Significance of Surface Formate Coverage on the Reaction Kinetics of

Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China. ACS Catal. , 2017, 7 (10), pp 7187–7195. DOI: 10.1021...
1 downloads 13 Views 2MB Size
Subscriber access provided by UNIVERSITY OF CALGARY

Article

Significance of Surface Formate Coverage on the Reaction Kinetics of Methanol Synthesis from CO2 Hydrogenation over Cu Panpan Wu, and Bo Yang ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b01910 • Publication Date (Web): 15 Sep 2017 Downloaded from http://pubs.acs.org on September 15, 2017

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 free 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 accessible to all readers and 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.

ACS Catalysis 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 31

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

1

Significance of Surface Formate Coverage on the Reaction Kinetics of Methanol

2

Synthesis from CO2 Hydrogenation over Cu

3

Panpan Wu,1,2,3 Bo Yang1,4,*

4 5 6

1

7

Huaxia Road, Shanghai 201210, China

8

2

9

China

School of Physical Science and Technology, ShanghaiTech University, 393 Middle

Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050,

10

3

University of Chinese Academy of Sciences, Beijing, 101407, China

11

4

Key Laboratory of Low-Carbon Conversion Science & Engineering, Shanghai

12

Advanced Research Institute, Chinese Academy of Sciences, Shanghai 201210, China

13

*Email address: [email protected]

14 15 16

Abstract

17

The hydrogenation of CO2 to methanol over copper-based catalysts has attracted

18

considerable attentions recently. Among all the proposed reaction mechanisms, a large

19

number of experimental and theoretical studies have focused on the one via HCOO

20

intermediate due to the fact that a high coverage of formate over catalyst surfaces

21

were observed experimentally. In order to systematically understand the influence of

22

formate species coverage on the reaction kinetics of methanol synthesis, energetics of

23

the CO2 hydrogenation pathway over clean and one/two formate pre-adsorbed Cu(211)

24

are obtained using density functional theory (DFT) calculations, and these energetics

25

are further employed for the microkinetic modeling. We find that the adsorption

26

energies of the intermediates and transition-states involved in the reaction pathway

27

are changed in the presence of spectating formate species, and consequently the

28

potential energy diagrams are varied. Microkinetic analysis shows that the turn-over

29

frequencies (TOFs) over different formate pre-adsorbed surfaces vary under the same 1 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

1

reaction condition. Particularly, the reaction rates obtained over clean Cu(211) are

2

generally the lowest, while those over one/two formate pre-adsorbed surfaces are

3

depending on the reaction temperatures and pressures. Meanwhile, we find that, only

4

when the formate coverage effect is considered, some of the TOFs obtained from

5

microkinetic modeling are in fair agreement with previous experimental results under

6

similar conditions. After the degree of rate control analysis, it is found that the

7

combination of HCOO and HCOOH hydrogenation steps can be treated as the

8

‘effective rate-determining step’, which can be written as HCOO* + 2H* →

9

H2COOH* + 2*. Therefore, the formation of methanol is mainly controlled by the

10

surface coverage of formate and hydrogen at steady state, as well as the free energy

11

barriers of the effective rate-determining step, i.e. effective free energy barriers.

12 13

Key words:

14

Density functional theory; formate coverage effect; CO2 hydrogenation; methanol

15

synthesis; degree of rate control; BEEF-vdW.

16 17

1. Introduction

18

Methanol synthesis from CO2 hydrogenation over copper-based catalysts has attracted

19

considerable interests owing to its industrial applications and environmental

20

significance. In industry, methanol is produced from synthesis gas, a mixture of CO,

21

CO2 and H2, at temperatures of 473~573 K and pressures of 50~100 bar over the

22

Cu/ZnO/Al2O3 catalyst.1,2 Several experimental studies have shown that CO2 in

23

synthesis gas is the main carbon source of methanol synthesis.3-5 For example,

24

Chinchen et al. previously found that, while adding 14C labelled 14CO or 14CO2 to the

25

reactant mixtures of CO2/CO/H2, the fraction of methanol made from carbon dioxide

26

rises with the increase of p C O 2 / p C O ratio, thus the authors concluded that methanol

27

is mainly produced from CO2 in the feedstock.4 The most puzzling questions in this

28

field till now are lying on the reaction mechanism and the nature of catalytic active

29

sites. 2 ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

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

1

Given the fact that the most abundant adsorbate on the catalyst surface within the

2

limit of detection is formate, it is easy to understand that a large number of

3

experimental and theoretical studies focused on the reaction pathway via HCOO

4

intermediate.6-13 For example, Rasmussen et al. developed a model to understand the

5

kinetics of methanol synthesis over Cu and found that formate is a stable

6

intermediate.9 Theoretically, Grabow and Mavrikakis revealed that the hydrogenation

7

of CO2 to methanol on Cu(111) might follow the formate mechanism, in which the

8

surface adsorbed HCOO, HCOOH, CH2OOH, CH2O and CH3O are involved, and it

9

was found that the formation of COOH, the precursor for CO formation, is much

10

more difficult than that of HCOO over Cu.10 Moreover, the hydrogenation of CO2,

11

following the formate mechanism, is found more favorable than CO hydrogenation

12

over Cu.11 Recently, Kattel and co-workers also suggested that methanol synthesis via

13

HCOO intermediate on either Cu-Zn alloy or Cu-ZnO interface is energetically

14

favored compared with the reverse water-gas shift (RWGS) pathway through COOH

15

and CO intermediates.12 Although the debate on whether the formate species is

16

participating the formation pathways of methanol over copper is still going on, its

17

spectating effect is widely recognized.14-16

18

The nature of the active site of copper-based catalysts for methanol synthesis is still

19

under debate. In situ X-ray photoelectron spectroscopy and surface X-ray diffraction

20

experiments suggested that metallic Cu is the active site over copper-based

21

catalysts.17,18 While focusing on pristine Cu catalysts, experiments using Cu(100),9

22

Cu(110),19 and polycrystalline Cu films mainly composed of Cu(111) facet20 indicated

23

that CO2 hydrogenation is structure-sensitive. Moreover, according to the

24

computational results reported by Behrens et al., the adsorption of intermediates

25

involved in CO2 hydrogenation over Cu(211) is much stronger than over Cu(111),21

26

which is consistent with the experimental phenomenon that the defects on the working

27

catalysts may act as the active site.

