Вы находитесь на странице: 1из 42

Accepted Manuscript

Title: Density-Functional Theory Study of Dimethyl


Carbonate Synthesis by Methanol Oxidative Carbonylation on
Single-Atom Cu1 /Graphene Catalyst

Authors: Wei Sun, Ruina Shi, Xuhui Wang, Shusen Liu,


Xiaoxia Han, Chaofan Zhao, Zhong Li, Jun Ren

PII: S0169-4332(17)31991-8
DOI: http://dx.doi.org/doi:10.1016/j.apsusc.2017.07.002
Reference: APSUSC 36536

To appear in: APSUSC

Received date: 27-4-2017


Revised date: 30-6-2017
Accepted date: 1-7-2017

Please cite this article as: Wei Sun, Ruina Shi, Xuhui Wang, Shusen
Liu, Xiaoxia Han, Chaofan Zhao, Zhong Li, Jun Ren, Density-Functional
Theory Study of Dimethyl Carbonate Synthesis by Methanol Oxidative
Carbonylation on Single-Atom Cu1/Graphene Catalyst, Applied Surface
Sciencehttp://dx.doi.org/10.1016/j.apsusc.2017.07.002

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Density-Functional Theory Study of Dimethyl Carbonate Synthesis by Methanol

Oxidative Carbonylation on Single-Atom Cu1/Graphene Catalyst

Wei Sun a, Ruina Shi a, Xuhui Wang a, Shusen Liu a, Xiaoxia Han b, Chaofan Zhao b,

Zhong Li a, Jun Ren a*

a
Key Laboratory of Coal Science and Technology (Taiyuan University of Technology),

Ministry of Education and Shanxi Province, No. 79 Yingze West Street, Taiyuan

030024, China
b
College of Information Engineering, Taiyuan University of Technology, No. 79

Yingze West Street, Taiyuan 030024, China

*Corresponding author. Mailing address for correspondence: No. 79 Yingze West

Street, Taiyuan 030024, China. Tel/Fax: +86 351 6018598.

E-mail address: renjun@tyut.edu.cn (J. Ren).


Graphical Abstract

The mechanism for dimethyl carbonate (DMC) formation by oxidation

carbonylation of methanol on a single-atom Cu1/graphene catalyst and the

rate-limiting steps (RDS) on Cu/MG and Cu/DG catalyst surfaces are depicted in this

picture. The RDS for the DMC sythesis on these two catalysts surfaces are CH3O +

CO → CH3OCO for the Cu/MG catalyst, CH3OCO + CH3O → DMC for the Cu/DG

catalyst, with maximum energy barriers of 73.5 kJ mol–1 and 190.9 kJ mol–1,

respectively. Compared with the energy barrier of these rate-determining steps, the

active performance of the Cu/MG catalyst is superior to the Cu/DG catalyst because

of the monvacancy in graphene sheet. Therefore, Cu/MG can be considred as the

promising catalyst for DMC foemation.


Highlights

 Mechanism for DMC synthesis was explored on a single-atom

Cu1/graphene catalyst.

 Carbon vacancy is beneficial to the stability of the Cu 1/graphene

catalyst.

 CO prefers to insert to methoxide species rather than dimethoxide

species.

 The productivity for DMC can be significantly improved by Cu/MG

catalyst.

ABSTRACT: The mechanism for dimethyl carbonate (DMC) synthesis by oxidation

carbonylation of methanol on a single-atom Cu1/graphene catalyst was investigated by

density-functional theory calculations. Carbon vacancies in graphene can significantly

enhance the interaction between Cu atoms and graphene supports, and provide an

increased transfer of electrons from Cu atoms to the graphene sheet. Compared with

Cu-doped divacancy graphene (Cu/DG), Cu-doped monovacancy graphene (Cu/MG)

provides a stronger interaction between adsorbents and the catalyst surface. Among

the reaction processes over Cu1/graphene catalysts, CO insertion into methoxide was

more favorable than dimethoxide. The rate-limiting step on the Cu/DG surface is the

carbomethoxide reaction with methoxide, which is exothermic by 164.6 kJ mol−1 and

has an activation barrier of 190.9 kJ mol−1 energy. Compared with that on the Cu

crystal surface, Cu4 and Cu3Rh clusters, and the Cu2O(111) surface, the

rate-determining step for DMC formation on Cu/MG, which is CO insertion into

methoxide, needs to overcome the lowest barrier of 73.5 kJ mol−1 and is exothermic
by 44.6 kJ mol−1. Therefore, Cu/MG was beneficial to the formation of DMC as a

single-atom catalyst.

Keywords: Single-atom catalyst, Cu-doped graphene, Vacancy, Dimethyl carbonate

1. Introduction

Dimethyl carbonate (DMC) has been deemed to be an environmentally green

chemical product and has been applied widely in different fields, such as an

oxygen-containing additive for gasoline, as a potential alternative substitute for

phosgene and dimethyl sulfide and as an important intermediate in synthetic medicine

[1]. Among various synthesis methods, oxidative carbonylation of methanol to DMC

has gained attention because of its abundant, inexpensive raw materials and its

environmentally friendly process [2, 3]. To solve several problems related to

chloride-containing Cu-based catalysts, such as equipment corrosion and deactivation,

which are caused by the loss of chloride during the reaction, the development of

chloride-free catalysts has attracted attention worldwide [4, 5]. Recently, highly

dispersed Cu species with various valence states (CuO, Cu2O, Cu0) on active carbon

(AC) have been identified as promising candidates to compete with Cu-exchanged

zeolites catalysts [6-16]. Ren et al. [17] reported that AC-supported hollow copper

nanoparticle catalysts displayed a prominent activity with methanol conversion and a

space–time yield of DMC of 5.8% and 596.9 mg g−1cat−1. The space–time yield of

DMC formation on CuO/AC, Cu2O/AC, and Cu0/AC catalysts is 134.0, 216.6, and

261.9 mg g−1h−1, respectively [5]. Our previous study [18] provided strong evidence

by density-functional-theory (DFT) calculations that the catalytic performance of

copper with various valence states in DMC formation increased as: Cu2+ < Cu+ < Cu0.
However, the aggregation and loss of active centers are the main reasons for a

significant decrease in activity of the Cu-based catalysts. More interestingly,

enveloping metal particles into a specific confined space would yield an improved

catalytic performance. Wang et al. [19] prepared ordered mesoporous carbon supports

for Cu catalysts for DMC synthesis. We [20] have reported that carbon-based yolk–

shell nanospheres with a distinct cavity can immobilize copper nanoparticles. Cu@C

catalysts show an excellent catalytic stability towards DMC synthesis. Therefore, a

limited aggregate and loss is necessary to design Cu-based catalysts for DMC

synthesis.

