Selectivity of Synthesis Gas Conversion to C2+ Oxygenates on fcc(111

SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, Menlo Park , California 94025 , United States. ACS Catal. , 2...
0 downloads 7 Views 3MB Size
Subscriber access provided by Kaohsiung Medical University 2+

Selectivity of synthesis gas conversion to C oxygenates on fcc(111) transition metal surfaces Julia Schumann, Andrew J. Medford, Jong Suk Yoo, Zhi-Jian Zhao, Pallavi Bothra, Ang Cao, Felix Studt, Frank Abild-Pedersen, and Jens K. Norskov ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.8b00201 • Publication Date (Web): 13 Mar 2018 Downloaded from http://pubs.acs.org on March 13, 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 14 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

ACS Catalysis

Selectivity of synthesis gas conversion to C2+ oxygenates on fcc(111) transition metal surfaces Julia Schumanna,b, Andrew J. Medforda,b,†, Jong Suk Yooa,◊, Zhi-Jian Zhaoa,b,‡, Pallavi Bothraa,b, Ang Caoa, Felix Studtb,§, Frank Abild-Pedersenb, and Jens K. Nørskova,b,* a

SUNCAT Center for Interface Science and Catalysis, Department of Chemical Engineering, Stanford University, Stanford, CA 94305, United States. b SUNCAT Center for Interface Science and Catalysis, SLAC National Accelerator Laboratory, Menlo Park, CA 94025, United States. ABSTRACT: Using a combined DFT and descriptor based microkinetic model approach we predict production rate volcanos for higher oxygenate formation on (111) transition metal surfaces. Despite their lower activity for CO conversion compared to stepped surfaces, (111) transition metal surfaces bring the potential for selectivity towards C2+ oxygenates. The volcano plots can be used to rationalize and predict activity and selectivity trends for transition metal based catalysts.

KEYWORDS Density Functional Theory, higher alcohol synthesis, syngas conversion, heterogeneous catalysis, micro-kinetic modelling, probability, catalyst design. INTRODUCTION The conversion of synthesis gas directly to higher-valued chemicals and fuels like ethanol is an extremely desirable process route.1–3 The advantage compared to ‘bioethanol’ produced by fermentation of sugars is that a heterogeneous thermal reaction can make use of all biomass or natural gas instead of competing with food resources.1 Another attractive feature is that if we have a way to produce renewable hydrogen, carbon-neutral fuels can be produced by CO2 recycling.4 The field of research has focused mainly on the development of a catalyst with an acceptable activity and selectivity towards C2+ oxygenates, and the only elemental metal found to exhibit some selectivity towards the desired product is Rh.1–3,5 Previous theoretical studies have shown that undercoordinated sites such as steps on transition metals suffer from

low selectivity towards C2+ oxygenates and experimental evidence that this is correct has been found.5 In the present paper we have focused on the (less active) terrace sites, where an increase in C2+ oxygenate production is believed to occur through a compromise between the lower activity of terrace sites and an enhanced selectivity combined with a higher number of active sites. To realize this, we need to understand the pattern of selectivity on different surface orientations, tie that to our modeling framework, which will then guide our catalyst design efforts and help predict better catalysts for the synthesis of higher oxygenates. In the last two decades, density functional theory (DFT) has become one of the most widely used computational methods

Scheme 1. The considered reaction pathway for the formation of ethanol, acetaldehyde, methanol and methane. Blue arrows represent hydrogenation reactions, green arrows C-O bond dissociation reactions, and red arrows C-C coupling reactions.

ACS Paragon Plus Environment

ACS Catalysis

E OH [ eV]

CH3

Cu 1.5

Pd Pt IrRh

2.0 2.5

2.5 E CH 3 = (0.55 ± 0.06) E CO + (-1.29 ± 0.24) 0.5

3.0

E CH CO = (0.76 ± 0.08) E CO + (-1.09 ± 0.28)

0.0

1.5

1.0

CH-CO

Pd Pt IrRh CO

1.0

0.5

Pd Pt IrRh

1

E CH

1.5

1.0

CH3-H

OH

1.0

E CH 3

IrPRh d Pt

2.0

0.5

Rh

0.5

= (0.88 ± 0.06) E CO + (-0.40 ± 0.23) 1.0

E CH 3 1.0

0.0

Cu

Pd

OH

= (0.36 ± 0.10) E CO + (1.00 ± 0.24) E OH + (0.03 ± 0.28)

0.4

E CO [ eV]

0.2

0.0

0.2

0.4

2.5 CH 2CO

= (1.02 ± 0.03) E CO + (-0.42 ± 0.29) 1.5

1.0

0.5

1.0 1.5

EH

CH 2OH

0.0

1.0

E CO [ eV]

H-CHOH [ eV]

Pt

Rh

0.5 1.0

CH O

Pd

EH

[ eV] CH OH

0.5

EH

CH OH

= (1.16 ± 0.04) E CO + (1.53 ± 0.21)

1.5

1.0

E CO [ eV]

0.5

0.0

Cu

0.5

Pt

IrPd Rh

0.5

0.8

0.6

0.4

E CH 2O [ eV]

0.2

[ eV]

0.0 0.5

CHO =

(0.96 ± 0.05) E CO + (0.17 ± 0.06) E OH + (1.02 ± 0.14)

1.5

1.0

0.5

0.0

0.96E CO + 0.17E OH [ eV]

1.0

H

0.5

0.0

E CH 3O [ eV]

E CO [ eV]

CH3-CO 0.0

Ag

Pt = (1.46 ± 0.08) E CO + (1.05 ± 0.20) 1.0

0.5

0.5

Cu

1.0 1.5

Pt Pd Rh

2.0 2.5 3.0

E CH 3

CO

0.0

= (0.97 ± 0.04) E CO + (-0.53 ± 0.30) 1.5

1.0

0.5

0.0

E CO [ eV]

CH3O-H Ag

1.5

Cu

2.0 2.5

IrPRh d

3.0

Pt

3.5 4.0

E CH3CHOH H = (0.89 ± 0.05) E CO + (0.01 ± 0.03) E OH + (-1.78 ± 0.42) 1.5

1.0

Pt

Rh EH

CH2O =

0.2

0.0

1.5 2.0

3.0

2.5

3.5

Pt

IrPd Rh EH

CH3CO

1.0

0.5

1.0 E CH3O H = (0.21 ± 0.06) E CO + (0.50 ± 0.11) E OH + (-0.55 ± 0.25) 0.1

0.0

1.0

2.0

Ag

Pd

1.5

Pt

IrRh

2.5 3.0

EH

CH CO

= (1.23 ± 0.08) E CO + (0.10 ± 0.25)