28

In addition to the analyses on the energetics of methanol formation pathways,

29

microkinetic models were also proposed. van Rensburg and co-workers performed 3 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

1

microkinetic analyses on the relationship between formate saturation coverage and the

2

turn-over rate of methanol synthesis based on the energetic data reported before,11 and

3

they found that the rate of methanol formation is sensitive to the coverage of formate

4

at steady-state.22 The approach the authors used was controlling the sites that formate

5

could occupy during the reaction, and therefore varying the limits of formate

6

saturation coverage, to find out the change of reaction rate. However, it should be

7

pointed out that the effect of formate coverage on the adsorption energies of surface

8

intermediates were not considered in their work. More recently, Studt et al. undertook

9

a microkinetic study on methanol synthesis from syngas over copper-based catalysts

10

combining experimental and theoretical methods, and they concluded that the catalyst

11

mixture optimizes the reaction kinetics, whereas the feed gas optimizes the

12

equilibrium thermodynamics of the system.23

13

In the current work, we present a systematic study regarding the influence of formate

14

coverage on the reaction rate of methanol synthesis with the coverage-dependent

15

adsorbate-adsorbate interactions included. We are using clean and one/two formate

16

pre-adsorbed Cu(211) surfaces to model the reaction pathway at different formate

17

coverages. The same approach has been proved effective in the literature for the

18

analysis of coverage effects in surface reaction studies.24,25 Such approach is also able

19

to reflect different reaction kinetics along with the process of formate build-up over

20

the catalyst at the initial stage of CO2 hydrogenation. An alternative approach is to

21

determine the relations between all the adsorption/activation energies and coverages,

22

which may be linear or non-linear, and then run the self-consistent iterative

23

simulations.26,27 However, this approach is quite time-consuming considering the

24

complex reaction network of methanol synthesis, and the results obtained are strongly

25

dependent on the relations obtained, and therefore is not taken here. The adsorption

26

energies of HCOO at different coverages are calculated and compared. Meanwhile,

27

the energetics of all elementary reactions involved in the mechanism are calculated at

28

incremental formate coverages. We then carry out a microkinetic modeling based on

29

the energies obtained from density functional theory (DFT) calculations to gain more 4 ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

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

1

insight into the effect of formate coverage on the reaction rate and the identification of

2

rate-determining steps in the whole pathway.

3 4

2. Computational details

5

2.1 DFT calculation methods

6

The

7

projector-augmented wave (PAW) method31,32 was used to perform all the density

8

functional calculations. The generalized gradient approximation (GGA) with the

9

BEEF-vdW exchange-correlation functional,33 which explicitly takes long-range

10

dispersion force into consideration and has been widely used to model surface

11

catalysis reactions,11,34-36 was employed to set the plane-wave basis. The BEEF-vdW

12

functional was used here because it was found that the inclusion of van der Waals

13

interactions gives a better description on the energetics of CO2 hydrogenation

14

compared to RPBE.11 The bulk lattice constant of copper was optimized yielding a

15

value of 3.658 Å, which is close to the experimental bulk lattice constant of 3.615 Å.

16

Twelve-layer and 3×4 slabs with the upmost six Cu layers relaxed during optimization

17

were used to model the adsorption and reaction processes over Cu(211) surfaces.

18

Therefore, one formate species on the surface would give rise to a coverage of 0.083

19

(1/12) monolayer (ML). The slab was set with a vacuum to be at least 11 Å to make

20

sure the reactions only take place on one side of the slabs. A 4×2×1 k-point grid

21

generated with the Monkhorst-Pack scheme was used. An energy cutoff of 500 eV and

22

convergence criteria of the force on each relaxed atoms below 0.05 eV/ Å were found

23

to give converged results in the current work. Transition-states were located with a

24

constrained minimization method.37-39 In this method, transition-states are identified

25

when (i) the force on the relaxed atoms vanish and (ii) the energy is a maximum along

26

the reaction coordinate, but a minimum with respect to all of the remaining degrees of

27

freedom.

28

Adsorption energies were defined as:

29

Vienna

Ab-initio

Simulation

Package

(VASP)

Ead = Etotal − Eslab − Eg 5 ACS Paragon Plus Environment

code28-30

with

the

(1)

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 6 of 31

1

where Etotal is the energy of the system after adsorption, Eg is the energy of the

2

gas-phase molecule, and Eslab is the energy of slab. The correction to the energies of

3

gas phase molecules resulted from the systematic DFT errors in describing the

4

carbon-oxygen double bond with BEEF-vdW was used according to the values

5

reported in the literature,11,23,40,41 which are +0.41 and +0.09 eV for the gaseous CO2

6

and H2, respectively. In order to compare the adsorption of formate at different

7

coverages, the average and differential adsorption energies of formate were calculated,

8

where the average adsorption energies (∆Eavg) were calculated from the equation,

∆Eavg ( N ) =

9

Eslab+ N ⋅ads. − Eslab − N ⋅ Eads N

10

and the differential adsorption energies (∆Ediff) were calculated from the equation,

11

∆Ediff ( N ) = Eslab+N ⋅ads − Eslab+(N -1)⋅ads − Eads

(2)

(3)

12

where Eslab+N·ads is the energy of the system containing adsorbates with the number of

13

N, Eads is the energy of formate with respect to the energy of gaseous CO2+1/2H2.

14 15

2.2 Microkinetic model

16

Microkinetic analysis in the current work is manipulated using CatMAP module

17

introduced by Nørskov’s group,42,43 which has been used widely to obtain more

18

insight into surface catalytic reactions. In this module, elementary reactions

19

considered are the same with those reported in the literature10-12,21-23 and are listed as

20

follows:

21

H2(g) + 2* → 2H*

(R1)

22

CO2(g) + H* → HCOO*

(R2)

23

HCOO* + H* → HCOOH* + *

(R3)

24

HCOOH* + H* → CH2OOH* + *

(R4)

25

CH2OOH* + *→ CH2O* + OH*

(R5)

26

CH2O* + H* → CH3O* + *

(R6)

27

CH3O* + H*→ CH3OH(g) + 2*

(R7)

28

OH* + H* → H2O(g) + 2*

(R8) 6

ACS Paragon Plus Environment

Page 7 of 31

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

1

It should be mentioned that we also compared the activation energies of COOH

2

formation and CO2 direct decomposition at different formate coverages over Cu(211).

