+,e-) - 4(H+,e-) - 5(H+,e-) - 6(H+,e-) - 7(H+,e-) - 8(H+,e-)

CO

Figure 3: Catalytic routes of glycerol dehydrogenation on Au(111), presented as (a) catalytic cycles and (b) energy landscape. The two most favorable complete dehydrogenation routes to CO are marked in red with squares and blue with circles, respectively, while a path ending in adsorbed glyceraldehyde is shown in green with triangles. Potential determining steps (PDS) are marked for each route.

10

ACS Paragon Plus Environment

Page 11 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

we find the two most favorable complete dehydrogenation paths to CO, shown in Figure 3. These paths require 0.50 V and 0.62 V vs. RHE, respectively, in order to be thermodynamically favorable. A path to glyceraldehyde is included inspired by experimental observation on platinum in acidic media and on gold at higher pH by Kwon et al. 17 Starting from adsorbed glycerol (0), the most favorable route begins with dehydrogenation of the secondary carbon (1ab, requiring 0.39 V vs. RHE), and the secondary hydroxyl group, forming DHA (2a, -0.31 V). DHA is weakly bound to the surface, with an adsorption energy of -0.15 eV, and may desorb. The next steps are removal of hydrogen from the primary carbons, forming a six-membered ring (3a, 0.42 V) and a five-membered ring (4a, 0.40 V) with surface gold, respectively. Continuing from 4a, a di-aldehyde is formed by deprotonization of the primary hydroxyl groups (5a, 0.08 V; 6a, 0.50 V). The transition from 5a to 6a is the potential determining step (PDS). The last two steps are the removal of hydrogens on the primary carbons. Assuming concerted C-C bond breaking, this leads to formation of CO (7a, 0.03 V; 8, -0.31 V), which readily desorbs. C-C bond breaking is exergonic by -0.31 eV and -0.99 eV for 7a and 8, respectively. Thus, the last dehydrogenation steps with decoupled C-C bond breaking requires potentials of 0.41 V and 0.31 V, respectively, which does not change the PDS. The second most favorable route starts also with steps 0 and 1ab, followed by dehydrogenation of a primary carbon. Surface gold binds into the C-C double bond, forming a π-complex (2b, -0.18 V). Next, a five-membered ring is formed by dehydrogenation of a primary hydroxyl group (3b, 0.38 V). The deprotonation of a primary carbon leads to formation of 2,3-dihydroxy-2-propenal (4b, -0.26 V). This species is stable, forming a weak van der Waals bond to the surface, with an adsorption energy of -0.19 eV, and may thus desorb. Removal of hydrogen from the aldehyde group (5b, 0.62 V) is the PDS. Deprotonation of the second primary carbon results in the formation of a five-membered ring with surface gold (6b, 0.40 V). Deprotonation of the secondary carbon (7b, 0.27 V) is followed by concerted C-C bond breaking and CO formation (8, -0.38 V). With decoupled C-C bond breaking, step

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

8 would have required 0.58 V, not changing the PDS. The glyceraldehyde route starts from glycerol (0), followed by dehydrogenation of a primary carbon (1c, 0.60 V), which is the PDS. Glyceraldehyde (2c, -0.66 V) is formed by dehydrogenation the corresponding primary hydroxyl group. Further dehydrogenation would require 0.70 V vs. RHE. From these results we conclude that activity, owing to the first deprotonation, should start at 0.39 V vs. RHE. This is in good agreement with our experimental onset potential of ∼0.5 V vs. RHE, considering that kinetic effects are not treated in our theoretical study. There are, to the best of our knowledge, no reports on oxidation products on gold in acidic media. In neutral media, glyceraldehyde is reported to be produced from 0.8 V vs. RHE, 17 which, if the computational model surface is valid, agrees with our results. Our reaction mechanism can be compared to Liu and Greeley’s computational work on glycerol dehydrogenation on late transition metals, in particular on Pt(111). 46,47 The study is conducted in the framework of heterogeneous catalysis, but as H2 is used as reference for the removed hydrogen, we can interpret their results using the theoretical NHE. 24 Using the approach above, one finds that the two first electrochemical dehydrogenation steps on platinum are identical to the second most favorable route (1ab and 2b in Figure 3a). Then, the second primary carbon is deprotonated, deviating from our routes on gold. The primary hydroxyl groups are deprotonated (0.38 V), which is most likely the PDS if concerted C-C bond breaking is allowed. Liu and Greeley consider complete dehydrogenation with decoupled C-C bond breaking. 46,47 Thus, their route ends with the formation of a surface-bound (C=O)-(C=O)-(C=O), which is then the PDS (0.73 V). Considering our calculated thermodynamic C-C bond breaking gain on gold above, and considering a reported experimental onset potential on platinum of ∼0.4 V vs. RHE, 16,17,22 we conclude that concerted C-C bond breaking occurs in electrochemical dehydrogenation on platinum, similar to our result on gold.