1.5

= (1.04 ± 0.06) E CO + (0.10 ± 0.04) E OH + (0.97 ± 0.13) 0.5

0.0

1.04E CO + 0.1E OH [ eV]

1.0

0.5

0.0

HCO-H Ag

Pd

0.3

E CO [ eV]

Cu

Rh

0.2

Cu

H-OH

1.0

0.1

0.5

0.91E CO + 0.11E OH [ eV]

1.5

0.0

0.5

0.0

Pt

Ir

0.5

1.0

Pt

0.8

0.2

= (0.91 ± 0.05) E CO + (0.11 ± 0.03) E OH + (-1.44 ± 0.34)

1.5

Pt

1.5

Rh

0.21E CO + 0.5E OH [ eV]

Cu

0.2

Cu

CO

0.6

H-CHCO

Ag

EH

Pd

0.4

1.2

Ag

H-CO

IrRh Pd

Cu

0.2

0.0

1.0

(0.43 ± 0.06) E CO + (0.41 ± 0.09) E OH + (-0.05 ± 0.24) 0.4

0.5

Ag

0.0

H-CH3CO

1.0 0.5

1.5

0.89E CO + 0.01E OH [ eV]

1.0 EH

E CH 2

0.0

Cu

1.5

Ag

0.0

IrRh

0.43E CO + 0.41E OH [ eV]

1.0

1.0

1.0

0.6

H-CHO Cu

Pt

ECH3CHO H = (0.71 ± 0.09) E CO + (0.04 ± 0.06) E OH + (-1.64 ± 0.40)

0.0

1.5

Ag

1.5

0.0

d IrP Rh

Pd

E CO [ eV]

2.0

1.0

0.5

E CH = (1.87 ± 0.10) E CO + (1.49 ± 0.14)

CH3CHOH-H

Ag

1.0

= (0.85 ± 0.05) E CO + (0.50 ± 0.24)

1.5

Pd

1

1.5

0.5

0.5

Pd Pt IrRh

E CO [ eV]

3.0

0.0

1

3

0.0

0

H-CH2O

Pt Pd Rh

0.5

Cu

0.5

Cu

2.5

[ eV]

3.0E H

1.0

Ag

2

1.0

0

E CHO = (0.92 ± 0.03) ECO + (0.14 ± 0.04) E OH + (0.55 ± 0.11)

0.71E CO + 0.04E OH [ eV]

Cu

0.0

0.9

0.2

1.2

0.5

0.5

0.0

2.0

CH 2O

Pd Ir Rh

0.5

EH

CH 2OH

Pt

1.5 2.0

[ eV]

Ag

EH

[ eV] CH 2CO

EH

Cu

1.0

0.8

Ag

CH2-H

1.5

3.5

Ag

0.5

1.5

0.7

2

1.5

= (1.63 ± 0.09) E CO + (0.87 ± 0.02) E OH + (2.07 ± 0.16) 2.0

0.6

Cu

0.92E CO + 0.14E OH [ eV]

Ag

0.6

0.0

1.0

Pt

IrPd Rh

1.0 1.5

Pt

H-CH2OH

1.0

0.5

CH3CHO-H

0.36E CO + 1.0E OH [ eV]

H-CH2CO

OH

2.5

1.0

Ag

0.0

0.0

0.0

Rh Ir E CH

0.5

1

Cu

1.63E CO + 0.87E OH [ eV]

Pt 0.5

0.5

1

0.0

1.0

[ eV]

Cu

H

[ eV]

0.5

1.5

0.5

CH3-OH Ag

H

1.0

Cu

0

0.4

CH

Ag

E CO [ eV]

0.0

2.5 E CH 3

E CH OH = (1.18 ± 0.05) E CO + (0.90 ± 0.15)

Pd

0.3

2

1

1

= (2.09 ± 0.14) E CO + (2.22 ± 0.24)

H

0.0

E CO [ eV]

1.5

[ eV] OH

1 0

2

= (2.00 ± 0.16) E CO + (1.57 ± 0.22) 1.5

Pt

2

E CH 3O = (1.00 ± 0.12) E OH + (-1.78 ± 0.24) 0.2

E OH [ eV]

0.5

3

H

E CH

Cu

2.0

0.0

Ag

CH-OH Ag

0.5

1.0

E CO [ eV]

CO

2

d IrPRh

1.0

EH

1

1.0

PdAg

Cu

Rh

1.5

CHO

0.5

1.5

E CH

[ eV] H

0

E CH

[ eV] CO

E CH

Cu

1.5

Ir

1.0

E CO [ eV]

0.0

CH-H

1

E CH 3CO = (0.86 ± 0.06) E CO + (-2.07 ± 0.33)

0.35

Cu

0.0

2

3

E CH 3

0.5

3

Ag

2

0.30

Ag

E CO [ eV]

[ eV]

1.0

E CO [ eV]

0.25

0.5

1.5

E CH 3CH O

1.5

0.20

CHOH E CH OH [ eV]

E CH CO [ eV]

Pt

0.15

1.0

Ag

1.0

Cu

1.5

0.10

CHCO

0.5

4.0

E CH 3CHO = (0.02 ± 0.09) E CO + (0.39 ± 0.09) E OH + (-3.36 ± 0.46) 0.05

0.02E CO + 0.39E OH [ eV]

Ag

Pd Rh Ir

3.8

0.0

E CO [ eV]

1.0

2.0

0.5

E CH [ eV]

1.0

CO

1.5

[ eV]

E CH 3CH OH = (0.74 ± 0.06) E CO + (-2.41 ± 0.44)

E CH 3

4.5

1.0

Pd IrRh Pt

3.5

[ eV]

0.9

3.0

0.0

Pt

H

0.8

3.6

Cu

2.5

E CH 3O

0.7

3.4

0.5

CH3O

[ eV]

0.6

4.0

Ag

Pd

Rh

3.2

1.0

0.5

CH CO

0.5

Pt

1.5

Ag

2.0

H

0.4

d IrPRh

E CH 2 = (1.11 ± 0.08) E CO + (0.26 ± 0.17)

E CO [ eV]

H

0.3

3.0

Ir

Cu

Pd IrRh Pt

1.5

CH3CO

1.5

Pt

2.8

1.0

0.15 0.10 0.05 0.00 0.05 0.10 0.15

2.6

3.0

Cu 0.5

0.13E CO + 0.25E OH [ eV]

2.0

[ eV]

E CH 3CH2OH = (0.05 ± 0.09) E OH + (-4.17 ± 0.53) 0.2

Cu

2.5

3.5

1.2 E CH2O = (0.13 ± 0.10) E CO + (0.25 ± 0.12) E OH + (-0.72 ± 0.23) 0.0

CH3CHO Ag

Ag

0.0

2.0

H

4.2

0.5

E CH 3CO [ eV]