3

The corresponding transition-state configurations and activation energies are

4

presented in Figure S1 in the SI. We find that these two possible reactions are

5

experiencing much higher barriers than formate formation (as listed in Table 1).

6

Here, two different sites were considered: one for hydrogen atoms and the other for

7

the remaining adsorbates. In this approach, hydrogen is adsorbed at a special

8

“hydrogen reservoir” site,27,44 and does not compete with other adsorbates for a free

9

site. The free energies of adsorbates and transition-states at temperature T were

10

estimated according to the harmonic approximation, and the entropy is evaluated

11

using the following equation:

12

S (T ) = kB

harm DOF

∑ i

 εi − ln 1 − e −ε i  ε i kBT k T ( e − 1)  B

(

kBT



) , 

13

where kB is Boltzmann constant; DOF is the number of harmonic energies (εi) used in

14

the summation denoted as degree of freedom, which is generally 3N, where N is the

15

number of atoms in the adsorbates or transition-states.

16

Meanwhile, the free energies of gas phase species are corrected as

17

G g (T ) = Eelec + E ZPE + ∫ C p d T − TS (T ) ,

18

where Cp is gas phase heat capacity as a function of temperature derived from

19

Shomate equations and the corresponding parameters in the equations were obtained

20

from

21

transition-state-theory. The temperature considered is from 373 to 573 K, consistent

22

with the typical industrial reaction conditions. The concentration of feed gases is CO2:

23

H2: CH3OH: H2O: inert gas = 0.10: 0.40: 0.01: 0.01: 0.48, which corresponds to a

24

CO2 conversion of around 10% and therefore the adsorption/desorption of methanol

25

and water are not considered in our kinetic model. The total pressure (ptotal) ranges

26

from 10 to 100 bar, corresponding to the reactant (CO2 + H2) partial pressure (preactant)

27

varying from 5 to 50 bar.

NIST.45

The

pre-exponential

factors

were

28

7 ACS Paragon Plus Environment

determined

using

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

1

3. Results and Discussion

2

3.1 Formate adsorption energies as a function of coverage

3

As mentioned above, it was reported in the literature that the reaction rate of methanol

4

synthesis is sensitive to HCOO saturation coverage.22 Meanwhile, the coverage of

5

adsorbates is relevant to the corresponding adsorption energies, i.e. stronger

6

adsorption of formate would increase its coverage, but higher coverage would in turn

7

weaken the adsorption strength to some extent due to the lateral interactions between

8

adsorbates.

9

different coverages are calculated and the trend is shown in Figure 1. One can see that

10

the adsorption of formate becomes weaker with the increase of coverage, and the

11

average (differential) adsorption energy varies from -1.34 (-1.34) eV to -1.06 (-0.78)

12

eV corresponding to the formate coverage from 0.083 to 0.250 ML, respectively.

13

Coverage-dependent adsorbate-adsorbate interactions have been studied previously

14

for the systems of CO methanation and NO oxidation,27,52 and it was found that the

15

reaction rates might change by several orders of magnitude with coverage variation.

16

Considering that formate is the most abundant species on the surface, it is of essence

17

to take the interactions between formate and the surface intermediates into

18

consideration while investigating the reaction kinetics of methanol synthesis.

19

In the reaction pathways investigated, it is postulated that only one formate species in

20

the slab participates in methanol synthesis reaction and the residual adsorbed HCOO

21

act as spectators during the process at different coverages. Therefore, we define θDFT

22

as the coverage of all the surface species while performing the DFT calculations, and

23

pre-ads , where θIM is the coverage of surface intermediates during the θ DFT = θ IM + θ HCOO

24

pre-ads reaction, which is 0.083 ML, and θ HCOO is the coverage of formate pre-adsorbed on

25

Cu(211), which is 0, 0.083 and 0.167 ML for the surfaces with zero, one and two

26

pre-adsorbed formate species, respectively.

2,46-51

Here, the average/differential adsorption energies of formate at

27 28

3.2 CO2 hydrogenation over clean Cu(211)

29

The reaction energetics over clean Cu(211), where θDFT is 0.083 ML and the 8 ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

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

1

adsorbate-adsorbate interactions can be neglected, are calculated and listed in Table 1.

2

The configurations of corresponding adsorption and transition states are presented in

3

Figure 2. It should be mentioned that we tried different possible adsorption and

4

transition state structures in the current work and those shown in the figures are with

5

the lowest energies. One can see that the transition-state structures of HCOO-H,

6

H-CHOOH, CH3O-H and H-OH are those with hydrogen atoms binding to the Cu

7

atoms at the lower terrace of Cu(211), in accordance with the structures reported

8

previously in the literature.11 As listed in Table 1, the hydrogenation of HCOO has the

9

highest activation energy of 1.27 eV in the reaction pathway, and the corresponding

10

reaction energy is calculated to be endothermic by 0.55 eV, indicating that the

11

hydrogenation of HCOO is relatively difficult and formate species may get

12

accumulated over clean Cu(211).

13 14

3.3 CO2 hydrogenation over one and two formate pre-adsorbed Cu(211)

15

We study the coverage effect by employing a model with one and two fomate species

16

pre-adsorbed on the catalyst surfaces with a bridging bi-dentate structure, leading to

17

θDFT values of 0.167 and 0.250 ML, respectively. The relevant adsorption and

18

transition state structures are presented in Figure 3.

19

When θDFT = 0.167 ML, corresponding to the situation of one formate and one

20

reaction intermediate on the surface, the spectating formate species adsorbs on the

21

surface through two oxygen atoms binding with two copper atoms at the step-edge

22

site. The adsorption configurations of the intermediates involved in the reactions are

23

quite similar to those over clean Cu(211), whereas the geometries of transition-states

24

are slightly different for CH2O-OH and H-CH2O. As shown in Figure 3, at the θDFT of

25

0.167 ML, both CH2OOH dissociation and CH2O hydrogenation occurs with CH2O

26

binding at the step-edge site whilst the hydroxyl/hydrogen moiety staying at the

27

terrace Cu atoms.

