You are on page 1of 26

Subscriber access provided by Columbia Univ Libraries

Article
Competition between Eley-Rideal and Langmuir-Hinshelwood
pathways of CO oxidation on Cun and CunO (n = 6, 7) clusters
Yanbiao Wang, Guangfen Wu, Mingli Yang, and Jinlan Wang
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/jp3122775 • Publication Date (Web): 05 Apr 2013
Downloaded from http://pubs.acs.org on April 12, 2013

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.

The Journal of Physical Chemistry C 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 25 The Journal of Physical Chemistry

1
2
3
4 Competition between Eley-Rideal and Langmuir-Hinshelwood
5
6 pathways of CO oxidation on Cun and CunO (n = 6, 7) clusters
7
8
9 Yanbiao Wang1, Guangfen Wu1, Mingli Yang2, and Jinlan Wang1*
10
11 1
Department of Physics, Southeast University, Nanjing, 211189, P. R. China
12
13 2
14 Institute of Atomic and Molecular Physics, Sichuan University, Chengdu 610065,
15
16 China
17
18 *
19 Corresponding author: jlwang@seu.edu.cn
20
21
22
23
24 Abstract
25
26
27 The competition between the Eley-Rideal (ER) and Langmuir-Hinshelwood
28
29
30 mechanisms of CO oxidation on Cun and CunO (n = 6, 7) clusters was explored by
31
32 means of spin-polarized density functional theory calculations. The separate and
33
34
35 successive adsorptions of CO and O2 on the clusters were studied. CO and O2
36
37 molecules exhibit different adsorption behaviors and a cooperative effect was noted
38
39
40 for their co-adsorption. The reaction pathways of CO oxidization were then
41
42 investigated by locating the transition-state and intermediate structures. The ER
43
44
45 mechanism was found to be more favorable for the reactions on Cu6,7 and Cu6O, but
46
47
less favorable on Cu7O. The ER or LH preference of the CO + O2 reaction on the
48
49
50 clusters was further rationalized. We found that activation of O2 is the key issue that
51
52
affects the ER-LH competition. The pathway, either ER or LH, in which O2 is highly
53
54
55 activated is always preferred, while the O2 activation depends on its adsorption
56
57
pattern, site, and sequence in the presence of CO.
58
59
60 1

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 2 of 25

1
2
3
4 І. INTRODUCTION
5
6 Copper-based catalysts or catalyst promoters have attracted persistent interests
7
8
9 because of their wide applications in a variety of industrial processes.1–17 For example,
10
11 copper-based nanoparticles supported on oxide substrates demonstrate superior
12
13
14 catalysis for low temperature CO oxidation and resistance against water
15
16 contamination;1–5,12–14 The addition of Cu (as a promoter) into Pd–Cu–CIx/Al2O3
17
18
19 catalysts remarkably accelerates CO oxidation in the presence of water in reactant
20
21 gas.5
22
23
24 The interaction of copper nano-catalysts with small molecules including CO, O2,
25
26 H2, N2, NCO, H2O, and C2H2 have also been extensively studied.18–35 Size,
27
28
29 charge/oxidation state and pre-adsorbate of the nano-particles were found to play
30
31 important roles in their reactivity. For instance, Cun (n = 2–10) clusters exhibit distinct
32
33
34 size dependent adsorption behaviors toward molecular oxygen. The clusters with
35
36 odd-n generally possess larger adsorption energy than the even-n ones except Cu6O2
37
38
39 with a relatively large adsorption energy due to the adsorption–induced 2D–to–3D
40
41
structural transition in Cu6.31–32 The charge states of Cun also have great influences on
42
43
44 the adsorption sites and adsorption strength toward NCO.32 The pre–adsorption of CO
45
46
was found to enhance the reactivity of Cu atoms toward N2.31
47
48
49 The CO oxidation follows either an Eley-Rideal (ER) or a Langmuir-Hinshelwood
50
51
(LH) mechanism on catalyst surfaces.36–38 In the ER case, CO reacts with the
52
53
54 chemisorbed O2 on the surfaces, that is, the role of the catalyst is determined only by
55
56
57
its effect on O2. On the contrary, in the LH mechanism, the reaction involves a
58
59
60 2