12

ACS Paragon Plus Environment

Page 12 of 21

Page 13 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Conclusions We have shown that glycerol oxidation on gold occurs to a measurable extent in acidic media, in particular on the bare metallic surface < 1.0 V, in contrast to previous reports. 15,17 The choice of electrolyte is important, as the activity in HClO4 is significantly higher than in H2 SO4 . We speculate that the difference in activity is due to competitive adsorption, as it is known that sulfate ions adsorb stronger than perchlorate ions. 39,40 In order to investigate the glycerol oxidation reaction mechanism, we have validated the surface coverage at reaction conditions by constructing a theoretical surface Pourbaix diagram for Au(111). It was found that hydroxy groups and oxo groups appear on the surface above 1.0 V vs. RHE. We have explored the most thermodynamically favorable glycerol dehydrogenation paths on Au(111). From the energy landscape (Figure 3b), we determine which products should be observed at different potentials. At 0.39 V vs. RHE, DHA and 2,3-dihydroxy-2-propenal (2a and 4b) start to form. This is in reasonable agreement with the experimental onset at ∼0.5 V.

At 0.50 V, CO is produced and at 0.60 V, glyceraldehyde (2c) is formed but expected

to stay adsorbed on the surface. All of these reactions occur well below the adsorption of *O and *OH at 1.0 V. As biodiesel production is increasing globally, so is the production of glycerol. Our results show the potential of glycerol electrooxidation on gold as a means of production of valuable chemicals, such as hydrogen gas, glyceraldehyde and dihydroxyacetone.

Acknowledgement The authors gratefully acknowledge support from Formas and the Swedish Research Council and the Röntgen-Ångström project HEXCHEM. The electronic structure calculations were performed on resources provided by the Swedish National Infrastructure for Computing (SNIC) at NSC and UPPMAX. This work used the Extreme Science and Engineering Dis13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

covery Environment (XSEDE), 48 which is supported by National Science Foundation grant number ACI-1053575.

Supporting Information Available The following file is available free of charge. • Supporting information for: Electrooxidation of glycerol on gold in acidic medium: a combined experimental and DFT study This material is available free of charge via the Internet at http://pubs.acs.org/.

References (1) Gerpen, J. V. Biodiesel Processing and Production. Fuel Processing Technology 2005, 86, 1097–1107. (2) OECD/FAO (2016), OECD-FAO Agricultural Outlook 2016-2025, OECD Publishing, Paris. (3) Pratas, M. J.; Freitas, S. V. D.; Oliveira, M. B.; Monteiro, S. C.; Lima, Á. S.; Coutinho, J. A. P. Biodiesel Density: Experimental Measurements and Prediction Models. Energy & Fuels 2011, 25, 2333–2340. (4) Johnson, D. T.; Taconi, K. A. The Glycerin Glut: Options for the Value-Added Conversion of Crude Glycerol Resulting from Biodiesel Production. Environmental Progress 2007, 26, 338–348. (5) Yang, F.; Hanna, M. A.; Sun, R. Value-Added Uses for Crude Glycerol–a Byproduct of Biodiesel Production. Biotechnology for Biofuels 2012, 5, 13.

14

ACS Paragon Plus Environment

Page 14 of 21

Page 15 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(6) Pagliaro, M.; Ciriminna, R.; Kimura, H.; Rossi, M.; Della Pina, C. From Glycerol to Value-Added Products. Angewandte Chemie International Edition 2007, 46, 4434– 4440. (7) Cortright, R. D.; Davda, R. R.; Dumesic, J. A. Hydrogen from Catalytic Reforming of Biomass-Derived Hydrocarbons in Liquid Water. Nature 2002, 418, 964–967. (8) Simões, M.; Baranton, S.; Coutanceau, C. Electrochemical Valorisation of Glycerol. ChemSusChem 2012, 5, 2106–2124. (9) Coutanceau, C.; Baranton, S. Electrochemical Conversion of Alcohols for Hydrogen Production: A Short Overview. Wiley Interdisciplinary Reviews: Energy and Environment 2016, 5, 388–400. (10) Ciriminna, R.; Palmisano, G.; Pina, C. D.; Rossi, M.; Pagliaro, M. One-Pot Electrocatalytic Oxidation of Glycerol to DHA. Tetrahedron Letters 2006, 47, 6993–6995. (11) Greeley, J.; Markovic, N. M. The Road from Animal Electricity to Green Energy: Combining Experiment and Theory in Electrocatalysis. Energy & Environmental Science 2012, 5, 9246–9256. (12) Kirubakaran, A.; Jain, S.; Nema, R. K. A Review on Fuel Cell Technologies and Power Electronic Interface. Renewable and Sustainable Energy Reviews 2009, 13, 2430–2440. (13) Hu, C.; Creaser, D.; Siahrostami, S.; Grönbeck, H.; Ojagh, H.; Skoglundh, M. Catalytic Hydrogenation of CC and CO in Unsaturated Fatty Acid Methyl Esters. Catalysis Science & Technology 2014, 4, 2427–2444. (14) Fukuoka, T.; Habe, H.; Kitamoto, D.; Sakaki, K. Bioprocessing of Glycerol into Glyceric Acid for Use in Bioplastic Monomer. Journal of Oleo Science 2011, 60, 369–373. (15) Beden, B.; Çetin, I.; Kahyaoglu, A.; Takky, D.; Lamy, C. Electrocatalytic Oxidation