28

Further increase of the number of pre-adsorbed formate changes θDFT to 0.250 ML

29

over Cu(211). Under such situation, the two spectating formate species tend to get 9 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

1

stabilized through adjusting their adsorption configurations. As we can see from

2

Figure 3, one of the formate species binds at the step-bridge site with two oxygen

3

atoms, while the other formate turns into the structure with two oxygen atoms

4

bonding to the step and terrace copper atoms, respectively. It should be noted that we

5

have examined the barriers of the formation and hydrogenation of all three formate

6

species, the structures of which are shown in Figure S2 in the Supporting Information

7

(SI), and those shown in Figure 3 are the most stable ones.

8

One can see that most of the configurations of surface intermediates at the θDFT of

9

0.250 ML are similar to those at 0.167 ML, and therefore only the different ones are

10

discussed here. As shown in Figure 3, CH2OOH prefers to stay at the terrace hcp

11

hollow site when θDFT = 0.250 ML, whilst the bridge site at step edge is favored at

12

lower θDFT. Upon the dissociation of CH2OOH, the produced CH2O still binds to one

13

step copper atom but with the C=O bond more vertical to the step-edge site.

14

Meanwhile, the hydroxyl moiety tends to stay at the terrace hcp hollow site when θDFT

15

is higher. These subtle structure changes imply that the self-adjustment of

16

intermediates are experiencing a steric effect caused by the increase of formate

17

coverage. The corresponding transition-state structures are presented in Figure 3,

18

which are similar to those at θDFT = 0.167 ML.

19 20

3.4 Potential energy diagrams at different θDFT

21

Based on the energies obtained above, the potential energy diagrams of CO2

22

hydrogenation to methanol at different θDFT are presented in Figure 4. It can be found

23

from the figure that the activation of CO2 is insensitive to surface formate coverage

24

variation, whilst the barriers of formate further hydrogenation to formic acid are

25

decreasing with the increase of formate coverage.

26

It is interesting to see that the adsorption of formic acid at the θDFT of 0.167 ML is the

27

strongest among the three studied coverages, and the trend is different from that

28

observed for the adsorption of formate. To explain this, charge density differences

29

induced by the adsorption of HCOOH is studied. The charge density difference (∆ρ) 10 ACS Paragon Plus Environment

Page 10 of 31

Page 11 of 31

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

1

is defined as

∆ ρ = ρ HCOOH+slab − ρ slab − ρ HCOOH

2

ρHCOOH+slab

(4)

3

where

is the charge density distribution of formic acid adsorbed on

4

slabs with different fomate species pre-adsorbed as shown in Figure 5, and

5

ρHCOOH

6

pre-adsorbed formate species and HCOOH only, respectively. The yellow contours

7

represent charge accumulations, and the blue contours denote charge depressions. The

8

scales of the charge densities are identical. It can be clearly seen that the charge

9

distribution of the copper atom that HCOOH adsorbs on differs from each other at

10

different formate coverages. In Figure 5b, the copper atom binding with HCOOH

11

transfers the most charge, corresponding to the θDFT of 0.167 ML. This is consistent

12

with the trend obtained for the adsorption of HCOOH and indicating that the presence

13

of spectating formate would influence the charge distribution of copper atoms.

14

In contrast to the formation of HCOOH, both the barriers and reaction energies for

15

CH2OOH generation and dissociation at different coverages are similar, therefore,

16

stronger HCOOH adsorption would give rise to more stable transition-states for

17

formic acid hydrogenation, as shown in Figure 4. It will be revealed later that the

18

transition-state of CH2OOH generation is in fact rate-determining in the reaction

19

pathway of methanol synthesis under the industrial reaction conditions, and the trend

20

of HCOOH adsorption energies revealed above is crucial to understanding the activity

21

trend of methanol synthesis. Subsequent hydrogenation of CH2O and OH lead to the

22

desired products CH3OH and H2O, respectively, and the corresponding barriers and

23

reaction energies are listed in Table 1. Based on the analyses of the energy diagrams

24

in Figure 4, it is obvious that the energetics of CO2 hydrogenation are indeed strongly

25

influenced by the number of pre-adsorbed formate.

ρslab

and

are the charge density distributions of Cu (211) with different numbers of

26 27

3.5 Reaction kinetics at different θDFT

28

The kinetics of methanol synthesis are explored based on the energetics obtained 11 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

1

above at different reaction conditions. In our study, the coverage of most

2

intermediates obtained are essentially zero at the steady state and only surface HCOO

3

are found. Figure 6a-c shows the trend of the obtained formate coverages from kinetic

4

k , ss analyses at steady state ( θ HCOO ), in which the coverage of pre-adsorbed formate is

5

included, and the corresponding calculated turn-over frequency (TOF) as a function of

6

temperature (373 - 573 K) and pressure (preactant = 5, 25 and 50 bar) are presented in

7

Figure 6d-f. The trend of the obtained TOFs at all the temperatures and pressures

8

studied are shown in Figure S3 in the SI. It is clear that the coverage of formate and

9

the TOF of methanol synthesis obtained from our microkinetic modeling at steady

10

state is strongly affected by the variation of θDFT.

11

k , ss When θDFT = 0.083 ML, as we can see from Figure 6a, θ HCOO behaves a minor

12

decrease from 1 to ~0.8 ML with the increase of temperature at low reactant pressure

13

(preactant = 5 bar). However, it remains almost 1 ML at higher pressures (preactant = 25

14

k , ss and 50 bar) as shown in Figure 6b and c. All of the θ HCOO s obtained are much higher

15

than the θDFT used, indicating that the energetics obtained from DFT calculations at

16

this low coverage cannot give rise to converged steady state information.

17

k , ss As also shown in Figure 6a, when θDFT = 0.167 ML, θ HCOO decreases with the

18

increase of temperature, and converges around 0.167±0.050 ML from 513 to 543 K at

19

preactant = 5 bar and at 573 K and preactant = 25 bar, whilst at preactant = 50 bar, no

20

k , ss convergence between θDFT and θ HCOO is found in the temperature range studied in

21

the current work. However, one can see from Figure 6b and c that it is possible for

22

k , ss to approach 0.167 ML when the temperature is further increased. In the case of θ HCOO

23

k , ss θDFT = 0.250 ML, the coverages obtained ( θ HCOO ) converges at several different

24

conditions, e.g. 373 K with the value of 0.220 ML at preactant = 25 bar.