ACS Paragon Plus Environment


Page 3 of 25 The Journal of Physical Chemistry

1
2
3
4 co-adsorption of O2 and CO on the catalyst and yields an intermediate in which both
5
6 molecules are activated.36 These two mechanisms are competitive depending on the
7
8
9 compositions of the catalysts and activation of adsorbates on them. For example, an
10
11 oscillatory mechanism of ER and LH was observed on platinum and iridium
12
13
14 surfaces,39–40 while the LH mechanism is preferential for the catalytic oxidation of CO
15
16 on Ni surfaces and Au55 cluster.37,41 Differently, the ER preference was found on W10
17
18
19 clusters and W(111) surface.43
20
21 What happens when CO comes to Cun–O2, or when O2 comes to Cun–CO? How do
22
23
24 the ER and LH mechanisms compete for the CO oxidation on the small copper
25
26 clusters? What is the effect of the pre-adsorption O atom on the reactivity and the
27
28
29 ER-LH competition? These issues are fundamental in understanding the catalytic
30
31 activity of Cu nanoparticles, but are scarcely known yet.
32
33
34 Size effect is important in catalysis. For example, small Aun– clusters with even n
35
36 exhibit relatively high activity toward O2, whereas the odd-sized clusters are inert.45
37
38
39 Small clusters are thus particularly interesting in the insight into the molecular-level
40
41
details of heterogeneous catalytic reactions.46 Cu6 is the largest Cu cluster with a
42
43
44 planar structure in ground state, while Cu7 is the smallest cluster with a
45
46
three-dimensional (3D) structure, making them proper representatives for small Cu
47
48
49 clusters. Moreover, structure change of clusters is an important phenomenon in gas
50
51
adsorption which has remarkable influence on the cluster reactivity. As we will show
52
53
54 below, a 2D–to–3D transition is found for Cu6 upon gas adsorption. In this work,
55
56
57
therefore, we choose Cu6, Cu7, Cu6O, and Cu7O clusters as prototypes to examine
58
59
60 3

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 4 of 25

1
2
3
4 their catalytic mechanisms of CO oxidation, as well as the ER-LH competition by
5
6 employing spin-polarized density functional theory (DFT) approach. Our calculations
7
8
9 show that the step of O2 activation is crucial for the ER or LH preference on small
10
11 copper clusters. If O2 itself is activated enough after adsorption, it may oxidize CO
12
13
14 directly through an ER pathway. If O2 activation is not enough without the
15
16 cooperative adsorption of CO, a LH pathway is preferred. Moreover, the reactions
17
18
19 proceed more easily if the transition-state geometries resemble those of the reactants
20
21 rather than the products.
22
23
24
25
26 II. COMPUTATIONAL METHOD
27
28
29 We exploited Becke exchange functional and the Lee-Yang-Parr correlation
30
31 functional (BLYP) as well as double numerical basis sets including d-polarization
32
33
34 functions (DNP), as implemented in the DMol3 package.47–55 Relativistic semi-core
35
36 pseudo potential (DSPP)56 for Cu atoms, and all electron basis sets for C and O atoms
37
38
39 were adopted. Complete linear synchronous transit (LST)/quadratic synchronous
40
41
transit (QST) calculations were performed to locate transition states. The obtained
42
43
44 local minima and the transition states were further verified via vibration frequency
45
46
calculations.57–61 The GGA/BLYP scheme is at a good level of theory for copper
47
48
49 cluster system,22,23 and has been evaluated in previous works.24-26,62 For example,
50
51
Fuentealba et al. studied the interaction of molecular oxygen with small copper
52
53
54 clusters Cun and Cun– (n = 1–7) at the same level, in which the calculated binding
55
56
57
energies, ionization potential, electron affinity, and vertical detachment energies of
58
59
60 4

ACS Paragon Plus Environment


Page 5 of 25 The Journal of Physical Chemistry

1
2
3
4 Cun or Cun– (n = 1–7) clusters are in good agreement with experimental data.24
5
6 Guzmán-Ramírez et al. compared a number of computations with the observations of
7
8
9 copper dimer and concluded that GGA/BLYP is appropriate for copper cluster
10
11 systems.25 Furthermore they studied the fragmentation channels and electronic
12
13
14 properties of Cunv (v = ±1, 0, 2; n = 3−13) clusters with BLYP calculations.26 In Table
15
16 1, we compared the BLYP and PBE results for bond lengths and frequencies of CO,
17
18
19 O2, CuO, and Cu2. They are also in good agreement with the measured values, which
20
21 suggests that the computational strategy employed in this work is adequate to produce
22
23
24 correct trend for the energetics and reactivity of the studied copper and copper oxide
25
26 clusters.
27
28
29 TABLE 1. Bond length (R), frequency (ω) and dissociation energy (DE) of CO, O2,
30
31 CuO, Cu2 from different theoretical methods and experiments.
32
33
34 BLYP PBE Expt.
35
R/Å ω/cm-1 DE/eV R/Å ω/cm-1 DE/eV R/Å ω/cm-1 DE/eV
36
37 CO 1.14 2124 11.15 1.14 2094.4 11.48 1.13a 2143.2a 11.1b
38
39 O2 1.23 1499 5.83 1.22 1580 6.26 1.21c 1580c 5.11d
40
41 CuO 1.72 684 3.31 1.71 694.3 3.51 1.72f 640f 2.75g
42
43 Cu2 2.26 259.6 1.98 2.25 261.8 2.13 2.22h 265h 2.08i
44 a
45 Ref. [63]. b Ref. [64]. c Ref. [65]. d Ref. [66]. f Ref. [67]. g Ref. [68]. h Ref. [69]. i Ref. [70].
46
47
48
49
50 To locate the energetically most favorable configurations, we considered various
51
52 possible adsorption patterns, including hollow (H)-, bridge (B)-, top (T)-site
53
54
55 adsorption, of CO and O2 on Cu6,7 and Cu6,7O clusters. The adsorption energy of
56
57 CO/O2 on Cun/CunO is defined as, Ead = E[cluster] + E[adsorbate] – E
58
59
60 5