The size of the metal particles that are anchored on the supports is one of the

most important factors to modify catalyst performance [21-24]. The free energy of the

metal cluster increases, whereas the metal particle size decreases [25]. Small,

single-atom catalysts (SACs) have received considerable attention because of their

higher catalytic activity, minimal material usage, and high atom efficiency [25, 26].

Theoretical calculations predict that atomically dispersed metal on oxide supports

would exhibit excellent catalytic and selectivity in heterogeneous catalysis [27, 28].

The first practical single-atom catalyst, Pt1/FeOx, was reported by Zhang and his

collaborators [29] and showed an excellent stability and high activity for CO

oxidation. Jones et al. [30] discovered that single-atom Pt on a ceria catalyst can

modify the rates of catalyst sintering and enhance the activity and selectivity. A

practical Pd SAC, which is supported by TiO2, was prepared experimentally by Liu et

al. [31], and it exhibited excellent catalytic performance in aldehyde hydrogenation.

In recent research, SACs with graphene support have been reported to be

excellent catalysts in particular reactions. Graphene is a novel carbon material with

one-atom-thick crystals, a high theoretical specific-surface area and unique electronic


and geometric properties, and it is regarded as a superior support for heterogeneous

catalysts [32-36]. For graphene-based SACs, vacancies in the graphene can promote

the interaction between the adatom and graphene sheet. Therefore, monoatomic metal

species prefer to occupy one or more carbon sites in the graphene lattices [37-41].

Wang et al. [42] has reported that substitutional single-metal-atom (Pt, Co and

In)-doped monolayer graphene was obtained via a two-step process: create vacancies

by high-energy atom/ion bombardment and fill these vacancies with metal dopants.

Chemical-vapor deposition and atomic layer deposition can also achieve the goal of

synthesis graphene-based SACs [43, 44]. Isolated Au [45], Cu [46], Fe [47], Mo [48],

Pt [49], Al, and Ce [50] that are embedded in a graphene lattice present great activity

and selectivity for the CO oxidation reaction. Atomic gold that is stabilized by

defective graphene sheets has been predicted to be a potential catalyst for ethylene

epoxidation [51]. Among various defects, monovacancy graphene (MG) and divacancy

graphene (DG) in graphene sheets can affect the stability and catalytic activity of metal

dopants. Recently, the influence of vacancies on transition-metal (Sc–Zn)-doped

graphene has been examined by DFT and the vacancy type has been found to contribute

to the adsorption [52]. It has been reported that the reaction mechanism of CO

oxidation on Fe-graphene (Fe/MG and Fe/DG) catalyst surfaces is different [53]. As a

result, metal atomical dispersion on defective graphene could be a better catalyst,

where the stability and lifetime can be improved.

Single-atom Cu-doped graphene catalyst has been used in CO oxidation [46] and

CO2 hydrogenation to formic acid [54]. As mentioned previously, intensive work to

explore the structure and catalytic performance of a single-atom Cu1/graphene catalyst

is highly desirable. In this work, we study the geometry and catalytic properties of

single-atom Cu-doped mono- and divacancy graphene (MG and DG), which can be
defined as Cu/MG and Cu/DG. Details of the geometrical structures, energies, and

reaction mechanisms along the DMC formation process will be discussed and

compared with related systems.

2. Computational methods

All the periodic DFT calculations were performed by using Dmol3 code

implemented in the Materials Studio 5.5 package (Accelrys Ltd.) [55-57]. The

hexagonal graphene supercell (6 × 6 unit cell) that contains 72 carbon atoms was used

to mimic pristine graphene and Cu-graphene composites. The modulus unit-cell

vector in the z direction was set to 20 Å to avoid interactions with its own periodic

image. The generalized gradient approximation (GGA) potentials with the Perdew,

Burke and Ernzerhof (PBE) functional [58] and double numeric polarized (DNP)

basis set [59] were used to evaluate the nonlocal exchange correlation energy. In our

calculation, the inner electrons of the copper atoms were kept frozen and they were

superseded by an effective core potential (ECP) [59, 60]. To search for the transition

states (TS) and to explore the reaction mechanisms of DMC, a complete linear

synchronous transit/quadratic synchronous transit (LST/QST) in the DMol3 code was

used [61]. A frequency analysis has been performed to validate the optimized TS

structures. Charge transfers were calculated by utilizing the Hirshfeld charge analysis

method [62]. Methfessel–Paxton smearing was set to 0.005 Hartree. The convergence

criteria of self-consistent field (SCF) energy was set to 1 × 10–5 Hartree and that of

displacement was 5 × 10–3 Å. The convergence criteria to optimize the energy and

maximum force were 2 × 10–5 Hartree and 0.004 Hartree/Å. We set the threshold

value for the self-consistent field density convergence to 1.0 × 10–5 Hartree.
3. Results and Discussion

3.1. Adsorption of Cu adatom on defective graphene

We investigated the structure of a single Cu atom that is anchored on defective

graphene (including MG and DG) first. The stable configuration of Cu/MG and Cu/DG

complexes after geometry optimization is shown in Fig. 1 and Table 1, which

summarizes the relevant binding energies, bond length, and transferred electrons. Here,

the binding energy (Eb) of single Cu-atom-doped graphene (Cu/MG and Cu/DG) is

calculated from:

Eb = ECu1/graphene − (ECu + EG) (1)

where ECu1/graphene, ECu, and EG are the total energies of Cu-doped graphene (Cu/MG

or Cu/DG), single Cu atom, and the defective graphene sheet (MG or DG).

The trapping of single-Cu atoms at monovacancy sites of graphene results in a

threefold coordination for the Cu atom and a saturation of all carbon bonds. Fig. 1(a)

shows a Cu atom that is stable on top of the monovacancy and interacts with three

neighboring dangling carbon atoms. The binding energy of the Cu/MG can be

enhanced by 544.2 kJ mol−1, which expresses an increased stability over Cu dopant on

pristine graphene (Cu/PG, 229.4 kJ mol−1, supplied in Fig. S1). Because of an unequal

atomic radius with respect to carbon atoms, the Cu dopant protrudes from the graphene

sheet and lies 1.45 Å above the graphene surface. The average bond length between

the Cu and the neighboring carbon atom is ~1.874 Å. A population analysis of the

Hirshfeld charge shows that, in the Cu/MG complex, the charge at Cu and at each ligate

C is +0.36689 e and –0.0891 e, respectively.