25

The TOFs of methanol synthesis are found increasing with temperature at all the

26

reactant pressures and coverages of surface species studied. We list the reaction

27

k , ss conditions, where the θ HCOO obtained converge with θDFT, and the corresponding

12 ACS Paragon Plus Environment

Page 12 of 31

Page 13 of 31

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

1

TOFs in Table S1 in the SI. Considering that the formation rate of methanol obtained

2

over clean Cu(211) surface is 1.11 × 10-6 site-1 s-1 at the temperature of 513K and

3

reactant pressure of 5 bar, it is found from this table that the reaction rate obtained at

4

θDFT = 0.167 ML (4.17 × 10-4 site-1 s-1) is in fair agreement with that measured from

5

experiments, i.e. 1.2 × 10-3 site-1 s-1 under the same reaction conditions,53 whilst the

6

TOF obtained based on θDFT =0.083 ML energetics is not.

7 8

3.6 Degree of rate control analyses

9

In terms of the kinetic analyses of multistep reactions, identifying the

10

rate-determining steps (RDSs) is pivotal to making good predictions of the reaction

11

rates and to improving the TOFs by optimizing the structure of catalysts or reaction

12

conditions in practical terms.54 According to the relevant investigations reported, the

13

tool named degree of rate control (DRC) facilitates the analyses of reaction

14

mechanisms and kinetics, serving a quite similar purpose to determining the RDS.54-56

15

In this approach, Xi is defined as the value of DRC for the adsorption or transition

16

state i, and can be written as:

17

 −∂ ln r   Xi =   ∂ ( Gio / RT )  o  G j≠i

(5)

18

where Gio denotes the free energies of surface adsorbates or transition-states and the

19

partial derivative is taken holding constant the standard-state free energy of all other

20

species j (intermediates, transition-states, reactants and products). Specifically, the Xi

21

can be defined as XRC,i (degree of rate control for transition-states) or XTRC,i (degree of

22

thermodynamic rate control for adsorbates). We here investigate the Xi of the

23

methanol synthesis reactions at preactant = 50 bar to gain insight into the variation of

24

rate-determining step with temperature at different θDFT.

25

It is obvious from Figure 7 that XRC,i is positive and, in opposite, that XTRC,i is negative,

26

and the sum of XRC,i is 1 all the time.54-56 Taking the case of θDFT = 0.083 ML as an

27

example, as shown in Figure 7a, XRC,H-CHOOH increases from 0 to 1 with the 13 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

1

temperature varying from 373 to 573K, while XRC,CH2O-OH decreases from 1 to 0

2

within the same temperature range. However, XRC,HCOO-H remains zero all the time,

3

indicating that the formation of formic acid is not rate-limiting under the conditions

4

studied. We note that the higher DRC of HCOOH hydrogenation than HCOO

5

hydrogenation is consistent with the computational results reported before.22

6

It was also reported in the literature that the degree of thermodynamic rate control of

7

the surface intermediate n is proportional to the coverage of such an intermediate (θn),

8

i.e. X TRC,n = −σ ⋅ θ n , where σ is the number of surface free site required by the

9

rate-determining elementary reactions identified.54-56 As we can see from Figure 7a,

10

the rate-determining transition state at θDFT = 0.083 ML is mainly the dissociation of

11

CH2OOH below 473K, which requires one more surface free sites to occur.

12

k , ss Meanwhile, the corresponding θ HCOO at low temperatures is almost 1 ML (see Figure

13

6c). However, as we can see from Figure S4 in the SI, with the increase of

14

temperature, the coverage of HCOO slightly decreases, and small amounts of free

15

sites released facilitate the dissociation of CH2OOH, resulting in the decrease of

16

XRC,CH2O-OH and increase of XRC,H-CHOOH. When the temperature is higher than 473 K,

17

the hydrogenation of HCOOH becomes the rate-determining transition-state.

18

Similar investigations are carried out at θDFT = 0.167 and 0.250 ML as well, the results

19

of which are shown in Figure 7b and c, respectively. It is clear that, at θDFT = 0.167

20

ML, the reaction rate is mainly sensitive to the free energy of H-CHOOH

21

transition-state above 423 K. With the increase of θDFT, the transition-state

22

determining the reaction rate of methanol synthesis is always H-CHOOH within the

23

whole temperature range.

24

As discussed above, we find that it is reasonable to consider that the adsorption-state

25

of HCOO and the transition-state of HCOOH hydrogenation as the rate-determining

26

initial state and transition state of methanol synthesis, respectively, under the typical

27

industrial reaction conditions (> 500 K and preactant = 50 bar). Therefore, the

28

combination of HCOO and HCOOH hydrogenation steps can be treated as the

29

‘effective rate-determining step’, which is written as HCOO* + 2H* → H2COOH* + 14 ACS Paragon Plus Environment

Page 14 of 31

Page 15 of 31

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

1

2*, for this process. The overall reaction rate of CO2 hydrogenation to methanol can

2

therefore be estimated using eff

3

r = k ⋅θHCOO ⋅θH2 = A ⋅ e−Ga

/ k BT

θHCOO ⋅θH2 ,

(6)

4

where Gaeff is the effective free energy barrier of the effective rate-determining step

5

and can be calculated with Ga , R 4 + ∆GR 3 , with Ga , R 4 and ∆GR 3 being the free

6

energy barrier of R4 and reaction free energy of R3, respectively. For example, at 503

7

K and preactant = 50 bar, the effective free energy barrier is 1.14 eV and the coverage of

8

HCOO and H over one formate pre-adsorbed Cu(211) surface is 0.796 and 0.026 ML,

9

respectively. Thus, the TOF estimated using equation (6) is 2.17 × 10-2 site-1 s-1, which

10

is identical to the rate obtained from microkinetic modeling at steady-state, as one can

11

see from Table S2 in the SI. The TOFs of methanol synthesis obtained from CatMAP

12

and those computed from equation (6) proposed in our work are shown in Figure 8 as

13

a function of temperature, and all the data used to plot this figure are listed in Table

14

S2 of the SI. We find that the TOFs fit well from the temperature where the XH-CHOOH

15

is approaching 1 and the transition state of HCOOH hydrogenation is rate determining

16

at different formate coverages, which strongly support the effective rate determining

17

step found in our work.