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 6 of 25

1
2
3
4 [cluster-adsorbate], where E [*] is the total energy of relaxed Cun/CunO clusters, CO
5
6 or/and O2 adsorbates, and the cluster-adsorbate complexes respectively.
7
8
9
10
11 III. RESULTS AND DISCUSSION
12
13
14 A. Adsorption of the CO or/and O2 on Cu6,7 and Cu6,7O clusters
15
16 The low-lying isomers of Cun, Cun–O2, Cun–CO, OC–Cun–O2, CunO, CunO–O2,
17
18
19 OC–CunO, OC–CunO–O2, n = 6, 7, together with relative energies with respect to the
20
21 ground state structures are presented in Figures. 1 and 2, respectively. The bond
22
23
24 lengths of C–O, O–O and charge on CO and O2 in the complexes are given in Table 2.
25
26 Overall, O2 prefers H- or B-type adsorption, while CO tends to occupy T-sites in all
27
28
29 the cluster-adsorbate systems; the adsorption of O2 induces heavier structure
30
31 distortion on Cu6 and Cu6,7O than that of CO does. The O2 can always be effectively
32
33
34 activated upon adsorption, especially in the H-type adsorption, on both Cu6,7 and
35
36 Cu6,7O clusters with obvious charge transfer from the clusters to O2. On the contrary,
37
38
39 the C–O bond is hardly influenced upon the adsorption in all the clusters studied here.
40
41
42
43
44 TABLE 2. Bond lengths of C–O, O–O in Å and Charge on CO and O2 adsorbed in e.
45
46
The bold 1b,2b etc. correspond to Figure. 1b, Figure. 2b etc.
47
48
System Bond length Charge System Bond length Charge
49
50 1b 1.55 -0.48 2b 1.51 -0.45
51 Cu6O2 Cu7O2
1c 1.52 -0.46 2c 1.41 -0.41
52
53 1d 1.46 -0.46 2d 1.31 -0.41
54
55 Cu7CO
1e 1.15 -0.04 2e 1.15 -0.04
56 Cu6CO
57 2f 1.15 0.00
58
59
60 6

ACS Paragon Plus Environment


Page 7 of 25 The Journal of Physical Chemistry

1
2
3 1g 1.15 -0.03 Cu7O-O2 2j 1.41 -0.36
4
5
6 1j 1.52 -0.45 Cu7OCO 2k 1.16 -0.04
7 Cu6O-O2 1k 1.53 -0.46
8
9 1l 1.39 -0.33 Cu7CO-O2
10 (O-O) 1.49 -0.45
11 2f
Cu6OCO 1m 1.15 0.00 (C-O) 1.15 -0.02
12
13 (O-O) 1.53 -0.47
14 2g
Cu6CO-O2 (C-O) 1.15 -0.03
15
16 (O-O) 1.57 -0.51 (O-O) 1.54 -0.50
1h 2h
17 (C-O) 1.15 0.00 (C-O) 1.18 -0.16
18
19
Cu6OCO-O2 Cu7OCO-O2
20
21 (O-O) 1.56 -0.49 (O-O) 1.41 -0.37
1n 2l
22 (C-O) 1.15 0.01 (C-O) 1.15 -0.01
23
(O-O) 1.41 -0.36 (O-O) 1.42 -0.40
24 1o 2m
25 (C-O) 1.15 0.03 (C-O) 1.15 0.00
26 (O-O) 1.50 -0.44 (O-O) 1.40 -0.37
27 1p 2n
(C-O) 1.15 0.02 (C-O) 1.15 0.01
28
29
30
The structures of small Cun have been well studied. The ground states of Cu6 and
31
32 Cu7 are equilateral triangle and pentagonal bipyramidal, respectively. The additional
33
34
35
O atom slightly distorts the structure of Cu6 and Cu7. The vertex deflects to one side
36
37 in Cu6O. The lowest energy structures of Cu6O and Cu7O are an O-capped distorted
38
39
40 triangle and pentagonal bipyramidal respectively, as shown in Figures. 1i and 2i.
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 7