Fig. 1(b) shows the most stable geometric configuration of Cu/DG. In this

configuration, the Cu atom is located in the middle site of the DG sheet. Cu dopant

should be stabilized in a fourfold coordination and the size of the divacancy seems to
be approximately right for the capture of Cu impurities. In this case, atomic Cu dopant

is bonded equally to four adjacent carbon atoms and the average Cu–C bond length is

1.888 Å. Compared with Cu/MG, Cu/DG provides more binding energy (657.8 kJ

mol−1). Based on the population analysis of the Hirshfeld charge, approximately 0.3115

e charge moves from Cu to the DG sheets. Although a single Cu atom provides more

electrons to the MG sheet, the relatively stronger interaction of Cu with the DG

substrate compared with the MG likely results because the bare DG is more stable than

the bare MG. The literature indicates that DG is more stable than MG

thermodynamically, from a comparison of their relative energies and evaluating

migration barriers. As shown in Fig. 1(a), the formation of an MG configuration leads

to the three dangling bonds. As shown in Fig. 1(c), no dangling bond exists in the DG

configuration, so two pentagons and one octagon (5-8-5) appear at the vacancy site

instead of two hexagons in the PG. The quenching of dangling bonds in the DG renders

a more stable structure than for the MG [53, 63, 64].

A strong interaction occurs between the metal nanoparticles and the support,

which prevents metal atoms from aggregating and maintains their catalytic activity for

long-term use. It can be concluded that the defective graphene provides more

transferred charges from the Cu dopant to the graphene sheets, which agrees with

previous reports [52, 65]. Cu/DG is more stable than Cu/PG and Cu/MG. The results

show that Cu adatoms can diffuse easily onto PG, but they are fixed in defect sites of

MG and DG.

3.2. Adsorption of reactants and possible intermediates

For the adsorption study, the adsorption energy Eads, is always considered a valid

measure of interaction between adsorbate and substrate. Eads is defined as:

Eads(A) = (ECu1/graphene + EA) – EA+Cu1/Graphene (2)


where EA, ECu1/Graphene, and EA+Cu1/Graphene represent the total energies of the free

molecule or intermediate of the corresponding adsorbate, the Cu-doped graphene

sheets (Cu/MG or Cu/DG) and the Cu-doped graphene sheets with adsorbate.

In this section, we consider the adsorptions of all possible reactants, intermediates,

and products involved in the reaction of DMC synthesis on Cu/MG and Cu/DG

catalysts, respectively. The most stable adsorption geometries of these species that is

obtained from our DFT calculations are shown in Figs 2 and 3, and the corresponding

adsorption sites and adsorption energies are listed in Table 2. Compared with Cu/DG,

Cu/MG provides a much larger adsorption energy for all possible reactants,

intermediates, and products. For CH3OH adsorption, the adsorption energy of CH3OH

on the Cu/MG surface is slightly higher than that on Cu/DG (77.9 vs 56.7 kJ mol–1).

The charge of CH3OH adsorbed on the Cu/MG is positive by 0.1821 e. However, the

charge of CH3OH on Cu/DG is −0.0870 e, which means that approximately 0.0870 e

electrons move from the Cu/DG catalyst to the CH3OH molecule. The result of the

charge analysis shows that the interaction between CH3OH molecule and the Cu/MG

catalyst is slightly stronger than that for the Cu/DG [66, 67]. In the structure of the

CH3OH adsorbed on Cu/DG, the bond lengths of O–H, C–O, and C–H are 0.972, 1.428

and 1.106 Å, respectively, which are close to those of the free molecule (O–H, 0.971 Å;

C–O, 1.428 Å; C–H, 1.105 Å). Although the bond lengths of O–H and C–H in CH3OH

on the Cu/MG surface (O–H, 0.972 Å; C–H, 1.100 Å) are close to those of the free

CH3OH molecule, the C–O bond length stretches from 1.428 to 1.448 Å. The reactions

of CO insertion into methoxide or dimethoxide are rate-determining steps of two


different synthesis pathways, which plays a decisive part in the reaction [6, 18, 68, 69].

CO interacts with Cu/DG more weakly than with Cu/MG, and the adsorption energy of

CO on Cu/DG is only 10.5 kJ mol−1. Hence, we predict that the Cu/MG catalyst will

exhibit a superior catalytic performance in DMC synthesis. In Section 3.3, we

investigate in detail the mechanism of DMC formation on a Cu1/graphene catalyst to

verify our assumption.


3.3. Reaction mechanism

DMC can be formed by two pathways: insertion of CO into dimethoxide species

(path 1) or methoxide species (path 2) according to the experimental and theoretical

calculations in literature [6-8, 68-71]. The proposed reaction mechanism is shown in

Scheme 1. The activation barrier and reaction energy will be investigated

systematically by DFT calculations to explore the reaction mechanism of DMC

synthesis over Cu/MG and Cu/DG surfaces. Differences in DMC synthesis between

the Cu/MG and Cu/DG surfaces are studied to determine the effect of structures of

Cu-doped graphene. Here, the energy barrier Ea and reaction energy ∆H are calculated

from Eqs. (3) and (4). The ∆H (kJ mol−1) and Ea (kJ mol−1) of these steps are

presented in Table 3.

Ea = ETS - EIS (3)

∆H = EFS - EIS (4)

where EIS, ETS, and EFS are defined as the total energies of the initial, transition, and

final states, respectively.

3.3.1 CH3OH oxidation

CH3O is formed by O-H bond cleavage of the CH3OH on the Cu1/graphene

surface in the oxidation of CH3OH. Before CH3OH oxidation, an oxygen molecule is

adsorbed on the Cu1/graphene surface and is decomposed to two individual O atoms.


The initial (ISs), transition (TSs), and final (FSs) states are shown in Fig. S2. The

imaginary frequencies of O2 decomposition on the Cu/MG and Cu/DG are –248.97

and –118.50 cm–1, respectively. On the Cu/MG surface, the activation barrier of O2

decomposition was 34.8 kJ mol−1 and the reaction energy was −304.3 kJ mol−1, which

is consistent with previous studies [18, 72, 73]. The decomposition barrier of O2 on

the Cu/DG surface is slightly lower than that on Cu/MG (25.8 kJ mol−1 vs. 34.8 kJ

mol−1), which is exothermic by −295.2 kJ mol−1.

The reaction mechanism for CH3O and OH formation is shown in R1. Fig. 4

shows the ISs, TSs, and FSs of CH3OH oxidation on two different Cu1/Graphene

sheets (Cu/MG and Cu/DG). The path to produce the CH3O and OH coadsorption

configuration on the Cu/MG surface is exothermic by 77.8 kJ mol−1 and needs to

overcome the relatively lower energy barrier by 57.6 kJ mol−1. In this reaction, atomic

O that is adsorbed on top of the catalyst surface is unstable and will react immediately

with methanol to form a methoxyl group and hydroxyl, which is coadsorbed on the Cu

site. However, CH3OH oxidation over the Cu/DG surface has a higher barrier of 177.5

kJ mol−1 with a reaction energy of 129.9 kJ mol−1. From the coadsorbed CH3OH and

O species, no bond exists between CH3OH and Cu/DG, whereas O is adsorbed at a

bridge site, which is more inactive. Cu/MG has a positive effect on CH3OH oxidation

but Cu/DG has a negative role.