Pt

4.4 4.6

1.0

CH3CHOH

Ag Pd

E CH 3CH OH [ eV]

E CH 3CH 2OH [ eV]

Ir

Rh

4.0

1.5

1.0

E CO [ eV]

2.0

Cu

3.8

2.0 E CH 2OH = (0.77 ± 0.09) E CO + (-0.18 ± 0.22) 0.0

E CH O [ eV]

CH3CH2OH 3.6

E CH 3 [ eV]

0.5

Rh

E H CO

1.0

E CO [ eV]

Ir

0.8

[ eV]

1.5

Ag

EH

E CH 2CO = (0.77 ± 0.06) E CO + (-1.41 ± 0.29)

Cu Pt

Pd 0.6

E CH 2

3.5 0.0

[ eV]

0.5

E CO [ eV]

Pt

1.5

E CH 3CH OH

1.0

Pt

Pd Rh

[ eV]

1.5

2.5 3.0

0.5 1.0

CH 3CO

E H = (0.41 ± 0.03) E CO + (0.35 ± 0.06)

0.6

Pd IrRh

0.4

EH

Pt

Rh Pd

2.0

Cu

CH2 0.5

Ag

OH

0.2 0.4

Ag

E CH 3CH O [ eV]

Ir

1.5

CH2O

0.2

0.0

[ eV]

Cu 0.0

Cu

E CH 2OH [ eV]

E CH 2CO [ eV]

E H [ eV]

0.2

CH2OH

1.0

E CH 2 [ eV]

CH2CO Ag

EH

H 0.4

EH

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 2 of 14

Ag

1.5

Cu

1.0 0.5 0.0 0.5

Pd Rh Ir

Pt

1.0 EH

OH

= (0.28 ± 0.05) E CO + (0.69 ± 0.12) E OH + (1.01 ± 0.12) 0.2

0.0

0.2

0.4

0.28E CO + 0.69E OH [ eV]

1.5

E H CO

H

= (1.29 ± 0.07) E CO + (1.57 ± 0.20) 1.5

1.0

0.5

0.0

E CO [ eV]

Figure 1: Formation energies of intermediates and transition states are plotted as a function of CO binding energy (∆ECO*) and / or OH binding energy (∆EOH*). Scaling parameters are given in the insets. Errorbars and uncertainties of scaling parameters represent the standard deviation of the DFT binding energy determined from the BEEF ensemble.

to study heterogeneously catalyzed processes.6 The atomic level information provided by DFT calculations offers insights into the mechanistic aspect of catalytic reactions, to a degree that currently is inaccessible experimentally. In addition, recent developments of new and more accurate exchange correlation functionals, combined with an exponential increase in computational speed have enhanced the quality and the quanti-

ty of studied systems, resulting in significant progress in screening studies for new catalytic materials based on DFT.7–9 The accuracy of modern DFT functionals is still not satisfactory. A simple estimation for a barrier with an ±0.2 eV uncertainty would correspond to a four-order-of-magnitude uncertainty in the rate constant at 523 K for this single step. The success of DFT calculations for catalysis comes from the abil-

ACS Paragon Plus Environment

Page 3 of 14 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

ACS Catalysis

ity to reliably predict trends. For a complex reaction network, it is difficult to predict how the energy uncertainties propagate through multiple layers of simulations, from DFT calculated barriers to the rate estimated by e.g. micro-kinetic modelling. As shown in a previous study by Medford et al.,10 although the rate and selectivity volcano plots for the synthesis gas conversion to higher alcohols were successfully constructed, the error bars on the calculated descriptor properties for different transition metals can be significant. This challenge was addressed in a recent study by Medford et al.11 It was shown based on a Bayesian error estimation scheme that the predicted error on calculated rates can be reduced because the errors of DFT-calculated energies are correlated with each other. In other words, a more accurate prediction on the catalytic rate, as well as the selectivity, can be made if one takes the correlation between the errors in DFT calculations into account. Based on this idea, we can combine the Bayesian error estimation ensemble analysis with the widely-used volcano plot procedure for catalyst screening. Unlike the descriptor-rate or descriptor-selectivity volcano plot, a new and more reliable descriptor-probability volcano plot can be generated, in which the correlation between the errors in DFT calculations are considered. METHODS DFT calculations are performed with the planewave based Quantum ESPRESSO code.12 We employ the generalized-gradient approximation applying the van der Waals corrected exchange-correlation functional BEEF-vdW.13 Details about the calculations and methods used here can be found in the Supporting Information.

The micro-kinetic modeling was carried out using the CatMAP code14 that applies a mean-field approach and the steady-state approximation including a self-consistent description of adsorbate-adsorbate interactions. Two surface sites are included in the simulation: a “hydrogen reservoir” site and a site for all other intermediates.5 The reasoning for the two site model is that hydrogen only interacts weakly with other adsorbates and we assume that it does not have to compete for adsorption sites with any of the other adsorbates. We adopt the reaction mechanism as suggested for a Rh(111) surface by Yang et al.5 In this model, C2+ oxygenates are formed through the coupling of CO and CH, the latter of which is formed by splitting the C-O bond in CHOH. (see Scheme 1) To widen the applicability of our model, two pathways for methanol formation were added. Additionally, a pathway was added where CO bond splitting occurs from CH3O and C-C coupling proceeds via CO insertion to CH3 as it was shown for (211) surfaces that it is a relevant pathway for higher alcohol formation.10 (see Scheme 1) All 37 intermediates and transition states were calculated on five or six fcc(111) transition metal surfaces. To describe the thermodynamic and kinetic properties of the reaction pathways we use that adsorption energies of surface species on a given surface orientation can be scaled with certain reaction intermediates or descriptors.15 These so-called scaling relations were already previously successfully used to describe the reactions for methane and methanol formation8,16, and also more complex reaction networks as higher alcohol

Figure 2. Rate volcano plots for a) ethanol, b) acetaldehyde, c) methanol and d) methane production on (111) transition metal surfaces as a function of the descriptors CO binding energy (∆ECO*) and OH binding energy (∆EOH*) at 523 K and 20 bar (CO:H2 1:2,  =  =  =  =  = 10-19 bar). The included error bars show the uncertainty in the DFT binding energies as determined from the standard deviation of the BEEF ensemble calculations. CO and OH binding energies are formation energies of the adsorbed species relative to CO, H2 and H2O gas phase energies.