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 8 of 25

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 Figure. 1. Low-lying isomers of Cu6, Cu6O2, Cu6CO, Cu6–CO–O2, Cu6O, Cu6O–O2,
27
28
29 Cu6O–CO, Cu6O–CO–O2 together with relative energy ∆E (eV). Orange: Cu, red: O,
30
31 black: C.
32
33
34 The adsorption of O2 turns the planar structure of Cu6 into a 3D capped-bipyramid
35
36 (1b). The B-type isomer (1d) is 0.16 eV higher than the lowest energy state. Three
37
38
39 low-lying isomers of O2 adsorption on the pentagonal bipyramidal ground state of Cu7
40
41
(2a) are presented in Figures. 2b–2d. The B- and T-type isomers are 0.11 and 0.33 eV
42
43
44 higher in energy, respectively, with respect to the H-type ground state (2b). The
45
46
ground state (1e) of Cu6CO is evolved from the meta-stable pentagonal isomer of Cu6
47
48
49 (0.29 eV higher in energy). The two other isomers (1f, 1g) are based on the planar
50
51
ground state of Cu6 and are 0.11 and 0.19 eV higher in energy than the most stable
52
53
54 structure respectively. CO also takes an end-on manner in Cu7–CO (2e). The
55
56
57
activation of O–O bonds is closely related to the adsorption patterns. H-type
58
59
60 8

ACS Paragon Plus Environment


Page 9 of 25 The Journal of Physical Chemistry

1
2
3
4 adsorption generally has larger stretch and charge transfer than the other two types.
5
6 The different activated O–O bond lengths, 1.55 Å in Cu6O2 (1b) versus 1.51 Å in
7
8
9 Cu7O2 (2b), indicates that Cu6 is slightly more active than Cu7, maybe resulting from
10
11 the higher stability of Cu7 due to 2D-to-3D structure variation.
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 Figure. 2. Low-lying isomers of Cu7, Cu7O2, Cu7CO, Cu7–CO–O2, Cu7O, Cu7O–O2,
42
43 Cu7O–CO, Cu7O–CO–O2 together with relative energy ∆E (eV). Orange: Cu, red: O,
44
45
46 black: C.
47
48
49
50
51 The incoming O2 attacks Cu6O by two H-type and one B-type patterns. Three
52
53 low-lying isomers of Cu6O–O2, whose energy differences are less than 0.05 eV, are
54
55
56 displayed in Figures. 1j–1i. The adsorption of O2 leads to a remarkable distortion of
57
58
59
60 9

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 10 of 25

1
2
3
4 Cu7O, as shown in Figure. 2j, with the O2 occupying a bridge site. In Cu6OCO (1m),
5
6 CO shares a corner Cu atom with the atomic O. The coadsorption of CO and the
7
8
9 pre-adsorption of O lead to a remarkable distortion in Cu7O (see 2k). The activation
10
11 of O2 on the oxide clusters is relatively reduced in comparison with that on the
12
13
14 corresponding pure Cu6,7 clusters, as indicated by their shorter O–O bonds (1.52 vs.
15
16 1.55 Å and 1.41 vs. 1.51 Å for n = 6, 7 respectively).
17
18
19 In the co-adsorption of CO and O2, the molecule with the large adsorption energy is
20
21 in general likely to pre-adsorb onto the clusters. For Cu6,7 clusters, the adsorption
22
23
24 energies of O2 are 1.01 and 1.16 eV, respectively, larger than those of CO (0.87 and
25
26 0.92 eV). Therefore, the O2 tends to pre-adsorb on these clusters, followed by CO
27
28
29 attack. In OC–Cu6–O2 (1h), O2 and CO adsorb on the H- and T-sites respectively. The
30
31 three low-lying isomers of OC–Cu7–O2 are shown in 2f–2h respectively with CO and
32
33
34 O2 sitting on different sites. On the whole, the activation of O2 is promoted slightly by
35
36 the nearby subsequent attack of CO, reflected by the O–O bond length variations, 1.57
37
38
39 (1h) vs. 1.55 (1b) and 1.53 (2g) vs.1.51 (2b) Å respectively.
40
41
The CO adsorbs to Cu6O before O2 according to the adsorption energies of 1.93
42
43
44 (CO) vs. 1.46 eV (O2), and three low-lying isomers of OC–Cu6O–O2 are displayed in
45
46
Figures. 1n, 1o, 1p. On the contrary, the O2 attacks prior to CO on the Cu7O due to
47
48
49 larger adsorption energies of 2.43 (O2) vs. 1.67 eV (CO). Three isomers of
50
51
OC–Cu7O–O2 are given in Figures. 2l–2n as well. The adsorption sequence of small
52
53
54 molecules shows an obvious influence on the structure of the clusters, which further
55
56
induces different activations to the adsorbates. For example, The pre-adsorbed O2
57
58
59
60 10