18

We further find that the hydrogenation of CO2 can be understood using the two-step

19

model developed before,2,49-51,57-59 in which the formation of methanol can be divided

20

into two processes, namely the ‘adsorption’ and ‘desorption’ of HCOO to/from the

21

catalyst surface. The ‘adsorption’ of HCOO is in fact the hydrogen-assisted adsorption

22

of CO2 onto the catalyst surface, whilst the ‘desorption’ process of HCOO from the

23

surface is a combination of a series of elementary steps including hydrogenation and

24

dissociation steps heading to the final formation of methanol. It is obvious from our

25

work that, over Cu(211), the ‘desorption’ of HCOO is rate-determining under reaction

26

conditions. However, the transition-state of HCOO hydrogenation is not necessarily

27

rate-determining, since the energy of this transition-state is lower than that of

28

HCOOH hydrogenation from the free energy landscape as shown in Figure S5 (T = 15 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

1

503 K, pCO2 = 10 bar, pH2 = 40 bar and pCH3OH = pH2O = 1 bar) in the SI. While

2

increasing the values of θDFT, the corresponding adsorption strength of HCOO is

3

weakened to a larger extent than the destabilization of H-CHOOH on the surface,

4

leading to lower surface formate coverage at steady-state and smaller effective free

5

energy barriers. Considering that the coverage of hydrogen at the same temperature

6

and pressure over the surface are identical at different θDFT, the formation of methanol

7

is mainly controlled by the surface coverage of formate and the effective free energy

8

barriers as shown in equation (6) accordingly.

9 10

4. Conclusions

11

In summary, a systematic analysis of the effect of formate coverage on the reaction

12

kinetics of methanol synthesis over Cu(211) is carried out in the current work. The

13

following conclusions are obtained:

14

(i) Increasing formate coverage would destabilize the adsorption of most of the

15

reaction intermediates and transition states, except for the adsorption of formic acid at

16

θDFT = 0.167 ML. Further charge density difference analysis reveals that the copper

17

atom binding with HCOOH transfers the most charge under this situation;

18

(ii) TOFs at different coverages of surface species are also changed under the same

19

reaction condition, particularly, the reaction rates at θDFT = 0.083 ML are the lowest in

20

general. Meanwhile, we find that taking the coverage effect into account while

21

carrying out DFT calculations would give rise to more reasonable steady-state

22

information, compared with those from previous experimental studies.

23

(iii) The combination of HCOO and HCOOH hydrogenation steps can be treated as

24

the ‘effective rate-determining step’, which can be written as HCOO* + 2H* →

25

H2COOH* + 2* over Cu(211). As a result, the formation of methanol is mainly

26

controlled by the surface coverage of formate and the effective free energy barriers,

27

i.e. r = A ⋅ e

28

Our work shows that, in order to obtain better understandings on the kinetics of

29

surface reactions and to design more effective catalysts, it is essential to consider the

−Gaeff / kBT

θHCOO ⋅θH2 .

16 ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31

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

1

coverage effect of surface abundant species while performing DFT calculations and

2

microkinetic studies.

3 4

Supporting Information

5

Turn-over frequencies, coverages and free energy barriers at different temperatures,

6

adsorption and transition state structures, free energy profiles

7 8

Acknowledgements

9

This work is financially supported by ShanghaiTech University, Shanghai Pujiang

10

Program (16PJ1406800), Shanghai Young Eastern Scholar Program (QD2016049)

11

and the Foundation of Key Laboratory of Low-Carbon Conversion Science &

12

Engineering, Shanghai Advanced Research Institute, Chinese Academy of Sciences.

13

We thank Shanghai Supercomputer Center for computing time.

14 15

References

16

(1) Lee, S. Methanol synthesis technology; CRC Press, 1989.

17

(2) Wu, P.; Yang, B. Phys. Chem. Chem. Phys. 2016, 18, 21720-21729.

18

(3) Liu, G.; Willcox, D.; Garland, M.; Kung, H. H. J. Catal. 1984, 90, 139-146.

19

(4) Chinchen, G. C.; Denny, P. J.; Parker, D. G.; Spencer, M. S.; Whan, D. A.

20 21 22 23 24 25 26 27 28 29

Appl. Catal. 1987, 30, 333-338. (5) Muhler, M.; Nielsen, L. P.; Törnqvist, E.; Clausen, B. S.; Topsøe, H. Catal. Lett. 1992, 14, 241-249. (6) Nakano, H.; Nakamura, I.; Fujitani, T.; Nakamura, J. J. Phy. Chem. B 2001, 105, 1355-1365. (7) Nakamura, I.; Nakano, H.; Fujitani, T.; Uchijima, T.; Nakamura, J. J. Vac. Sci. Technol. A, 1999, 17, 1592-1595. (8) Nakamura, J.; Kushida, Y.; Choi, Y.; Uchijima, T.; Fujitani, T. J. Vac. Sci. Technol. A, 1997, 15, 1568-1571. (9) Rasmussen, P. B.; Holmblad, P. M.; Askgaard, T.; Ovesen, C. V.; Stoltze, P.; 17 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

1

Nørskov, J. K.; Chorkendorff, I. Catal. Lett. 1994, 26, 373-381.

2

(10) Grabow, L. C.; Mavrikakis, M. ACS Catal. 2011, 1, 365-384.