3.3.2. Formation of dimethoxide species

The formation of a dimethoxide intermediate in path 1 was initialized by the

coadsorption complex between CH3OH, the CH3O group and OH over the Cu/MG, and

Cu/DG catalyst surfaces via R2. The CH3O group was formed through activation of the

O-H bond in the CH3OH. Once the dimethoxide intermediate was generated, H2O was

eliminated from the reaction system. Fig. 5 presents the structures of ISs, TSs, and FSs
of R2 on Cu/MG and Cu/DG surfaces. In the IS configurations of Cu/MG, CH3O and

OH groups are coadsorbed at the top of the Cu atom, whereas CH3OH exists above the

Cu-graphene surface, and the geometric structure of CH3OH is close to that of gaseous

free molecules. On the Cu/DG surface, although CH3O and OH are coadsorbed at the

top of the C atom, CH3OH also lies above the Cu-graphene surface. This occurs

because the adsorption energy of methanol on the catalyst surface is much weaker than

CH3O and OH. The CH3O+OH/CH3OH configuration (see Figs 5(a) and (d)) goes

through a transition state TS2 or TS7 (see Figs 5(b) and (e)), which leads to the

formation of (CH3O)2 and H2O species (see Figs 5(c) and (f)). The formation of

(CH3O)2 over a Cu/DG catalyst has a higher barrier of 133.2 kJ mol−1 with a slight

reaction energy of −5.4 kJ mol−1, but this reaction on Cu/MG is energetically

compatible, and is exothermic by 22.5 kJ mol−1 with an activation barrier of 68.3 kJ

mol−1.

3.3.3. CO insertion to dimethoxide

For R3, once (CH3O)2 formed in R2, CO can insert directly on the dimethoxide

via the transition state TS3 or TS8 to produce DMC on the Cu/MG and Cu/DG

catalyst surface in path 1, respectively. Fig. 6 presents the structures of ISs, TSs, and

FSs in R3 on the Cu/MG and Cu/DG surfaces. In the ISs of Cu/MG and Cu/DG, CO

is placed on the top site of the Cu1/graphene sheets. Thus, DMC is formed via the

Eley–Rideal (E–R) reaction in R3. In the TSs configuration (Figs 6(b) and (e)), the

CO molecule moves towards and reacts with the (CH3O)2 species to produce DMC.

This step on the Cu/MG needs to overcome a barrier of 154.8 kJ mol−1 and releases

160.7 kJ mol−1 of heat, whereas on Cu/DG, the barrier is higher by 305.2 kJ mol−1 and

the reaction energy is −280.5 kJ mol−1. The activation energy of CO insertion into

(CH3O)2 on the Cu1/graphene catalyst, especially on the Cu/MG surface, is lower than
that on the Cu4 catalyst surface [18]. This result shows that the insertion of CO into

dimethoxide is a rate-limiting step for path 1 of DMC formation, which agrees with

that in previous work [18, 69].

3.3.4. CO insertion to methoxide

In path 2, carbomethoxide CH3OCO has been regarded as a vital intermediate in

DMC synthesis. Once CO and CH3O coadsorbed over the Cu-graphene surface,

CH3OCO species can be formed via R4. The structures of ISs, TSs, and FSs of the

CH3OCO formation of the atomic Cu-graphene surface (Cu/MG and Cu/DG) are

described in Fig. 7. In IS9, CO is adsorbed physically on the Cu/DG surface.

Therefore, the reaction of CH3OCO formation is followed by an E–R mechanism. The

activation barrier of this step on Cu/DG is calculated as 147.7 kJ mol−1 and the

reaction energy is −164.6 kJ mol−1. Different from the Cu/MG surface, the CO and

CH3O are coadsorbed on top of the Cu, and CH3OCO formation is followed by the

Langmuir–Hinshelwood (L–H) reaction on the Cu/MG surfaces. This step is slightly

exothermic by 44.6 kJ with a lower activation barrier of 73.5 kJ mol−1. These results

show that the single-atom catalyst of Cu/MG is more beneficial to the formation of a

CH3OCO intermediate than a Cu surface [17], a Cu4 cluster [18], or Cu2O(111) [69],

whereas the Cu/DG catalyst prevents the insertion of CO into CH3O.

3.3.5. DMC formation by carbomethoxide reaction with methoxide

The CH3OCO intermediate is important in DMC synthesis. In our calculation,

DMC is the product of the direct reaction of carbomethoxide with methoxide. The ISs,

TSs, and FSs structures of the methoxide reaction with carbomethoxide on the

Cu/MG and Cu/DG surfaces are presented in Fig. 8. When R5 commences, CH3OCO

and CH3O move closer and a new C–O bond is formed. The step that occurs on

Cu/MG is exothermic (64.5 kJ mol−1) and needs to overcome a rather low activation
barrier of 64.5 kJ mol−1, compared with Cu/DG. On Cu/DG, the barrier and reaction

energies are calculated as 190.9 kJ mol−1 and −115.7 kJ mol−1.

3.4. Main formation pathway of DMC

Based on our calculation, the potential-energy diagram of two possible pathways

of DMC formation from the oxidative carbonylation of methanol on Cu/MG and

Cu/DG are shown in Figs S2 and S3, respectively.

As shown in Fig. S2, the formation mechanism of DMC on Cu/MG can be

obtained. In path 1 of DMC formation, which includes R2 and R3 (black line), the

corresponding activity barriers are 54.6 kJ mol−1 and 154.8 kJ mol−1, respectively. CO

insertion into dimethoxide (R3) has the highest energy barrier and is considered the

rate-determining step for the first path of DMC formation. In the second path (red

line), the corresponding energy barriers of R4 and R5 are 73.5 kJ mol−1 and 64.5 kJ

mol−1, respectively. Therefore, R4 is the rate-limiting step for this path. The energy

profiles of the elementary steps for DMC formation on Cu/MG are shown in Fig. S2.

We conclude that CO adsorbed on Cu/MG prefers to insert into methoxide to produce

the vital intermediate, CH3OCO. The main formation pathway of DMC involves R1,

R4, and R5 steps. The first pathway is uncompetitive with the second pathway, which

agrees with previous literature [69].

As shown in Fig. S3, the mechanism of DMC formation on Cu/DG can be

described. In R1, the formation of CH3O on Cu/DG by the E–R mechanism needs to

overcome a higher energy than that on Cu/MG by the L–H mechanism. For the first

pathway of DMC formation on the Cu/DG surface, which includes R2 and R3 (see the

black line in Fig. S3), the corresponding energy barrier is 133.2 kJ mol−1 and 305.2 kJ

mol−1. For the second pathway of DMC formation on the Cu/DG, which involves R4

and R5 (see the red line in Fig. S3), the corresponding energy barrier is 147.7 kJ mol−1
and 190.9 kJ mol−1. A comparison of the energy diagram of DMC formation by two

possible pathways is given. The formation of DMC on Cu/DG goes through a second

pathway, which consists of R1, R4, and R5, which is the same as the synthesis

pathway of DMC on the Cu/MG catalyst surface. In this pathway, R5 has the highest

energy barrier of 190.9 kJ mol−1. Hence, the rate-controlling step is the reaction of

carbomethoxide with methoxide (R5) for DMC formation over the Cu/DG surface,

and is different from that over the Cu/MG surface.