ACS Paragon Plus Environment


Page 11 of 25 The Journal of Physical Chemistry

1
2
3
4 results in a significant distortion in Cu7O–O2, and further affects the subsequent attack
5
6 of CO in OC–Cu7O–O2 (2k), which is totally different in structure from the case of
7
8
9 CO direct adsorption on Cu7O (2l).
10
11 In summary, the adsorption of O2 and CO on the Cu6,7 and Cu6,7O clusters has the
12
13
14 following features: 1) the activation of O2 is mainly decided by the adsorption
15
16 patterns and partially by the cluster size and oxidation state and the subsequent attack
17
18
19 of CO. The activation of the O–O bonds follows the order of H-type > B-type >
20
21 T-type adsorption. 2) The incoming CO is always T-type adsorbed on the clusters and
22
23
24 can slightly improve the activation of O2 near it. The cluster size and oxidation states
25
26 have rather small influence on the activation of the C–O bond here. 3) The adsorbing
27
28
29 sequence of CO/O2 leads to different deformation of the clusters, and thus activates
30
31 the small molecules in different extents.
32
33
34
35
36
B. Competition between ER and LH mechanisms
37
38
39 Now we consider the kinetics and energetics of the reaction paths of the CO
40
41
oxidation on Cun and CunO, which are displayed in Figures. 3–7. First we explore CO
42
43
44 oxidation on pure Cu6,7 clusters following ER and LH mechanisms, respectively. As
45
46
discussed above, O2 pre-adsorbs on the pure Cu6,7 clusters. The rate-determining steps
47
48
49 for the ER mechanism are the activation of the pre-adsorbed O2 molecule with
50
51
barriers of 0.22 (TS1, 3c) and 0.42 eV (TS1, 4c) on the Cu6 and Cu7, respectively. For
52
53
54 the LH mechanism, the migrations of CO towards the pre-adsorbed O2 determine the
55
56
57
rate of the reactions with the barriers of 1.13 (TS1, 3g) and 0.89 eV (TS2, 4j) on the
58
59
60 11

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 12 of 25

1
2
3
4 Cu6 and Cu7, respectively. Apparently, ER is energetically preferred over LH for the
5
6 pure Cu6,7 clusters, maybe stemming from the high-degree activation of O2 upon the
7
8
9 H-type adsorption. The O–O bond lengthened by 28.1% in Cu6 and 24.8% in Cu7 with
10
11 respect to free O–O bond 1.21 Å, is active enough to oxidize the free CO via the ER
12
13
14 mechanism. Although the O2 is slightly further activated by CO in the LH path, the
15
16 CO migration is hindered by its relatively strong binding with Cun, resulting in higher
17
18
19 barriers than in the ER path.
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43 Figure. 3. Structures of initial states (IS), intermediates (IM), transition states (TS),
44
45 final states (FS) and relative energy (eV) with respect to IS in the paths which CO is
46
47
48 oxidized by molecular O2 on Cu6 via Eley-Rideal mechanism (ER(Cu6+O2+CO), red
49
50 dash arrow) and Langmuir-Hinshelwood mechanism (LH(Cu6+O2+CO), black solid
51
52
53 arrow).
54
55
56
57
58
59
60 12

ACS Paragon Plus Environment


Page 13 of 25 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25 Figure. 4. Structures of initial states (IS), intermediates (IM), transition states (TS),
26
27 final states (FS) and relative energy (eV) with respect to IS in the paths which CO is
28
29
oxidized by molecular O2 on Cu7 via Eley-Rideal mechanism (ER(Cu7+O2+CO), red
30
and dash arrow) and Langmuir-Hinshelwood mechanism (LH(Cu7+O2+CO), black
31
32 and solid arrow).
33
34
35
36
37 Next we turn to the ER-LH competition of CO oxidization on Cu6,7O clusters. From
38
39
40
above discussion, CO adsorbs on Cu6O first. In the ER pathway, it is the free O2 that
41
42 attacks the pre–adsorbed CO. Since the activation of CO is relatively weak (the C–O
43
44
45 bond is only 1.8% longer than in free CO), its oxidization faces an energy barrier of
46
47 1.13 eV (TS1, 5c). In the LH mechanism, in the presence of CO at the apex near the
48
49
50 intrinsic O atom, the incoming O2 attacks the farther hollow site, forming the
51
52 co–adsorbed OC–Cu6O–O2 cluster (5g). The oxidization does not occur unless CO
53
54
55 and/or O2 migrate towards each other. However, the pre-adsorbed O has some
56
57 restraint against the CO migration since OC–CunO constitutes a donor
58
59
60 13

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 14 of 25

1
2
3
4 (D)–Cun–acceptor (A) system in which CO(O) serves as the electron donor (acceptor).
5
6 As a result, the LH pathway is blocked for the CO + O2 reaction on Cu6O.
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22 Figure. 5. Structures of initial states (IS), intermediates (IM), transition states (TS),
23
24 final states (FS) and relative energy (eV) with respect to IS in the paths which CO is
25
26 oxidized by atomic O of Cu6O–CO–O2 in Langmuir-Hinshelwood mechanism
27
28 (LH(Cu6O+O2+CO), black solid arrow); and oxidized by molecular O2 via
29
30 Eley-Rideal mechanism (ER(Cu6O+O2+CO), red dash arrow).
31
32
33
34
35
Nevertheless, O2 adsorbs on Cu7O first and prefers the B–type pattern (6b). Its
36
37 O–O activation (lengthened by 16.5%) is less than those in Cu6,7O2 (28.1% and
38
39
40
24.8%). As a result, the ER mechanism has a relatively large barrier of 1.86 eV (TS1,
41
42 6c) for the formation of Cu7O–O2–CO precursor (IM2, 6d), which is the rate–limiting
43
44
45 step in the pathway. In the LH mechanism, after O2 adsorption, CO has the tendency
46
47 to adsorb at the sites near to O2 since O2 boosts the adsorption of CO by forming the
48
49
50 D–CuO–A system in which CO(O2) serves as the electron donor (acceptor). Such
51
52 cooperative adsorption effect helps the activation of both CO and O2. The LH path,
53
54
55 named LH2 in Figure. 6 is designed for this reaction, whose rate–limiting barriers is
56
57 1.04 eV (TS1, 6k), about 0.82 eV lower than the ER pathway. Therefore, LH– is
58
59
60 14