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of Saturated Oxygenated Compounds on Gold Electrodes. Journal of Catalysis 1987, 104, 37–46. (16) Kahyaoglu, A.; Beden, B.; Lamy, C. Oxydation Electrocatalitique Du Glycerol Sur Electrodes d’Or et de Platine En Milieu Aqueux. Electrochimica Acta 1984, 29, 1489– 1492. (17) Kwon, Y.; Schouten, K. J. P.; Koper, M. T. M. Mechanism of the Catalytic Oxidation of Glycerol on Polycrystalline Gold and Platinum Electrodes. ChemCatChem 2011, 3, 1176–1185. (18) Schnaidt, J.; Heinen, M.; Denot, D.; Jusys, Z.; Behm, J. R. Electrooxidation of Glycerol Studied by Combined in Situ IR Spectroscopy and Online Mass Spectrometry under Continuous Flow Conditions. Journal of Electroanalytical Chemistry 2011, 661, 250– 264. (19) Cheng, W.; Singh, N.; Maciá-Agulló, J. A.; Stucky, G. D.; McFarland, E. W.; Baltrusaitis, J. Optimal Experimental Conditions for Hydrogen Production Using Low Voltage Electrooxidation of Organic Wastewater Feedstock. International Journal of Hydrogen Energy 2012, 37, 13304–13313. (20) Kwon, Y.; Lai, S. C. S.; Rodriguez, P.; Koper, M. T. M. Electrocatalytic Oxidation of Alcohols on Gold in Alkaline Media: Base or Gold Catalysis? Journal of the American Chemical Society 2011, 133, 6914–6917. (21) Roquet, L.; Belgsir, E. M.; Léger, J. M.; Lamy, C. Kinetics and Mechanisms of the Electrocatalytic Oxidation of Glycerol as Investigated by Chromatographic Analysis of the Reaction Products: Potential and pH Effects. Electrochimica Acta 1994, 39, 2387–2394. (22) Kim, Y.; Kim, H. W.; Lee, S.; Han, J.; Lee, D.; Kim, J.-R.; Kim, T.-W.; Kim, C.U.; Jeong, S.-Y.; Chae, H.-J.; et al., The Role of Ruthenium on Carbon-Supported 16