Fig. 9 presents the optimal DMC formation pathway from the oxidative

carbonylation of methanol on the Cu/MG and Cu/DG surfaces. Although anchored

atomic Cu on the DG sheet (Cu/DG) is more stable than that on the MG sheet

(Cu/MG), Cu dopants in the Cu/MG provide more transferred electrons to

monovacancy graphene and are more positively charged, which regulates the

adsorption of CH3OH, O, and CO species. A comparison shows that Cu atoms in the

Cu/MG are more active. Therefore, Cu/MG catalysts provide excellent catalytic

performance for DMC formation via oxidative carbonylation of methanol, which

confirms the prediction in Section 3.2. As a result, different carbon vacancies in

graphene have a varying effect on the catalytic properties of Cu1/graphene catalysts

and monovacancies in the Cu1/graphene catalyst can improve the catalytic activity in

DMC formation.

3.5. Catalytic performance of Cu1/graphene catalyst

The catalytic performance of a single-atom Cu catalyst with a graphene support

can be influenced by monovacancies, which move more charge from Cu dopants to

graphene sheets. A comparison of the energy barrier of the rate-determining step on

the Cu/MG together with that on Cu4 [18], Cu3Ru [74], Cu(111), Cu(220), Cu(200)

[17], and Cu2O(111) [69] is presented in Fig. 10. The catalytic performance of Cu/MG
is superior to other Cu-based catalysts, although the energy barrier of the

rate-determining step on Cu/DG is slightly higher than that on the Cu(200) surface.

Therefore, a single Cu atom that is anchored on monovacancy graphene (Cu/MG) is

regarded as the best single-atom catalyst for the reaction of DMC formation via

oxidative carbonylation of methanol.

3.6. General discussion

The adsorption energies of single-atom Cu on PG, MG, and DG sheets are

−229.4, −544.2, and −657.8 kJ mol−1, respectively. This result indicates that vacancies

in graphene can enhance the interaction between atomic Cu and graphene substrates

and promote charge transfer from Cu to the graphene sheets. With a significant

increase in the stability of the Cu1/graphene catalyst, this result is relevant in terms of

limiting the mobility of Cu species and their tendency to aggregate and sinter, which

may solve the problem of a loss of active centers in the reaction of DMC synthesis.

Reaction mechanisms of DMC formation over atomic Cu-doped graphene

catalyst surfaces are determined from our calculation. The key issue in DMC

synthesis by CH3OH oxidative carbonylation is CO insertion into methoxide or

dimethoxide species. The basic calculations reveal that CO insertion into methoxide is

more thermodynamically and kinetically favorable. As evaluated by the reaction

barriers, the reaction of CO insertion into methoxide on the Cu/MG catalyst is a

rate-determining step with a barrier of 73.5 kJ mol−1. However, the rate-determining

step in the main pathway of DMC on the Cu/DG catalyst surface is the

carbomethoxide reaction with methoxide together with an activation barrier of 190.9

kJ mol−1. During the reaction, Cu/MG is a better catalyst and presents outstanding

catalytic performance over Cu/DG.


4. Conclusion

SAC Cu1/graphene for DMC formation via the oxidative carbonylation of

methanol was investigated by DFT calculations. Carbon vacancies in graphene

stabilize single-metal Cu atoms on the graphene surface. The formation mechanisms

of DMC synthesis over Cu1/graphene catalysts in two different pathways are

explained in detail. Our calculation reveals that the second pathway is dominant for

DMC synthesis where CH3OCO is the significant reactive intermediate. The

activation energy of the rate-determining step on the Cu/MG catalyst is lower than the

rate-determining step on the Cu/DG catalyst (73.5 kJ mol−1 vs. 190.9 kJ mol−1). In

conclusion, single-atom Cu-doped monovacancy graphene substrates (Cu/MG)

exhibit a more promising activity for DMC synthesis via the oxidative carbonylation

of CH3OH than other Cu-based catalysts, which provides a new guide for the

experimental design of single-atom Cu catalysts in DMC synthesis.

Acknowledgements

This work has been supported by a grant from the National Natural Science

Foundation of China (21376159 and 21606159).

Appendix A. Supplementary data

Supplementary data associated with this article can be found in the attachment.
Reference

[1] M.A. Pacheco, C.L. Marshall, Review of dimethyl carbonate (DMC) manufacture

and its characteristics as a fuel additive, Energ. Fuel. 11 (1997) 2-29.

[2] S. King, Reaction mechanism of oxidative carbonylation of methanol to dimethyl

carbonate in Cu-Y zeolite, J. Catal. 161 (1996) 530-538.

[3] I.J. Drake, K.L. Fujdala, A.T. Bell, T.D. Tilley, Dimethyl carbonate production via

the oxidative carbonylation of methanol over Cu/SiO2 catalysts prepared via molecular

precursor grafting and chemical vapor deposition approaches, J. Catal. 230 (2005)

14-27.

[4] R. Jiang, Y. Wang, X. Zhao, S. Wang, C. Jin, C. Zhang, Characterization of catalyst

in the synthesis of dimethyl carbonate by gas-phase oxidative carbonylation of

methanol, J. Mol. Catal. A: Chem. 185 (2002) 159-166.

[5] R. Wang, Z. Li, H. Zheng, K. Xie, Preparation of chlorine-free Cu/AC catalyst and

its catalytic properties for vapor phase oxidative carbonylation of methanol, Chinese J.

Catal. 31 (2010) 851-856.

[6] Y. Zhang, A.T. Bell, The mechanism of dimethyl carbonate synthesis on

Cu-exchanged zeolite Y, J. Catal. 255 (2008) 153-161.

[7] S.A. Anderson, T.W. Root, Kinetic studies of carbonylation of methanol to dimethyl

carbonate over Cu+X zeolite catalyst, J. Catal. 217 (2003) 396-405.

[8] S.A. Anderson, T.W. Root, Investigation of the effect of carbon monoxide on the

oxidative carbonylation of methanol to dimethyl carbonate over Cu+X and Cu+ZSM-5

zeolites, J. Mol. Catal. A: Chem. 220 (2004) 247-255.

[9] H. Zheng, J. Qi, R. Zhang, Z. Li, B. Wang, X. Ma, Effect of environment around the

active center Cu+ species on the catalytic activity of CuY zeolites in dimethyl carbonate

synthesis: a theoretical study, Fuel Process. Technol. 128 (2014) 310-318.