ACS Paragon Plus Environment


Page 15 of 25 The Journal of Physical Chemistry

1
2
3
4 preferred over ER–mechanism for the CO + O2 reaction on Cu7O. One needs to note
5
6 that although the adsorption energy of O2 on Cu7O (2.43 eV) is much larger than
7
8
9 those on the other three investigated clusters (1.01, 1.16 and 1.46 eV on Cu6,7, Cu6O
10
11 respectively), it has higher activation energy of 1.86 eV in the ER mechanism. The
12
13
14 adsorption energy has two origins, O2 activation and Cun reconstruction. Therefore the
15
16 large adsorption energy does not correspond solely to great O2 activation in some
17
18
19 cases. The large adsorption energy of O2 on Cu7O mainly stems from the
20
21 reconstruction of Cu7O cluster upon adsorption of O2 as shown in Figure 2i and 2j.
22
23
24 Thus it is not much helpful in the O-O activation. As discussed above, the activation
25
26 of the O–O bond follows the order of H-type > B-type > T-type adsorption. The O2
27
28
29 adsorption on Cu7O is B-type, which is less active than H-type adopted by the other
30
31 three clusters and leads to the high activation energies in the ER pathway of O2-Cu7O.
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 Figure. 6. Structures of initial states (IS), intermediates (IM), transition states (TS),
52
53 final states (FS) and relative energy (eV) with respect to IS in the paths which CO is
54
55 oxidized by molecular O2 on Cu7O via Eley-Rideal and Langmuir-Hinshelwood
56
57 mechanisms (ER(Cu7O+O2+CO), red dash arrow; and LH2(Cu7O+O2+CO), green dot
58
59
60 15

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 16 of 25

1
2
3 arrow), and oxidized by atomic O of Cu7O–O2–CO in Langmuir-Hinshelwood
4
5 mechanism (LH1(Cu7O+O2+CO), black solid arrow).
6
7
8
9
10 Furthermore, we assess the reactivity of the intrinsic O in the Cu6,7O clusters
11
12
13
towards CO. Neither the ER nor the LH mechanism is favored for these reactions,
14
15 as shown in Figures. 5-7. In the ER pathway, the intrinsic O atom in the oxide clusters
16
17
18
is inert because it already forms three O–Cu bonds with the neighboring Cu atoms. In
19
20 the LH pathway, the rate–determining barriers are as high as 2.65 eV (Cu6O) and 3.52
21
22
23 eV (Cu7O) without the participation of O2 (blue lines in Figures. 7b and 7d). In case
24
25 of CO-O2 co-adsorption, the CO oxidation by the intrinsic O atom also has high
26
27
28 barriers, 2.69 eV on Cu6O (TS1, 5h) and 1.32 eV on Cu7O (6g to 6i).
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49 Figure. 7. Potential energy surfaces of catalytic CO oxidation on Cu6,7 and Cu6,7O for
50
51
each reaction path. ER: Eley-Rideal mechanism, LH: Langmuir-Hinshelwood
52
mechanism.
53
54
55
56
57
Finally, by comparing the geometrical structures of IS, TS, IM and FS, we find that
58
59
60 16

ACS Paragon Plus Environment


Page 17 of 25 The Journal of Physical Chemistry

1
2
3
4 the reactions have lower energy barriers when the TS structures are closer to the
5
6 reactant structures, but have higher barriers when the TS structures are closer to the
7
8
9 product structures. This is similar to the concept of “early/late transition state”
10
11 (namely the Hammond–Leffler postulate, early transition states with TS structures are
12
13
14 closer to the reactants; late transition states with TS structures are closer to the
15
16 products) in organic reactions. For example, in the reactions of Cu6 + O2 + CO →
17
18
19 Cu6O + CO2 via the ER mechanism, the TS structure 3c is closer to the reactant 3b
20
21 than to the product 3d, corresponding to a lower barrier of 0.22 eV; in contrast, in the
22
23
24 LH pathway, the TS structure 3g is closer to the product 3h than the intermediate 3f,
25
26 corresponding to a high barrier of 1.13 eV.
27
28
29 The competition between ER and LH mechanisms is common in surface reactions.
30
31 For example, the ER-LH competition has also been studied for reactions on Au55, W10
32
33
34 clusters and W(111) surface, copper-embedded graphene and Cu2O(111) surface,41
35
36 although the driving force for these two pathways has not yet been fully addressed. In
37
38
39 this study, we found that the step of O2 activation is crucial for the ER or LH
40
41
preference of the surface reactions on metal clusters or solids. Since CO usually acts
42
43
44 as an electron donor while O2 as an acceptor, the activation of CO is very difficult on
45
46
a metal system since both the metal and CO tend to donate electrons during
47
48
49 interaction. In contrast, O2 can accept one electron from the metal into its
50
51
anti-bonding π* orbital, boost the CO adsorption, and become activated. If O2 itself is
52
53
54 activated enough after adsorption, it may oxidize CO directly through an ER pathway.
55
56
57
If O2 activation is not enough without further activation by the cooperative adsorption
58
59
60 17