ACS Paragon Plus Environment

ACS Catalysis 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

synthesis on (211) transition metal surfaces.10 We identified the binding energy of CO and OH as good descriptors for all adsorption energies (due to scaling relations) with a reasonable mean average error (MAE) of 0.15 eV. Thus, a two-descriptor based model was chosen to build a micro-kinetic model of the CO hydrogenation on (111) transition metal surfaces. (See Figure 1 for scaling relations) The uncertainty on the calculated adsorption energies and the effect on the scaling relations is estimated using the BEEFvdW error estimation ensemble.11,13 Due to error correlations, the deviation in the slope of the individual scaling lines become small. The average standard deviation of the slope is 0.07 and the standard deviation of the intercept is 0.25 eV. (Figure 1) It was shown previously that including adsorbate-adsorbate interactions is important for surfaces that bind intermediates strongly.16 In our model we specifically included adsorbateadsorbate effects from co-adsorbed CO on the Rh(111) surface.5 Including adsorbate-adsorbate interaction limits the CO coverage to a maximum of ~0.5 ML at typical reaction conditions. (See Supporting Information for coverage plots) By preventing CO poisoning on strongly adsorbing surfaces, empty sites are available which enable CO dissociation. Additionally, adsorbate-adsorbate self-interaction was included for CHO*, CHOH*, CH* and OH* adsorbates. Those effects are not relevant for the maxima of the rate volcano for acetaldehyde production, but lead to changes of the strong binding side of the methane volcano. RESULTS AND DISCUSSION In qualitative agreement with what has previously been observed on the more active stepped fcc(211) transition metal surfaces10, we see a maximum in the rate of methane formation at strong CO and OH binding energies. (See Figure 2) We note that the maximum is shifted to stronger CO binding energies compared to the stepped sites to compensate for the lower CO dissociation activity of terrace sites. The dependence on OH binding energy is less strong close to the maximum of the methane rate volcano. The volcano for the methanol production rate shows two maxima. The maximum located in the region of weak hydroxyl and strong carbon monoxide binding corresponds to a pathway on which C-O bond dissociation is slow and methanol is formed via hydrogenation of CHOH to CH2OH. This is in agreement with the reaction pathway that has been found for Pd(111).17 The second maximum, which is more relevant for the currently applied copper-based methanol synthesis catalysts is located at weaker CO and stronger OH binding energies. Here, methanol formation proceeds via hydrogenation of CHO to CH2O and further to methanol as was reported previously for stepped surfaces.10 Of course, in the industrial-like copper-based catalyst the activity towards methanol will be dominated by the steps. However, the first pathway has not been considered before for the reaction network on (211) transition metal surfaces.10 New candidates for highly active methanol synthesis catalysts might be found in this region of the descriptor space. However, the challenge will be to identify new materials with strong CO, but weak OH binding energy, compared to e.g. Pd, Ir or Rh. Many oxide promoters result in the opposite, namely strengthening of oxygen binding, with only a weak effect on CO binding.18 The unusual, sharp change of the rate towards the strong binding side of the volcanos in Figure 2b)-2d) is a result of the inclusion of adsorbate-adsorbate interactions. In those regions

Page 4 of 14

of the volcanos, there is a sharp increase in OH*, CH* or CH2* coverage which leads to a non-continuous change in adsorption energies and consequently rates. As can be seen from Figure 3, the region of increased acetaldehyde selectivity is a narrow window between the regions with high methane and high methanol selectivity. In contrast to the previous volcano plot developed for the (211) transition metal surfaces, this narrow region is now shifted to stronger CO-binding energies and closer to pure transition metal surfaces. This provides a possible explanation for the observed selectivity towards higher C2+ oxygenates on Rh.19 Although the TOF is several orders of magnitudes higher on step sites, the selectivity towards higher oxygenates suffers from the fast C-O bond dissociation. On terraces the rates are much lower,

Figure 3. Selectivity towards acetaldehyde for fcc (111) transition metal surfaces as a function of the descriptors CO binding energy and OH binding energy at the same conditions as in Figure 2. Points with white labels were used to construct the scaling relations, while orange labels belong to metals and binary alloys for which the CO and OH descriptor energies only were calculated.

but due to the more difficult C-O bond splitting, higher oxygenate production can compete with methane formation. This structure sensitivity of Rh was previously reported and explains the observed inverse relationship of rate vs. higher oxygenate selectivity.5 Additionally, our calculations show that Co(111) may be a promising candidate with selectivity towards higher oxygenates. As for Rh, we propose that if the steps can be blocked, Co should show similar or higher selectivity towards higher oxygenates. Although pure Co catalysts behave as a traditional Fischer-Tropsch catalyst associated with high alkane selectivity, addition of e.g. sodium leads to an increase of higher oxygenate selectivity.20 Also for CuCobased21,22 or PtCo-based23 catalysts the selectivity is tuned towards oxygenated products, according to literature. The activity is most likely dominated by the alloyed step, but another possible explanation for the role of Cu or Pt respectively could be to block the hydrocarbon-selective Co step sites, such that Co terraces are responsible for the observed higher oxygenate productivity. For certain Co-rich alloys, such as Co3Pt and Co3Fe, we predict that also the (111) terrace could show some selectivity towards C2+ oxygenates. However, the more important role must be blocking of the Co steps to prevent the much faster hydrocarbon formation there. Most Cu-based alloy terraces are not active enough for higher oxygenate formation (see Figure 3). This will most likely be different on the

ACS Paragon Plus Environment

Page 5 of 14 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

ACS Catalysis

more active Cu-alloy step, in agreement with our earlier work on stepped surfaces.10 One Cu-based alloy terrace, CuNi (111), seems promising in terms of higher oxygenate selectivity. Our micro-kinetic model does estimate a very low rate for ethanol formation (max. TOF = 2.0 * 10-6 s-1), while the formation of acetaldehyde is appreciable (max. TOF = 0.02 s-1).

Figure 4. Acetaldehyde selectivity volcano for fcc (111) surfaces at increasing product pressures corresponding to simulated increase of CO conversion.

This is in qualitative agreement with previous results on Rh(111).5 One reason for the low ethanol rate is that the simulated conditions correspond to 0 % conversion, with extremely small background pressures of the products. That leads to a very low free energy of acetaldehyde in the gas phase, thus the desorption of acetaldehyde is favorable vs. further hydrogenation towards ethanol. At higher product pressures we see an

extremely fast equilibration between acetaldehyde and ethanol on surfaces with a strong CO adsorption energy. The equilibrium pressures between ethanol and acetaldehyde differ by about an order of magnitude, depending on the hydrogen gas pressure. Thus, it is likely that the alcohol will dominate over the aldehyde in the product distribution of the oxygenated species. The volcano maximum for the acetaldehyde selectivity from CO hydrogenation on the other hand does not shift with increased conversion. (See Figure 4) Furthermore, experimental results have shown, that for Rh-based catalysts Febased impurities or promoters are responsible for a higher ethanol/acetaldehyde ratio.5,24 That means that the intrinsic Rh(111) surface is selective towards acetaldehyde. Other experimental studies have found a particle size dependency for the ethanol/acetaldehyde ratio.25 However, the most important and difficult step for higher oxygenates is the C-C coupling while maintaining the oxygen functionality. Hydrogenation of acetaldehyde to the alcohol is comparably easy and not the limiting factor for the reaction. When doing this kind of analysis, it is important to keep in mind the uncertainties and possible errors introduced into the modeling.26–28 Especially when evaluating selectivity small differences in the transition state energies can result in very large differences in the product selectivity. On one hand, there are errors that come from the DFT calculated formation energies, typically around ~0.2 eV, on the other hand there are deviations from the scaling relations that can lead to differ-

Figure 5. Probability plots for fcc (111) surfaces. Plotted is the probability, that the TOF towards a) acetaldehyde, b) methanol and c) methane is higher than 10-7 s-1. The second row shows the probability that the selectivity towards d) acetaldehyde, e) methanol and f) methane is higher than 10 %.