[10] A. Rybakov, I. Bryukhanov, A. Larin, G. Zhidomirov, Theoretical aspects of

methanol carbonylation on copper-containing zeolites, Petrol. Chem+. 56 (2016)

259-266.

[11] I.J. Drake, Y. Zhang, D. Briggs, B. Lim, T. Chau, A.T. Bell, The local environment

of Cu+ in Cu-Y zeolite and its relationship to the synthesis of dimethyl carbonate, J.

Phys. Chem. B 110 (2006) 11654-11664.

[12] G. Zhang, Z. Li, H. Zheng, Z. Hao, X. Wang, J. Wang, Influence of surface

oxygenated groups on the formation of active Cu species and the catalytic activity of

Cu/AC catalyst for the synthesis of dimethyl carbonate, Appl. Surf. Sci. 390 (2016)

68-77.

[13] B. Yan, S. Huang, S. Wang, X. Ma, Catalytic oxidative carbonylation over Cu2O

nanoclusters supported on carbon materials: the role of the carbon support,

ChemCatChem 6 (2014) 2671-2679.

[14] G. Zhang, Z. Li, H. Zheng, T. Fu, Y. Ju, Y. Wang, Influence of the surface

oxygenated groups of activated carbon on preparation of a nano Cu/AC catalyst and

heterogeneous catalysis in the oxidative carbonylation of methanol, Appl. Catal. B:

Environ. 179 (2015) 95-105.

[15] J. Ren, M. Ren, D. Wang, J. Lin, Z. Li, Mechanism of microwave-induced

carbothermic reduction and catalytic performance of Cu/activated carbon catalysts in

the oxidative carbonylation of methanol, J. Therm. Anal. Calorim. 120 (2015)

1929-1939.

[16] K. Shi, S.-Y. Huang, Z.-Y. Zhang, S.-P. Wang, X.-B. Ma, Novel fabrication of

copper oxides on AC and its enhanced catalytic performance on oxidative

carbonylation of methanol, Chinese Chem. Lett. 28 (2016) 70-74.

[17] M. Ren, J. Ren, P. Hao, J. Yang, D. Wang, Y. Pei, J.-Y. Lin, Z. Li, Influence of
microwave irradiation on the structural properties of carbon-supported hollow copper

nanoparticles and their effect on the synthesis of dimethyl carbonate, ChemCatChem 8

(2016) 861-871.

[18] J. Ren, W. Wang, D. Wang, Z. Zuo, J. Lin, Z. Li, A theoretical investigation on the

mechanism of dimethyl carbonate formation on Cu/AC catalyst, Appl. Catal. A: Gen.

472 (2014) 47-52.

[19] X. Wang, T. Fu, H. Zheng, G. Zhang, Z. Li, The influence of the pore structure in

ordered mesoporous carbon over the formation of Cu species and their catalytic activity

towards the methanol oxidative carbonylation, J. Mater. Sci. 51 (2016) 5514-5528.

[20] P. Hao, J. Ren, L. Yang, Z. Qin, J. Lin, Z. Li, Direct and generalized synthesis of

carbon-based yolk-shell nanocomposites from metal-oleate precursor, Chem. Eng. J.

283 (2016) 1295-1304.

[21] E.C. Tyo, S. Vajda, Catalysis by clusters with precise numbers of atoms, Nat.

nanotechnol. 10 (2015) 577-588.

[22] Q. Wu, S. Xiong, P. Shen, S. Zhao, Y. Li, D. Su, A. Orlov, Exceptional activity of

sub-nm Pt clusters on CdS for photocatalytic hydrogen production: a combined

experimental and first-principles study, Catal. Sci. Technol. 5 (2015) 2059-2064.

[23] A.A. Herzing, C.J. Kiely, A.F. Carley, P. Landon, G.J. Hutchings, Identification of

active gold nanoclusters on iron oxide supports for CO oxidation, Science 321 (2008)

1331-1335.

[24] C.R. Henry, Catalytic activity of supported nanometer-sized metal clusters, Appl.

Surf. Sci. 164 (2000) 252-259.

[25] X.-F. Yang, A. Wang, B. Qiao, J. Li, J. Liu, T. Zhang, Single-atom catalysts: a new

frontier in heterogeneous catalysis, Accounts Chem. Res. 46 (2013) 1740-1748.

[26] J. Liu, Catalysis by supported single metal atoms, ACS Catal. 7 (2016) 34-59.
[27] C.-M. Wang, K.-N. Fan, Z.-P. Liu, Oxide-supported single gold catalyst for

selective hydrogenation of acrolein predicted from first principles, J. Catal. 266 (2009)

343-350.

[28] C.-Q. Lv, J.-H. Liu, Y. Guo, X.-M. Li, G.-C. Wang, DFT+U investigation on the

adsorption and initial decomposition of methylamine by a Pt single-atom catalyst

supported on rutile (110) TiO2, Appl. Surf. Sci. 389 (2016) 411-418.

[29] B. Qiao, A. Wang, X. Yang, L.F. Allard, Z. Jiang, Y. Cui, J. Liu, J. Li, T. Zhang,

Single-atom catalysis of CO oxidation using Pt1/FeOX, Nat. Chem. 3 (2011) 634-641.

[30] J. Jones, H. Xiong, A.T. DeLaRiva, E.J. Peterson, H. Pham, S.R. Challa, G. Qi, S.

Oh, M.H. Wiebenga, X.I.P. Hernández, Thermally stable single-atom

platinum-on-ceria catalysts via atom trapping, Science 353 (2016) 150-154.

[31] P. Liu, Y. Zhao, R. Qin, S. Mo, G. Chen, L. Gu, D.M. Chevrier, P. Zhang, Q. Guo,

D. Zang, Photochemical route for synthesizing atomically dispersed palladium

catalysts, Science 352 (2016) 797-800.

[32] A.K. Geim, K.S. Novoselov, The rise of graphene, Nat. Mater. 6 (2007) 183-191.

[33] H. Zhao, Y.-C. Lin, C.-H. Yeh, H. Tian, Y.-C. Chen, D. Xie, Y. Yang, K. Suenaga,

T.-L. Ren, P.-W. Chiu, Growth and raman spectra of single-crystal trilayer graphene

with different stacking orientations, ACS nano. 8 (2014) 10766-10773.

[34] S. Frindy, A. El Kadib, M. Lahcini, A. Primo, H. García, Copper nanoparticles

supported on graphene as an efficient catalyst for A3 coupling of benzaldehydes, Catal.

Sci. Technol. 12 (2016) 4306-4317.

[35] Y. Devrim, A. Albostan, Graphene-Supported Platinum Catalyst-Based Membrane

Electrode Assembly for PEM Fuel Cell, J. Electron. Mater. 45 (2016) 3900-3907.