3

(11) Studt, F.; Abild-Pedersen, F.; Varley, J. B.; Nørskov, J. K. Catal. Lett. 2012,

4 5 6 7 8 9 10 11 12 13 14 15 16

143, 71-73. (12) Kattel, S.; Ramírez, P. J.; Chen, J. G.; Rodriguez, J. A.; Liu, P. Science 2017, 355, 1296-1299. (13) Zhao, Y.-F.; Yang, Y.; Mims, C.; Peden, C. H. F.; Li, J.; Mei, D. J. Catal. 2011, 281, 199-211. (14) Yang, Y.; Mims, C. A.; Disselkamp, R. S.; Kwak, J.-H.; Peden, C. H. F.; Campbell, C. T. J. Phy. Chem. C 2010, 114, 17205-17211. (15) Yang, Y.; Mims, C. A.; Mei, D. H.; Peden, C. H. F.; Campbell, C. T. J. Catal. 2013, 298, 10-17. (16) Yang, Y.; Mei, D.; Peden, C. H. F.; Campbell, C. T.; Mims, C. A. ACS Catal. 2015, 5, 7328-7337. (17) Clausen, B. S.; Steffensen, G.; Fabius, B.; Villadsen, J.; Feidenhans'l, R.; Topsøe, H. J. Catal. 1991, 132, 524-535.

17

(18) Fleisch, T. H.; Mieville, R. L. J. Catal. 1984, 90, 165-172.

18

(19) Yoshihara, J.; Campbell, C. T. J. Catal. 1996, 161, 776-782.

19

(20) Fujitani, T.; Nakamura, I.; Watanabe, T.; Uchijima, T.; Nakamura, J. Catal.

20

Lett. 1995, 35, 297-302.

21

(21) Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen,

22

F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B.-L.; Tovar, M.; Fischer, R. W.; Nørskov,

23

J. K.; Schlögl, R. Science 2012, 336, 893-897.

24 25

(22) van Rensburg, W. J.; Petersen, M. A.; Datt, M. S.; van den Berg, J. A.; van Helden, P. Catal. Lett. 2015, 145, 559-568.

26

(23) Studt, F.; Behrens, M.; Kunkes, E. L.; Thomas, N.; Zander, S.; Tarasov, A.;

27

Schumann, J.; Frei, E.; Varley, J. B.; Abild-Pedersen, F.; Nørskov, J. K.; Schlögl, R.

28

ChemCatChem 2015, 7, 1105-1111.

29

(24)Calaza, F.; Stacchiola, D.; Neurock, M.; Tysoe, W. T. J. Am. Chem. Soc. 2010, 18 ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31

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

1 2 3

132, 2202-2207. (25) Smeltz, A. D.; Getman, R. B.; Schneider, W. F.; Ribeiro, F. H. Catal. Today 2008, 136, 84-92.

4

(26) Grabow, L. C.; Hvolbæk, B.; Nørskov, J. K. Top. Catal. 2010, 53, 298-310.

5

(27) Lausche, A. C.; Medford, A. J.; Khan, T. S.; Xu, Y.; Bligaard, T.;

6

Abild-Pedersen, F.; Nørskov, J. K.; Studt, F. J. Catal. 2013, 307, 275-282.

7

(28) Kresse, G.; Furthmuller, J. Comput. Mater. Sci. 1996, 6, 15-50.

8

(29) Kresse, G.; Furthmuller, J. Phys. Rev. B 1996, 54, 11169-11186.

9

(30) Kresse, G.; Hafner, J. Phy. Rev. B 1994, 49, 14251-14269.

10

(31) Blochl, P. E. Phys. Rev. B 1994, 50, 17953-17979.

11

(32) Kresse, G.; Joubert, D. Phys. Rev. B 1999, 59, 1758-1775.

12

(33) Wellendorff, J.; Lundgaard, K. T.; Møgelhøj, A.; Petzold, V.; Landis, D. D.;

13 14 15 16 17 18 19 20 21

Nørskov, J. K.; Bligaard, T.; Jacobsen, K. W. Phy. Rev. B 2012, 85, 235149. (34) Latimer, A. A.; Kulkarni, A. R.; Aljama, H.; Montoya, J. H.; Yoo, J. S.; Tsai, C.; Abild-Pedersen, F.; Studt, F.; Norskov, J. K. Nature Mater. 2017, 16, 225-229. (35) Medford, A. J.; Wellendorff, J.; Vojvodic, A.; Studt, F.; Abild-Pedersen, F.; Jacobsen, K. W.; Bligaard, T.; Norskov, J. K. Science 2014, 345, 197-200. (36) Ulissi, Z. W.; Medford, A. J.; Bligaard, T.; Norskov, J. K. Nat. Commun. 2017, 8. (37) Alavi, A.; Hu, P.; Deutsch, T.; Silvestrelli, P. L.; Hutter, J. Phys. Rev. Lett. 1998, 80, 3650-3653.

22

(38) Liu, Z.-P.; Hu, P. J. Am. Chem. Soc. 2003, 125, 1958-1967.

23

(39) Michaelides, A.; Liu, Z. P.; Zhang, C. J.; Alavi, A.; King, D. A.; Hu, P. J. Am.

24 25 26 27 28 29

Chem. Soc. 2003, 125, 3704-3705. (40) Christensen, R.; Hansen, H. A.; Vegge, T. Catal. Sci. Technol. 2015, 5, 4946-4949. (41) Yoo, J. S.; Christensen, R.; Vegge, T.; Nørskov, J. K.; Studt, F. ChemSusChem 2016, 9, 358-363. (42) Medford, A. J.; Shi, C.; Hoffmann, M. J.; Lausche, A. C.; Fitzgibbon, S. R.; 19 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

1 2 3 4 5

Bligaard, T.; Nørskov, J. K. Catal. Lett. 2015, 145, 794-807. (43)Yang, N.; Medford, A. J.; Liu, X.; Studt, F.; Bligaard, T.; Bent, S. F.; Nørskov, J. K. J. Am. Chem. Soc. 2016, 138, 3705-3714. (44) Garcia-Pintos, D.; Voss, J.; Jensen, A. D.; Studt, F. J. Phy. Chem. C 2016, 120, 18529-18537.

6

(45) NIST website: http://physics.nist.gov/cuu/Constants/index.html.