ACS Paragon Plus Environment

ACS Catalysis 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

ences between predictions from the volcano plot and single surface calculations.27 The uncertainties associated with the DFT energy have been evaluated using the BEEF ensemble averages. By solving the micro-kinetic model for an ensemble of 350 formation energies for each adsorbate and gas phase species included in the model, we checked how robust the predictions are. More specifically, with this procedure we assessed the error associated with the exchange-correlation part of the DFT energies. All other possible sources of error, due to the surface model, reaction mechanism, and descriptor-based dimensionality reduction are not included in this uncertainty estimation.11 We are analyzing the ensemble of solved micro-kinetic models by plotting the probability, that a certain point on the map has a rate higher than a threshold TOF, or selectivity respectively. The results are shown in Figure 5. We find that the maxima of the production rate and the selectivity volcanos agree with our original prediction. This is mainly due to the correlation of DFT errors and as a result provides confidence in the outcome of our predictions. CONCLUSION Our results demonstrate the structure sensitivity of transition metal catalysts for higher oxygenate synthesis. Despite the lower activity of (111) transition metal surfaces for CO conversion, different step-to-terrace ratios can not only influence the overall turnover frequency, but also lead to a significant change in selectivity for a given material. The C2+ oxygenate volcano maximum for stepped (211) surfaces was not close to a pure transition metal surface; however, for (111) surfaces the maximum is close to Rh, a well-known catalyst with some selectivity towards oxygenates. Furthermore, we predict that a step-free Co surface also should exhibit an increased selectivity towards higher oxygenates, although we note that even a few defects could change this picture since Co(211) is highly active for the production of methane and other hydrocarbons. This new insight into the activity and selectivity pattern for transition metal terraces can lead to a better understanding of existing catalysts for higher oxygenate synthesis and hence a knowledge-based design of better catalysts in the future. We propose that step blocking by metal oxides on transition metals such as Rh and Co will result in more selective catalysts for higher oxygenate synthesis and our results help understand one of the most important functions of promoter oxides added to Fischer-Tropsch and Rhbased catalysts1,20,29,30

ASSOCIATED CONTENT

†Andrew J. Medford: Georgia Institute of Technology, Atlanta, Georgia 30316 (United States) ◊ Jong Suk Yoo: Department of Mechanical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Ave, Cambridge, MA, 02139, USA ‡ Zhi-Jian Zhao: Key Laboratory for Green Chemical Technology of the Ministry of Education, School of Chemical Engineering, Tianjin University; Collaborative Innovation Center of Chemical Science and Engineering, Tianjin 300072, China § Felix Studt: Institute of Catalysis Research and Technology, Karlsruhe Institute of Technology, Hermann-vonHelmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany, and Institute for Chemical Technology and Polymer Chemistry, Karlsruhe Institute of Technology, Engesserstr. 18,
76131 Karlsruhe, Germany. Author Contributions All authors have given approval to the final version of the manuscript.

Funding Sources We gratefully acknowledge support from the U.S. Department of Energy, Office of Basic Energy Sciences to the SUNCAT Center for Interface Science and Catalysis through the SUNCAT-FWP. JS thanks the German Academic Exchange Service (DAAD) for a post-doc fellowship.

ACKNOWLEDGMENT We acknowledge generous computing resources at Carbon HighPerformance Computing Cluster at Argonne National Laboratory.

ABBREVIATIONS DFT, Density Functional Theory; BEEF Bayesian Error Estimation functional; TOF, Turnover Frequency;

REFERENCES (1)

(2)

Supporting Information The Supporting Information is available free of charge on the ACS Publications website. Detailed information about the computational methods as well as DFT energies used as input for the micro-kinetic modelling can be found in the supporting information. (PDF)

(3)

(4)

AUTHOR INFORMATION Corresponding Author

Page 6 of 14

(5)

* Email: [email protected].

Present Addresses (6)

Luk, H. T.; Mondelli, C.; Ferré, D. C.; Stewart, J. A.; Pérez-Ramírez, J.; Ting, H.; Ho, L.; Luk, T. Status and Prospects in Higher Alcohols Synthesis from Syngas. Chem. Soc. Rev 2017, 46, 1358–1426. Spivey, J. J.; Egbebi, A.; Chang, L.; Zhu, Z. B.; Ni, Y. H.; Cheng, L. J.; Yu, F.; Bruhwiler, P. A.; Baumer, M.; Martensson, N.; et al. Heterogeneous Catalytic Synthesis of Ethanol from Biomass-Derived Syngas. Chem. Soc. Rev. 2007, 36, 1514. Subramani, V.; Gangwal, S. K. A Review of Recent Literature to Search for an Efficient Catalytic Process for the Conversion of Syngas to Ethanol. Energy & Fuels 2008, 22, 814–839. Perathoner, S.; Centi, G. CO 2 Recycling: A Key Strategy to Introduce Green Energy in the Chemical Production Chain. ChemSusChem 2014, 7, 1274–1282. Yang, N.; Medford, A. J.; Liu, X.; Studt, F.; Bligaard, T.; Bent, S. F.; Nørskov, J. K. Intrinsic Selectivity and Structure Sensitivity of Rhodium Catalysts for C 2+ Oxygenate Production. J. Am. Chem. Soc. 2016, 138, 3705–3714. Burke, K. Perspective on Density Functional Theory. J.