[36] S. Hsieh, M. Hsu, W. Liu, W. Chen, Study of Pt catalyst on graphene and its

application to fuel cell, Appl. Surf. Sci. 277 (2013) 223-230.


[37] L. Zhang, Q. Xu, J. Niu, Z. Xia, Role of lattice defects in catalytic activities of

graphene clusters for fuel cells, Phys. Chem. Chem. Phys. 17 (2015) 16733-16743.

[38] K.T. Chan, J. Neaton, M.L. Cohen, First-principles study of metal adatom

adsorption on graphene, Phys. Rev. B 77 (2008) 235430.

[39] X. Liu, C. Z. Wang, M. Hupalo, W.C. Lu, M.C. Tringides, Y.X. Yao, K.M. Ho,

Metals on graphene: correlation between adatom adsorption behavior and growth

morphology, Phys. Chem. Chem. Phys. 14 (2012) 9157-9166.

[40] L. Hu, X. Hu, X. Wu, C. Du, Y. Dai, J. Deng, Density functional calculation of

transition metal adatom adsorption on graphene, J. Phys. Condens. Mat. 405 (2010)

3337-3341.

[41] H. Valencia, A. Gil, G. Frapper, Trends in the adsorption of 3d transition metal

atoms onto graphene and nanotube surfaces: a DFT study and molecular orbital

analysis, J. Phys. Chem. C 114 (2010) 14141-14153.

[42] H. Wang, Q. Wang, Y. Cheng, K. Li, Y. Yao, Q. Zhang, C. Dong, P. Wang, U.

Schwingenschlögl, W. Yang, X.X. Zhang, Doping monolayer graphene with single

atom substitutions, Nano lett. 12 (2011) 141-144.

[43] H.J. Qiu, Y. Ito, W. Cong, Y. Tan, P. Liu, A. Hirata, T. Fujita, Z. Tang, M. Chen,

Nanoporous graphene with single-atom nickel dopants: an efficient and stable catalyst

for electrochemical hydrogen production, Angew. Chem. 127 (2015) 14237-14241.

[44] S. Sun, G. Zhang, N. Gauquelin, N. Chen, J. Zhou, S. Yang, W. Chen, X. Meng, D.

Geng, M.N. Banis, R, Li, S. Ye, S. Knights, G. Botton, T.-K. Sham, X. Sun,

Single-atom catalysis using Pt/graphene achieved through atomic layer deposition, Sci.

Rep-UK (2013) 1775-1784.

[45] Y.-H. Lu, M. Zhou, C. Zhang, Y.-P. Feng, Metal-embedded graphene: a possible

catalyst with high activity, J. Phys. Chem. C 113 (2009) 20156-20160.


[46] E. Song, Z. Wen, Q. Jiang, CO catalytic oxidation on copper-embedded graphene,

J. Phys. Chem. C 115 (2011) 3678-3683.

[47] Y. Li, Z. Zhou, G. Yu, W. Chen, Z. Chen, CO catalytic oxidation on iron-embedded

graphene: computational quest for low-cost nanocatalysts, J. Phys. Chem. C 114 (2010)

6250-6254.

[48] Y. Tang, L. Pan, W. Chen, C. Li, Z. Shen, X. Dai, Reaction mechanisms for CO

catalytic oxidation on monodisperse Mo atom-embedded graphene, Appl. Phys. A 119

(2015) 475-485.

[49] X. Liu, Y. Sui, T. Duan, C. Meng, Y. Han, CO oxidation catalyzed by Pt-embedded

graphene: A first-principles investigation, Phys. Chem. Chem. Phys. 16 (2014)

23584-23593.

[50] M.D. Esrafili, P. Nematollahi, H. Abdollahpour, A comparative DFT study on the

CO oxidation reaction over Al- and Ge-embedded graphene as efficient metal-free

catalysts, Appl. Surf. Sci. 378 (2016) 418-425.

[51] X. Liu, Y. Yang, M. Chu, T. Duan, C. Meng, Y. Han, Defect stabilized gold atoms

on graphene as potential catalysts for ethylene epoxidation: a first-principles

investigation, Catal. Sci. Technol. 6 (2016) 1632-1641.

[52] H. Luo, H. Li, Z. Xia, Y. Chu, J. Zheng, Z. Hou, Q. Fu, Novel insights into

L-cysteine adsorption on transition metal doped graphene: influences of the dopant and

the vacancy, RSC Adv. 6 (2016) 29830-29839.

[53] Y. Tang, J. Zhou, Z. Shen, W. Chen, C. Li, X. Dai, High catalytic activity for CO

oxidation on single Fe atom stabilized in graphene vacancies, RSC Adv. 6 (2016)

93985-93996.

[54] J. Sirijaraensre, J. Limtrakul, Hydrogenation of CO2 to formic acid over a

Cu-embedded graphene: A DFT study, Appl. Surf. Sci. 364 (2016) 241-248.
[55] B. Delley, An all-electron numerical method for solving the local density

functional for polyatomic molecules, J. Chem. Phys. 92 (1990) 508-517.

[56] B. Delley, Fast calculation of electrostatics in crystals and large molecules, J. Phys.

Chem. 100 (1996) 6107-6110.

[57] B. Delley, From molecules to solids with the DMol3 approach, J. Chem. Phys. 113

(2000) 7756-7764.

[58] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made

simple, Phys. Rev. Lett. 77 (1996) 3865-3868.

[59] M. Dolg, U. Wedig, H. Stoll, H. Preuss, Energy-adjusted abinitio pseudopotentials

for the first row transition elements, J. Chem. Phys. 86 (1987) 866-872.

[60] A. Bergner, M. Dolg, W. Küchle, H. Stoll, H. Preuß, Ab initio energy-adjusted

pseudopotentials for elements of groups 13–17, Mol. Phys. 80 (1993) 1431-1441.

[61] T.A. Halgren, W.N. Lipscomb, The synchronous-transit method for determining

reaction pathways and locating molecular transition states, Chem. Phys. Lett. 49 (1977)

225-232.

[62] F.L. Hirshfeld, Bonded-atom fragments for describing molecular charge densities,

Theor. Chim. Acta 44 (1977) 129-138.

[63] A. Hashimoto, K. Suenaga, A. Gloter, K. Urita, S. Iijima, Direct evidence for

atomic defects in graphene layers, Nature 430 (2004) 870-873.

[64] F. Banhart, J. Kotakoski, A.V. Krasheninnikov, Structural defects in graphene,

ACS nano 5 (2010) 26-41.

[65] S. Back, J. Lim, N.Y. Kim, Y.H. Kim, Y. Jung, Single-atom catalysts for CO2

electroreduction with significant activity and selectivity improvements, Chem. Sci. 8

(2017) 1090-1096.