7

(46) Yang, B.; Burch, R.; Hardacre, C.; Headdock, G.; Hu, P. ACS Catal. 2012, 2,

8 9 10 11 12 13 14 15 16 17 18 19 20 21 22

1027-1032. (47) Yang, B.; Burch, R.; Hardacre, C.; Hu, P.; Hughes, P. J. Phys. Chem. C 2014, 118, 3664-3671. (48) Yang, B.; Burch, R.; Hardacre, C.; Hu, P.; Hughes, P. J. Phys. Chem. C 2014, 118, 1560-1567. (49) Yang, B.; Burch, R.; Hardacre, C.; Headdock, G.; Hu, P. J. Catal. 2013, 305, 264-276. (50) Yang, B.; Burch, R.; Hardacre, C.; Hu, P.; Hughes, P. Surf. Sci. 2016, 646, 45-49. (51) Yang, B.; Gong, X.-Q.; Wang, H.-F.; Cao, X.-M.; Rooney, J. J.; Hu, P. J. Am. Chem. Soc. 2013, 135, 15244-15250. (52) Wu, C.; Schmidt, D. J.; Wolverton, C.; Schneider, W. F. J. Catal. 2012, 286, 88-94. (53)Yoshihara, J.; Parker, S. C.; Schafer, A.; Campbell, C. T. Catal. Lett. 1995, 31, 313-324.

23

(54) Campbell, C. T. ACS Catal. 2017, 2770-2779.

24

(55)Stegelmann, C.; Andreasen, A.; Campbell, C. T. J. Am. Chem. Soc. 2009, 131,

25 26 27 28 29

8077-8082. (56) Wolcott, C. A.; Medford, A. J.; Studt, F.; Campbell, C. T. J. Catal. 2015, 330, 197-207. (57) Yang, B.; Burch, R.; Hardacre, C.; Headdock, G.; Hu, P. ACS Catal. 2014, 4, 182-186. 20 ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31

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

1

(58) Yang, B.; Burch, R.; Hardacre, C.; Hu, P.; Hughes, P. Catal. Sci. Technol.

2

2017, 7, 1508-1514.

3

(59) Wang,

4

K.;

Yang,

B.

Catal.

Sci.

Technol.

10.1039/C1037CY01500C.

5 6 7

21 ACS Paragon Plus Environment

2017,

DOI:

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 22 of 31

1

Table 1. Activation energies (Ea) and reaction energies (∆E) (in eV) of the elementary

2

reactions of methanol synthesis over Cu(211) at different covarges of surface spescies.

3

ZPE corrections are not included. 0.083 ML

0.167 ML

0.250 ML

Ea

∆E

Ea

∆E

Ea

∆E

CO2(g)+H*→HCOO*

0.20

-1.27

0.23

-0.99

0.29

-0.71

HCOO*+H*→HCOOH*+*

1.27

0.55

1.02

0.17

0.88

0.04

HCOOH*+H*→CH2OOH*+*

0.76

-0.11

0.79

0.03

0.82

0.05

CH2OOH*+*→CH2O*+OH*

0.42

0.35

0.49

0.34

0.30

0.31

CH2O*+H*→CH3O*+*

0.11

-1.09

0.21

-0.91

0.61

-0.53

CH3O*+H*→CH3OH(g)+2*

0.89

0.37

0.72

0.29

0.59

0.02

H*+OH*→H2O(g)+2*

1.01

0.32

0.87

0.20

0.83

-0.05

elementary reactions

4

22 ACS Paragon Plus Environment

Page 23 of 31

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

1 2

Figure 1. Differential adsorption energies (red square) and average adsorption

3

energies (black circle) of formate on Cu(211). All the adsorption energies are with

4

respect to the energy of CO2(g) + 1/2 H2(g).

5

23 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

1 2

Figure 2. Adsorption structures (above) of HCOO, HCOOH, CH2OOH, CH2O, OH,

3

CH3O and H, and transition-state structures (below) of H-COO, HCOO-H,

4

H-CHOOH, CH2O-OH, H-CH2O, CH3O-H and H-OH on clean Cu(211). In this figure

5

and those hereafter, the Cu, C, H, and O atoms are represented in orange, gray, white

6

and red, respectively.

24 ACS Paragon Plus Environment

Page 24 of 31

Page 25 of 31

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

1 2

Figure 3. Configurations of the adsorbed intermediates and corresponding

3

transition-states of the elementary steps of methanol synthesis on Cu(211) with one

4

and two pre-adsorbed formate. The spectating formate species are shown in stick type,

5

while the reaction intermediates are presented as ball-and-stick.

6

25 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

1 2

Figure 4. Potential energy diagrams of the reaction pathway for methanol synthesis

3

on Cu(211) with different coverages of surface species (θDFT) on the surface. The

4

elementary reactions considered are listed in Table 1.

5

26 ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31

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

1 2

Figure 5. Charge density difference plots for the adsorption of HCOOH over Cu(211)

3

with different number of formate species pre-adsorbed. The yellow contours represent

4

charge accumulations, and the blue contours denote charge depressions. The scales of

5

the charge densities are identical.

6

27 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

1 2

Figure 6. (a-c) The calculated coverages of HCOO as a function of temperature at

3

preactant = 5, 25 and 50 bar, respectively. θDFT = 0.083, 0.167 and 0.250 ML are shown

4

in yellow square, red circle and blue triangle, respectively. The corresponding

5

calculated TOFs as a function of temperature are presented in (d-f). The feed

6

composition is CO2(g) : H2(g) : CH3OH(g) : H2O(g) : inert gas = 0.10: 0.40 : 0.01 :

7

0.01 : 0.48 at the steady-state in two tables and those hereafter.

28 ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31

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

1 2

Figure 7. Degree of rate control for methanol synthesis from CO2 hydrogenation via

3

HCOO intermediate for the rate-determining elementary steps versus temperature at

4

industrial reaction conditions (preactant = 50 bar). a, b and c are corresponding to θDFT =

5

0.083, 0.167 and 0.250 ML, respectively. Values for those transition-states and

6

adsorbates close to zero are not shown here. The XHCOO-H, XH-CHOOH, XCH2O-OH and

7

XHCOO are marked as black square, red circle, yellow triangle and blue invert-triangle,

8

respectively.

29 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

1 2

Figure 8. The TOFs of methanol synthesis as a function of temperature at preactant = 50

3

bar obtained from CatMAP and equation (6) at θDFT = (a) 0.083, (b) 0.167 and (c)

4

0.250 ML.

5

30 ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31

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

1

TOC

2 3 4 5

6 7

31 ACS Paragon Plus Environment