ACS Paragon Plus Environment

Page 7 of 14 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

(7)

(8)

(9) (10)

(11)

(12)

(13)

(14)

(15)

(16)

(17)

(18)

ACS Catalysis Chem. Phys. 2012, 136, 150901. Sehested, J.; Larsen, K. E.; Kustov, A. L.; Frey, A. M.; Johannessen, T.; Bligaard, T.; Andersson, M. P.; Nørskov, J. K.; Christensen, C. H. Discovery of Technical Methanation Catalysts Based on Computational Screening. Top. Catal. 2007, 45, 9–13. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C. F.; Hummelshøj, J. S.; Dahl, S.; Chorkendorff, I.; Nørskov, J. K. Discovery of a Ni-Ga Catalyst for Carbon Dioxide Reduction to Methanol. Nat. Chem. 2014, 6, 320–324. Greeley, J.; Mavrikakis, M. Alloy Catalysts Designed from First Principles. Nat. Mater. 2004, 3, 810–815. Medford, A. J.; Lausche, A. C.; Abild-Pedersen, F.; Temel, B.; Schjødt, N. C.; Nørskov, J. K.; Studt, F. Activity and Selectivity Trends in Synthesis Gas Conversion to Higher Alcohols. Top. Catal. 2014, 57, 135–142. Medford, A. J.; Wellendorff, J.; Vojvodic, A.; Studt, F.; Abild-Pedersen, F.; Jacobsen, K. W.; Bligaard, T.; Nørskov, J. K. Assessing the Reliability of Calculated Catalytic Ammonia Synthesis Rates. Science (80-. ). 2014, 345, 197–200. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G. L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys. Condens. Matter 2009, 21, 395502. 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, 235149. Medford, A. J.; Shi, C.; Hoffmann, M. J.; Lausche, A. C.; Fitzgibbon, S. R.; Bligaard, T.; Nørskov, J. K. CatMAP: A Software Package for Descriptor-Based Microkinetic Mapping of Catalytic Trends. Catal. Letters 2015, 145, 794–807. Abild-Pedersen, F.; Greeley, J.; Studt, F.; Rossmeisl, J.; Munter, T. R.; Moses, P. G.; Skúlason, E.; Bligaard, T.; Nørskov, J. K. Scaling Properties of Adsorption Energies for Hydrogen-Containing Molecules on Transition-Metal Surfaces. Phys. Rev. Lett. 2007, 99, 16105. Lausche, A. C.; Medford, A. J.; Khan, T. S.; Xu, Y.; Bligaard, T.; Abild-Pedersen, F.; Nørskov, J. K.; Studt, F. On the Effect of Coverage-Dependent Adsorbate– adsorbate Interactions for CO Methanation on Transition Metal Surfaces. J. Catal. 2013, 307, 275–282. Lin, S.; Ma, J.; Ye, X.; Xie, D.; Guo, H. CO Hydrogenation on Pd(111): Competition between Fischer–Tropsch and Oxygenate Synthesis Pathways. J. Phys. Chem. C 2013, 117, 14667–14676. Mehta, P.; Greeley, J.; Delgass, W. N.; Schneider, W. F. Adsorption Energy Correlations at the Metal–Support

(19)

(20)

(21)

(22)

(23)

(24)

(25)

(26)

(27)

(28)

(29)

(30)

Boundary. ACS Catal. 2017, 7, 4707–4715. We Note That There Are Some Minor Differences between the Previously Published Rh(111) Acetaldehyde Selectivity and the Results from This Study Which Originate from Fitting the Adsorption Energies with Linear Scaling Relations. Those Small Deviations from from the linear fit lead to the observed differences. The overall trends can be trusted as is demonstrated by the uncertainty estimation. Ishida, T.; Yanagihara, T.; Liu, X.; Ohashi, H.; Hamasaki, A.; Honma, T.; Oji, H.; Yokoyama, T.; Tokunaga, M. Synthesis of Higher Alcohols by Fischer– Tropsch Synthesis over Alkali Metal-Modified Cobalt Catalysts. Appl. Catal. A, Gen. 2013, 458, 145–154. Prieto, G.; Beijer, S.; Smith, M. L.; He, M.; Au, Y.; Wang, Z.; Bruce, D. A.; De Jong, K. P.; Spivey, J. J.; De Jongh, P. E. Design and Synthesis of Copper-Cobalt Catalysts for the Selective Conversion of Synthesis Gas to Ethanol and Higher Alcohols. Angew. Chemie - Int. Ed. 2014, 53, 6397–6401. Smith, M. L.; Kumar, N.; Spivey, J. J. CO Adsorption Behavior of Cu/SiO 2 , Co/SiO 2 , and CuCo/SiO 2 Catalysts Studied by in Situ DRIFTS. J. Phys. Chem. C 2012, 116, 7931–7939. Christensen, J. M.; Medford, A. J.; Studt, F.; Jensen, A. D. High Pressure CO Hydrogenation over Bimetallic PtCo Catalysts. Catal. Letters 2014, 144, 777–782. Haider, M. A.; Gogate, M. R.; Davis, R. J. Fe-Promotion of Supported Rh Catalysts for Direct Conversion of Syngas to Ethanol. J. Catal. 2009, 261, 9–16. Arakawa, H.; Takeuchi, K.; Matsuzaki, T.; Sugi, Y. Effect of Metal Dispersion on the Activity and Selectivity of Rh/SiO2 Catalyst for High Pressure CO Hydrogenation. Chem. Lett. 1984, 13, 1607–1610. Ulissi, Z.; Prasad, V.; Vlachos, D. G. Effect of Multiscale Model Uncertainty on Identification of Optimal Catalyst Properties. J. Catal. 2011, 281, 339– 344. Sutton, J. E.; Vlachos, D. G. Effect of Errors in Linear Scaling Relations and Bronsted-Evans-Polanyi Relations on Activity and Selectivity Maps. J. Catal. 2016, 338, 273–283. Sutton, J. E.; Guo, W.; Katsoulakis, M. A.; Vlachos, D. G. Effects of Correlated Parameters and Uncertainty in Electronic-Structure-Based Chemical Kinetic Modelling. Nat. Chem. 2016, 8, 331–337. Johnson, G. R.; Bell, A. T. Role of ZrO2 in Promoting the Activity and Selectivity of Co-Based Fischer-Tropsch Synthesis Catalysts. ACS Catal. 2016, 6, 100–114. Li, Z.; Zhong, L.; Yu, F.; An, Y.; Dai, Y.; Yang, Y.; Lin, T.; Li, S.; Wang, H.; Gao, P.; et al. Effects of Sodium on the Catalytic Performance of CoMn Catalysts for Fischer–Tropsch to Olefin Reactions. ACS Catal. 2017, 7, 3622–3631.