[66] J. Gomes, J. Gomes, A DFT study of the methanol oxidation catalyzed by a


copper surface, Surf. Sci. 471 (2001) 59-70.

[67] W.K. Chen, S.H. Liu, M.J. Cao, C.H. Lu, Y. Xu, J.Q. Li, Adsorption of methanol

and methoxy on Cu (111) Surface: a first-principles periodic density functional theory

study, Chinese J. Chem. 24 (2006) 872-876.

[68] Y. Shen, Q. Meng, S. Huang, J. Gong, X. Ma, DFT investigations for the reaction

mechanism of dimethyl carbonate synthesis on Pd(Ⅱ)/β zeolites, Phys. Chem. Chem.

Phys. 15 (2013) 13116-13127.

[69] R. Zhang, L. Song, B. Wang, Z. Li, A density functional theory investigation on the

mechanism and kinetics of dimethyl carbonate formation on Cu2O catalyst, J. Comput.

chem. 33 (2012) 1101-1110.

[70] U. Romano, R. Tesel, M.M. Mauri, P. Rebora, Synthesis of dimethyl carbonate

from methanol, carbon monoxide, and oxygen catalyzed by copper compounds, Ind.

Eng. Chem. Prod. Res. Dev. 19 (1980) 396-403.

[71] S. King, Oxidative carbonylation of methanol to dimethyl carbonate by solid-state

ion-exchanged CuY catalysts, Catal. Today 33 (1997) 173-182.

[72] Y. Xu, M. Mavrikakis, Adsorption and dissociation of O2 on Cu(111):

thermochemistry, reaction barrier and the effect of strain, Surf. Sci. 494 (2001)

131-144.

[73] J. Torras, C. Lacaze-Dufaure, N. Russo, J. Ricart, Chemisorption of molecular

oxygen on Cu(100): a Hartree-Fock and density functional study, J. Mol. Catal. A:

Chem. 167 (2001) 109-113.

[74] J. Ren, J. Yang, W. Wang, H. Guo, Z. Zuo, J. Lin, Z. Li, A DFT study of DMC

formation on Rh-doped Cu/AC surfaces, Int. J. Quantum Chem. 115 (2015) 853-858.
FIGURE CAPTION

Fig. 1. Top (left) and Side (right) views of optimized structure of (a) MG, (b) Cu/MG

and (c) DG and (d) Cu/DG. orange-black balls Cu atoms; red, O; grey ,C atoms.

Fig. 2. Optimized configurations and corresponding structure parameters of each

species and intermediates involved in DMC formation adsorbed on Cu/MG surface.

orange-black balls Cu atoms; red, O; grey ,C atoms; white, H atoms.

Fig. 3. Optimized configurations and corresponding structure parameters of each

species and intermediates involved in DMC formation adsorbed on Cu/DG surface.

orange-black balls Cu atoms; red, O; grey ,C atoms; white, H atoms.

Fig. 4. The configurations and structure parameters of the initial, transition and final

states for R1 on Cu/MG (a-c) and Cu/DG (d-f) surfaces.

Fig. 5. The configurations and structure parameters of the initial, transition and final

states for R2 on Cu/MG (a-c) and Cu/DG (d-f) surfaces.

Fig. 6. The configurations and structure parameters of the initial state, transition state

and final states for R3 on Cu/MG (a-c) and Cu/DG (d-f) surfaces.

Fig. 7. The configurations and structure parameters of the initial, transition and final

states for R4 on Cu/MG (a-c) and Cu/DG (d-f) surfaces.

Fig. 8. The configurations and structure parameters of the initial, transition and final

states for R5 on Cu/MG (a-c) and Cu/DG (d-f) surfaces.

Fig. 9. Optimal pathway of DMC formation via oxidative carbonylation of methanol on

Cu/MG and Cu/DG.

Fig. 10. Comparison of energy barrier of rete-determing steps on Cu/MG in together

with that on Cu4, Cu3Ru, Cu(111), Cu(220), Cu(200), Cu2O(111).


Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Scheme 1

CH3OH + O → CH3O + OH (R1)

Path 1

CH3O + OH + CH3OH → (CH3O)2 + H2O (R2)

(CH3O)2 + CO → CH3OCOOCH3 (R3)

Path 2

CH3O + CO → CH3OCO (R4)

CH3OCO + CH3O → CH3OCOOCH3 (R5)

The proposed mechanism of DMC formation.


Table 1

The binding energy (kJ mol−1), Charges Q of Cu and C atom as well as bond length of

Cu-C (Å) involved in Cu/PG, Cu/MG and Cu/DG sheets.

Catalysts Eb QCu * QC lCu-C

Cu/PG −229.4 0.2545 −0.0167 2.175

Cu/MG −544.2 0.3688 −0.0891 1.874

Cu/DG −657.8 0.3115 −0.0965 1.888

*
The positive charge of Cu means that the electrons transfer from Cu to MG or DG

sheet.
Table 2

Adsorption energy (kJ mol−1) and Charges Q of adsorption species involved in

dimethyl carbonate synthesis over Cu/MG and Cu/DG.

Cu/MG Cu/DG
species
Eads Sites Q* Eads Sites Q*

CH3OH 77.9 Top Cu 0.1821 56.7 Top Cu −0.0870

O 386.1 Top Cu −0.3855 109.3 Top Cu −0.2292

CH3O 269.5 Top Cu −0.2373 48.3 Top C −0.0880

OH 254.9 Top Cu −0.2616 99.6 Top C −0.0694

CO 136.3 Top Cu −0.0515 10.5 Top Cu 0.0508

CH3OCO 229.8 Top Cu −0.1620 74.2 Top C −0.0677

DMC 71.6 Top Cu 0.1561 -27.7 Top Cu −0.0191

H2 O 76.5 Top Cu 0.1714 66.3 Top Cu 0.0247

*
The positive charge of adsorbate means that there are some electrons moving from

adsorbate to Cu/MG or Cu/DG sheet, whereas for the negative charge, the electrons

transfer is reverse.
Table 3

Calculated reaction energies, ∆H(kJ mol−1) and activation energies, Ea(kJ mol−1) of the

elementary steps in DMC formation.

Elementary Cu/MG Cu/DG

steps TS (cm–1) Ea ∆H TS (cm–1) Ea ∆H

R1 TS1 (-220.87) 57.6 −78.9 TS6 (-457.61) 230.0 123.7

R2 TS2 (-368.86) 54.6 −1.0 TS7 (-402.61) 133.2 −5.4


Path 1
R3 TS3 (-409.32) 154.8 −160.7 TS8 (-213.44) 305.2 −280.5

R4 TS4 (-217.63) 73.5 −44.6 TS9 (-246.32) 147.7 −164.6


Path 2
R5 TS5 (-260.95) 64.5 −73.4 TS10 (-431.79) 190.9 −115.7

Вам также может понравиться