ACS Paragon Plus Environment

Page 16 of 21

Page 17 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

PtRu Catalysts for Electrocatalytic Glycerol Oxidation under Acidic Conditions. ChemCatChem 2017, 9, 1683–1690. (23) Vitos, L.; Ruban, A. V.; Skriver, H. L.; Kollár, J. The Surface Energy of Metals. Surface Science 1998, 411, 186–202. (24) 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. The Journal of Physical Chemistry B 2004, 108, 17886–17892. (25) Rouya, E.; Cattarin, S.; Reed, M. L.; Kelly, R. G.; Zangari, G. Electrochemical Characterization of the Surface Area of Nanoporous Gold Films. Journal of The Electrochemical Society 2012, 159, K97–K102. (26) Kresse, G.; Hafner, J. Ab Initio Molecular Dynamics for Liquid Metals. Physical Review B 1993, 47, 558–561. (27) Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Computational Materials Science 1996, 6, 15–50. (28) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Physical Review B 1996, 54, 11169–11186. (29) Klimeš, J.; Bowler, D. R.; Michaelides, A. Chemical Accuracy for the van Der Waals Density Functional. Journal of Physics: Condensed Matter 2010, 22, 022201. (30) Klimeš, J.; Bowler, D. R.; Michaelides, A. Van Der Waals Density Functionals Applied to Solids. Physical Review B 2011, 83, 195131. (31) Becke, A. D. On the Large-Gradient Behavior of the Density Functional Exchange Energy. The Journal of Chemical Physics 1986, 85, 7184–7187.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(32) Baltrusaitis, J.; Valter, M.; Hellman, A. Geometry and Electronic Properties of Glycerol Adsorbed on Bare and Transition-Metal Surface-Alloyed Au(111): A Density Functional Theory Study. The Journal of Physical Chemistry C 2016, 120, 1749–1757. (33) Blöchl, P. E. Projector Augmented-Wave Method. Physical Review B 1994, 50, 17953– 17979. (34) Björneholm, O.; Hansen, M. H.; Hodgson, A.; Liu, L.-M.; Limmer, D. T.; Michaelides, A.; Pedevilla, P.; Rossmeisl, J.; Shen, H.; Tocci, G.; et al., Water at Interfaces. Chemical Reviews 2016, 116, 7698–7726. (35) Frydendal, R.; Busch, M.; Halck, N. B.; Paoli, E. A.; Krtil, P.; Chorkendorff, I.; Rossmeisl, J. Enhancing Activity for the Oxygen Evolution Reaction: The Beneficial Interaction of Gold with Manganese and Cobalt Oxides. ChemCatChem 2015, 7, 149–154. (36) Busch, M.; Halck, N. B.; Kramm, U. I.; Siahrostami, S.; Krtil, P.; Rossmeisl, J. Beyond the Top of the Volcano? – A Unified Approach to Electrocatalytic Oxygen Reduction and Oxygen Evolution. Nano Energy 2016, 29, 126–135. (37) Monkhorst, H. J.; Pack, J. D. Special Points for Brillouin-Zone Integrations. Physical Review B 1976, 13, 5188–5192. (38) Diaz-Morales, O.; Calle-Vallejo, F.; de Munck, C.; Koper, M. T. M. Electrochemical Water Splitting by Gold: Evidence for an Oxide Decomposition Mechanism. Chemical Science 2013, 4, 2334–2343. (39) Angerstein-Kozlowska, H.; Conway, B. E.; Hamelin, A.; Stoicoviciu, L. Elementary Steps of Electrochemical Oxidation of Single-Crystal Planes of Au—I. Chemical Basis of Processes Involving Geometry of Anions and the Electrode Surfaces. Electrochimica Acta 1986, 31, 1051–1061.

18

ACS Paragon Plus Environment

Page 18 of 21

Page 19 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(40) Edens, G. J.; Gao, X.; Weaver, M. J. The Adsorption of Sulfate on Gold(111) in Acidic Aqueous Media: Adlayer Structural Inferences from Infrared Spectroscopy and Scanning Tunneling Microscope. Journal of Electroanalytical Chemistry 1994, 375, 357–366. (41) Hansen, H. A.; Rossmeisl, J.; Nørskov, J. K. Surface Pourbaix Diagrams and Oxygen Reduction Activity of Pt, Ag and Ni(111) Surfaces Studied by DFT. Physical Chemistry Chemical Physics 2008, 10, 3722–3730. (42) Peuckert, M.; Coenen, F. P.; Bonzel, H. P. On the Surface Oxidation of a Gold Electrode in 1N H2S04 Electrolyte. Surface Science 1984, 141, 515–532. (43) Angerstein-Kozlowska, H.; Conway, B. E.; Hamelin, A.; Stoicoviciu, L. Elementary Steps of Electrochemical Oxidation of Single-Crystal Planes of Au Part II. A Chemical and Structural Basis of Oxidation of the (111) Plane. Journal of Electroanalytical Chemistry and Interfacial Electrochemistry 1987, 228, 429–453. (44) Rodriguez, P.; Garcia-Araez, N.; Koper, M. T. M. Self-Promotion Mechanism for CO Electrooxidation on Gold. Physical Chemistry Chemical Physics 2010, 12, 9373–9380. (45) Rodriguez, P.; Kwon, Y.; Koper, M. T. M. The Promoting Effect of Adsorbed Carbon Monoxide on the Oxidation of Alcohols on a Gold Catalyst. Nature Chemistry; London 2012, 4, 177–82. (46) Liu, B.; Greeley, J. Decomposition Pathways of Glycerol via C–H, O–H, and C–C Bond Scission on Pt(111): A Density Functional Theory Study. The Journal of Physical Chemistry C 2011, 115, 19702–19709. (47) Liu, B.; Greeley, J. A Density Functional Theory Analysis of Trends in Glycerol Decomposition on Close-Packed Transition Metal Surfaces. The Journal of Physical Chemistry C 2013, 15, 6475–6485.

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(48) Towns, J.; Cockerill, T.; Dahan, M.; Foster, I.; Gaither, K.; Grimshaw, A.; Hazlewood, V.; Lathrop, S.; Lifka, D.; Peterson, G. D.; et al., XSEDE: Accelerating Scientific Discovery. Computing in Science Engineering 2014, 16, 62–74.

20

ACS Paragon Plus Environment

Page 20 of 21

Page 21 ofThe 21 Journal of Physical Chemistry 1 2 3 4 5 6

0.4 V

0.6 V 0.4 V

V ACS Paragon Plus0.5 Environment 0.4 V