ACS Paragon Plus Environment

ACS Catalysis 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 8 of 14

TOC Graphic

ACS Paragon Plus Environment

8

Page 9 of 14 1 2 3 4 5 6 7 8 9 10 11 12

ACS Catalysis

ACS Paragon Plus Environment

1.5

0.0

1.5

Ir

Ag Pd

Pt

1.5

0.5

0.6

0.7

0.8

0.9

2.5 3.0 3.5

d IrPRh

Pt

4.0 4.5

1.5

1.0

E CHCO [eV]

Cu

Pt

0.5

E CH3 = (0.55 ±0.06) E CO + (-1.29 ±0.24) 1.0

0.5

Pd Pt IrRh

1.0

0.0 0.5

1.5

0.5

d IrPRh

Pt

E CHOH = (1.18 ±0.05) E CO + (0.90 ±0.15)

0.0

1.5

1.0

0.3

0.4

0.5

0.6

0.7

0.5 1.0

0.0

Pt

IrPd Rh

0.5

Ag Cu

1

Pd Pt IrRh

E CH2 [eV]

E CH = (1.87 ±0.10) E CO + (1.49 ±0.14) 3

0.0

1.5

0.5

E CO [eV]

E CO [eV]

CH3-H

CH3-OH

1.0

0.5

0.0

Pt

Rh

0.5 E CH3−OH = (0.36 ±0.10) E CO + (1.00 ±0.24) E OH + (0.03 ± 0.28) 1.0

0.0

0.4

0.2

0.0

0.2

0.4

1.0

Pt

E H−CH2OH [eV]

Ag

1.0

0.5

3.0E H−CH2CO = (1.02 ±0.03) E CO + (-0.42 ±0.29) 1.0

0.5

1.5

Pt

Ag Cu

Pd Rh

Pt

1.0

E H−CHOH = (1.16 ±0.04) E CO + (1.53 ±0.21) 1.0

E CO [eV]

0.5

0.0

E CH3CHO−H = (0.71 ±0.09) E CO + (0.04 ±0.06) E OH + (-1.64 ± 0.40) 1.0

0.8

0.6

0.4

0.2

E CH3−CO [eV]

Cu

IrPRh d Pt

3.5 E CH3CHOH−H = (0.89 ±0.05) E CO + (0.01 ±0.03) E OH + (-1.78 ± 0.42) 1.5

1.0

1.0

0.0

0.5

0.5

CH3O-H

0.2

Ag

2.5

4.0

1.5

E CO [eV]

2.0

0.0

Pt

E H−CHO = (0.96 ±0.05) E CO + (0.17 ±0.06) E OH + (1.02 ± 0.14) 1.0

0.5

0.5

0.0

0.96E CO + 0.17E OH [eV]

Pt

Rh

Ag

0.0

Cu

0.2

Pd

0.4 0.6

Pt

Rh

0.8 1.0 1.2 E CH3O−H = (0.21 ±0.06) E CO + (0.50 ±0.11) E OH + (-0.55 ± 0.25) 0.2

0.0

0.6

0.4

0.2

0.0

0.1

0.0

0.1

0.2

0.3

Cu

2.0 2.5 3.0 3.5

Pt

IrPd Rh

0.5

E H−CH3CO = (0.91 ±0.05) E CO + (0.11 ±0.03) E OH + (-1.44 ± 0.34) 1.5

1.0

Cu

0.5 1.0 1.5 2.0 2.5 3.0

0.0

0.5

Ag

0.0

Pd IrRh

Pt

E H−CHCO = (1.23 ±0.08) E CO + (0.10 ±0.25) 1.5

1.0

0.43E CO + 0.41E OH [eV]

0.91E CO + 0.11E OH [eV]

E CO [eV]

H-CO

H-OH

HCO-H

Ag

Ag Cu

0.5 0.0

1.0

Ag

1.5

0.2

1.0

0.5

1.0

E H−CH2O = (0.43 ±0.06) E CO + (0.41 ±0.09) E OH + (-0.05 ± 0.24)

1.5

Cu

1.5

Cu Pd

1.0

Ag

IrPd Rh

Ag

0.0

0.0

0.5

0.5

1.0

3.0

0.5

E H−CH2OH = (0.85 ±0.05) E CO + (0.50 ±0.24)

1.0

0.5

Pt

0.0

1.5

3.0

E CH3−CO = (0.97 ±0.04) E CO + (-0.53 ±0.30)

H-CHCO

1.5

0.0

d IrPRh

2.5

H-CH3CO

1.5

E H−CHO [eV]

0.5

Cu

Pd Rh

H-CH2O

H-CHO Cu

Ag

2.0

H-CH2OH

H-CHOH 1.0

CH3CHOH-H

Pt

1.5

3.0

0.5

CH3CHO-H

2.0

3.5

1.0

E CO [eV]

1.5

2.5

1.5

Cu

1.0

0.21E CO + 0.5E OH [eV]

E CO [eV]

1.5

0.5

0.5

0.89E CO + 0.01E OH [eV]

0.0

Ag

0.0

0.5

Ag

0.71E CO + 0.04E OH [eV]

E CO [eV] 2.0

1.0

E CH2−H = (1.46 ±0.08) E CO + (1.05 ±0.20)

0.0

0.36E CO + 1.0E OH [eV]

0.0

1.0

1.5

Pd Pt IrRh

1.63E CO + 0.87E OH [eV]

1.2

0.5

0.5

2.0

1 2

E CH−OH = (1.63 ±0.09) E CO + (0.87 ±0.02) E OH + (2.07 ± 0.16) 2.5

0.6

0.0 0.5

Ag

Cu

Pd

1.0

Cu

0.0

Pt

Rh Ir

1.0

1.0 0.5

0 1

E CH−H = (2.09 ±0.14) E CO + (2.22 ±0.24) 1.0

Pd

0

E CH3O−H [eV]

Pd Pt Rh Ir

1

Cu

E H−CHCO [eV]

1

Cu

1

IrRh Pd

Pt

Cu

1.5

Pd 1.0

Rh

Pt

Ir

0.5 E H−CO = (1.04 ±0.06) E CO + (0.10 ±0.04) E OH + (0.97 ± 0.13) 1.5

1.0

0.5

1.04E CO + 0.1E OH [eV]

ACS Paragon Plus Environment

0.0

E H−OH = (0.28 ±0.05) E CO + (0.69 ±0.12) E OH + (1.01 ± 0.12) 0.2

0.0

0.2

0.4

0.28E CO + 0.69E OH [eV]

0.5

2.0

E HCO−H [eV]

0

E CH2−H [eV]

1

2

E CH3CHOH−H [eV]

Cu

0.0

Ag

Ag

1.0

0

E CHO = (0.92 ±0.03) E CO + (0.14 ±0.04) E OH + (0.55 ± 0.11) 1.0

0.9

2

2

1.5

0.8

1

Cu

0.0

1.5

0.5

3

E CH−OH [eV]

E CH−H [eV]

Ag

1.5

E CH3−OH [eV]

E CH−CO [eV] E CH3−H [eV]

0.2

Ag

0.5

Cu

0.5

1.0

1.0

Ag

1.0

E CHCO = (0.76 ±0.08) E CO + (-1.09 ±0.28)