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 18 of 25

1
2
3
4 of CO, an LH pathway is preferred. Since the adsorption on a cluster has a variety of
5
6 patterns with quite different adsorption energies, the O2 activation and the subsequent
7
8
9 ER-LH competition become complicated for reactions on the cluster. Now we extend
10
11 our findings to some other systems. Gong et al.41 have found that LH is more
12
13
14 favorable than ER for CO oxidization on Au55. CO and O2 have adsorption energy of
15
16 0.50 and 0.08 eV on Au55, respectively. In this case, O2 adsorbs subsequently on Au55
17
18
19 after CO adsorption instead of reaction with CO directly. The co-adsorption of CO
20
21 and O2 yields not only larger adsorption energy of 0.83 eV but also an activated O2
22
23
24 near to CO, making the LH the preferred mechanism for CO oxidization on Au55.
25
26 Differently, O2 is activated greatly on W10 clusters and W(111) surface.42 As a result,
27
28
29 CO reacts directly with the activated O2 instead of adsorption on the surface, leading
30
31 to the ER preference of this reaction. It suggests that the activation of O2 determines
32
33
34 the priority of ER or LH mechanism. The degree of O2 activation is much less on the
35
36 copper-embedded graphene, in which the O–O bond is elongated by 11.6%, than that
37
38
39 on Cu6, and Cu7 in which O-O is lengthened by 24% or more. Therefore, the O2
40
41
prefers reacting with CO via LH mechanism on the copper-embedded graphene.43
42
43
44 An ER mechanism was suggested for CO oxidization on Cu2O(111) surface since the
45
46
high O concentration makes the LH mechanism infeasible because the adsorbed CO a
47
48
49 nd O2 are distant from each other on the surface44.
50
51
52
53
54 IV. CONCLUSION
55
56
57
We have performed systematic DFT calculations to study the competition between
58
59
60 18

ACS Paragon Plus Environment


Page 19 of 25 The Journal of Physical Chemistry

1
2
3
4 the ER and LH mechanism for the CO oxidation reactions on Cu6, Cu7, Cu6O, and
5
6 Cu7O clusters. Our calculations reveal that 1) CO and O2 exhibit quite different
7
8
9 adsorption behaviors (sites, patterns, and energetics) on the cluster surfaces, which
10
11 have great influence on the subsequent surface reactions; 2) The ER- is preferred over
12
13
14 LH-mechanism for Cu6, Cu7, and Cu6O, while Cu7O favors the LH-mechanism; 3)
15
16 The step of O2 activation is crucial for the ER or LH preference of the surface
17
18
19 reactions on metal clusters or solids. It is conjectured if O2 itself is activated enough
20
21 after adsorption, it may oxidize CO directly through an ER pathway; if O2 activation
22
23
24 is not enough without the cooperative adsorption of CO, a LH pathway is preferred. In
25
26 addition, the concept of “early/late transition state” in organic reactions is confirmed
27
28
29 in the CO oxidization on Cu6,7 and Cu6,7O surfaces. It is worthy to mention that in this
30
31 work both the Cun and CunO clusters were treated in gas phase. Their adsorption and
32
33
34 reactivity may be different on metal oxide substrates. This is an interesting issue for
35
36 future study.
37
38
39
40
41
ACKNOWLEDGEMENTS
42
43
44 This work is supported by NBRP (2011CB302004, 2010CB923401 and
45
46
2009CB623200), the NSF (21173040 and 11074035) and Peiyu Foundation of SEU.
47
48
49 MY Thanks the Open research Fund of State Key Laboratory of Oil and Gas
50
51
Reservoir Geology and Exploration (No. PLN1118). The authors thank the
52
53
54 computational resource at Department of Physics, SEU and National Supercomputing
55
56
57
Center in Tianjin.
58
59
60 19