0.0

H-CH2CO E H−CH2CO [eV]

0.0

CH

2

E CO [eV]

E H−CHOH [eV]

0.5

E CH3O = (1.00 ±0.12) E OH + (-1.78 ±0.24)

CHO

E CH3CHO−H [eV]

0.5

IrPRh d Pt

1.5

1.0

2.0

CHOH Cu

3

2.5 E CH3−H = (0.88 ±0.06) E CO + (-0.40 ±0.23)

1.0

1.5

1.5

Cu

CHCO

E H−CH2O [eV]

1.0

Cu

Rh

0.35

Rh

PdAg

E OH [eV]

2.5

2

2.0

0.5

E CH3CO = (0.86 ±0.06) E CO + (-2.07 ±0.33)

Ir

1.0

CH3-CO

Ag

Pd

Pd IrRh Pt

Pt

CH2-H

1.0

0.0

Cu

0.0

0.5

Ag

CH-OH

E CH−CO = (2.00 ±0.16) E CO + (1.57 ±0.22)

1.5

CH3O

CH-H

0.5

2.5

CH3CO

CH-CO

0.0

2.0

4.0

0.5

E CO [eV]

2.5

3.5

1.0

0.13E CO + 0.25E OH [eV]

2.0

3.0

1.5

E CO [eV]

Pd Pt IrRh

Pd Ir Rh

E CH2 = (1.11 ±0.08) E CO + (0.26 ±0.17)

0.92E CO + 0.14E OH [eV]

0

1.5

2.0

E CO [eV]

Cu

1.5

0.30

1.5

E CO [eV]

1

1.5

0.25

Pd IrRh Pt

E CO [eV]

1.5

Ag

1.5

0.20

1.0

E CO [eV]

2

3

0.15

1.0

Cu

0.5

0.15 0.10 0.05 0.00 0.05 0.10 0.15

E CH3CHO = (0.02 ±0.09) E CO + (0.39 ±0.09) E OH + (-3.36 ± 0.46) 0.10

Rh

0.0

0.02E CO + 0.39E OH [eV]

Ag

0.0

3.6

0.05

E H−CO [eV]

1.5

Rh

3.4

0.0

1.5

3.0

3.2

Ag

Ir

1.5

Ag

Pd

Ag

E CO [eV]

1.0

2.0

3.0

Ir

Cu

0.8

Cu Pt

Pd

0.6

0.0

Pt

2.8

3.8

2.5

2

0.5

2.6

E CH3CHOH = (0.74 ±0.06) E CO + (-2.41 ±0.44)

1.0

Ag

Ag

Cu

0.5

1.5

1

1.0

0.4

1.2 E CH2O = (0.13 ±0.10) E CO + (0.25 ±0.12) E OH + (-0.72 ± 0.23)

E H−CH3CO [eV]

0.4

1.0

E CH3 [eV]

0.0

2.0

CH3

Pd Rh Ir

0.5

CH3CHO

E OH [eV]

2.0

2.0 E CH2OH = (0.77 ±0.09) E CO + (-0.18 ±0.22)

CH3CHOH

E CH3CH2OH = (0.05 ±0.09) E OH + (-4.17 ±0.53) 0.3

1.5

CH3CH2OH

4.4

0.2

1.0

Pt

E CO [eV]

4.2

4.6

E CH2CO = (0.77 ±0.06) E CO + (-1.41 ±0.29)

Pd Rh

E CO [eV]

Rh

4.0

Pt

3.0 3.5

0.5

Pd IrRh

1.0

E CO [eV]

Cu

3.8

2.5

0.5

E H−OH [eV]

3.6

1.0

2.0

E CH3CHO [eV]

0.6

Ag

Cu

Page 10 of 14

0.5

E CH3O [eV]

E H = (0.41 ±0.03) E CO + (0.35 ±0.06)

1.5

Ag

CH2

E CH [eV]

Ir Pt Rh Pd

0.2

0.0

E CH2O [eV]

0.0

0.4

E CH3CH2OH [eV]

E CH2CO [eV]

Cu

E CH3CHOH [eV]

E H [eV]

0.2

Cu

E CH3CO [eV]

1.0

0.4

CH2O

0.2

E CHO [eV]

Ag

CH2OH

E CHOH [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

CH2CO E CH2OH [eV]

H

ACS Catalysis

0.0

Ag

1.5

Cu

1.0 0.5 0.0 0.5 1.0 1.5

Pd Rh Ir Pt E HCO−H = (1.29 ±0.07) E CO + (1.57 ±0.20) 1.5

1.0

E CO [eV]

0.5

0.0

Page 11 of 14 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

ACS Catalysis

ACS Paragon Plus Environment

ACS Catalysis 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19

ACS Paragon Plus Environment

Page 12 of 14

Page 13 of 14 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

ACS Catalysis

ACS Paragon Plus Environment

ACS Catalysis

a)

0.64 0.56

Pt

1.0

E OH

1.5

Ag

Pd Ir

0.5

Cu

0.24 Co

0.08

1.5

1.5

0.90 Pt

1.0

1.0

0.5

0.0

Ag Cu

Rh

0.30 Co

0.15

0.5 2.0 1.5

0.36

0.60

1.5

1.0

0.5

Rate CH4

c)

0.90 Pt

1.0

0.0

e)

0.95

Ag Cu

Rh

0.30 Co

0.15

0.5 2.0 1.5

0.60 0.45

0.0

0.00

Selectivity MeOH > 10%

0.75

Pd Ir

0.5

0.45

0.0

0.00

Selectivity CH3 CHO > 10%

d)

0.75

Pd Ir

0.5

0.16

0.5 2.0

b)

0.32

0.0

1.5

0.48 0.40

Rh

Rate MeOH

1.5

1.0

0.5

0.0

0.00

Selectivity CH4 > 10%

f)

0.90

0.85 Pt

1.0

0.30 Ag

Pd Ir

0.5

Cu

Rh

0.24

0.12 Co

0.06

0.5 1.5

1.0

E CO

0.5

0.0

0.00

0.75 Ag

Pd Ir

0.5

0.35

0.5

0.15 1.5

0.5

1.0

0.5

0.0

E CO

ACS Paragon Plus Environment

0.05

Ag Cu

Rh

0.60 0.45

0.0

0.30

0.25

Co

0.75

Pd Ir

0.45

0.0

2.0

Pt

1.0

0.65 0.55

Cu

Rh

0.18

0.0

2.0

Pt

1.0

Rate Probability

Rate CH3 CHO

Co

0.15

0.5 2.0

1.5

1.0

E CO

0.5

0.0

0.00

Selectivity Probability

1.5

E OH

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

Page 14 of 14