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 20 of 25

1
2
3
4
5
6 References
7
8
(1) Kugai, J.; Kitagawa, R.; Seino, S.; Nakagawa, T.; Ohkubo, Y.; Nitani, H.; Daimon,
9
10 H.; Yamamoto, T. A. Appl. Catal. A: Gen. 2011, 406, 43.
11
12 (2) Wu, G.; Guan, N.; Li, L. Catal. Sci. Technol. 2011, 1, 601.
13
14 (3) Abe, Y.; Yamaura, H.; Yamaguchi, S.; Yahiro, H. Catal. Commun. 2010, 11, 820.
15
16 (4) Yu, Q.; Wu, X.; Yao, X.; Liu, B.; Gao, F.; Wang, J.; Dong, L. Catal. Commun.
17
18 2011, 12, 1311.
19
20 (5) Shen, Y.; Lu, G.; Guo, Y.; Wang, Y.; Guo, Y.; Gong, X. Catal. Today 2011, 175,
21
22 558.
23
24
(6) Pasini, T.; Piccinini, M.; Blosi, M.; Bonelli, R.; Albonetti, S.; Dimitratos, N.;
25 Lopez–Sanchez, J. A.; Sankar, M.; He, Q.; Kiely, C. J.; Hutchings, G. J.; Cavania, F.
26
27 Green Chem. 2011, 13, 2091.
28
29 (7) Chen, C. S.; Lai, T. W.; Chen, C. C. J. Catal. 2010, 273, 18.
30
31 (8) Kappen, P.; Grunwaldt, J. D.; Hammershøi, B. S.; Tröger, L.; Clausen, B. S. J.
32
33 Catal. 2001, 198, 56.
34
35 (9) Chen, J.; Zhan, Y.; Zhu, J.; Chen, C.; Lin, X.; Zheng, Q. Appl. Catal. A: Gen.
36
37 2010, 377, 121.
38
39 (10) DekkeP, F. H. M.; Dekker, M. C.; Bliek, A.; Kapteijn, F.; Moulijn, J. A. Catal.
40
Today 1994, 20, 409.
41
42 (11) Dekker, F. H. M.; Kraneveld, S.; Bliek, A.; Kapteijn, F.; Moulijn, J. A. J. Catal.
43
44 1997, 170, 168.
45
46 (12) Ayastuy, J. L.; Gurbani, A.; González–Marcos, M. P.; Gutiérrez–Ortiz, M. A. Int.
47
48 J. Hydrogen Energ. 2010, 35, 1232.
49
50 (13) (a) White, B.; Yin, M.; Hall, A.; Le, D.; Stolbov, S.; Rahman, T.; Turro, N.;
51
52 O’Brien, S. Nano Lett. 2006, 6, 2095. (b) Kosmambetova, G. R.; Moroz, E. M.;
53
54 Guralsky, A. V.; Pakharukova, V. P.; Boronin, A. I.; Ivashchenko, T. S.; Gritsenko, V.
55
56 I.; Strizhak, P. E. Int. J. Hydrogen Energ. 2011, 36, 1271.
57
(14) Gokhale, A. A.; Dumesic, J. A.; Mavrikakis, M. J. Am. Chem. Soc. 2008, 130,
58
59
60 20

ACS Paragon Plus Environment


Page 21 of 25 The Journal of Physical Chemistry

1
2
3 1402.
4
5 (15) Estrella, M.; Barrio, L.; Zhou, G.; Wang, X.; Wang, Q.; Wen, W.; Hanson, J. C.;
6
7 Frenkel, A. I.; Rodriguez, J. A. J. Phys. Chem. C 2009, 113, 14411.
8
9 (16) Bridier, B.; Hevia, M. A. G.; López, N.; Pérez–Ramírez, J. J. Catal. 2011, 278,
10
11 167.
12
13 (17) Lapidus, A. L.; Gaidai, N. A.; Nekrasov, N. V.; Tishkova, L. A.; Agafonov, Y. A.;
14
15 Myshenkova, T. N. Petr. Chem. 2007, 47, 75.
16
17 (18) Padilla–Campos, L. J. Molec. Struct. (Theochem.) 2008, 851, 15.
18
19
(19) Cao, Z.; Wang, Y.; Zhu, J.; Wu, W.; Zhang, Q. J. Phys. Chem. B 2002, 106, 9649.
20 (20) Poater, A.; Duran, M.; Jaque, P.; Toro–Labbé, A.; Solà, M. J. Phys. Chem. B 2006,
21
22 110, 6526.
23
24 (21) Holmgren, L.; Grönbeck, H.; Andersson, M.; Rosén, A. Phys. Rev. B 1996, 53,
25
26 16644.
27
28 (22) Becke, A.D. Phys. Rev. A 1988, 38, 3098.
29
30 (23) Lee, C.; Yang, W.; Parr, R.G. Phys. Rev. B 1988, 37, 785.
31
32 (24) Florez, E.; Tiznado, W.; Mondragón, F.; Fuentealba, P. J. Phys. Chem. A 2005,
33
34 109, 7815.
35
(25) Guzmán-Ramírez, G.; Aguilera-Granja, F.; Robles, J. Eur. Phys. J. D 2010, 57,
36
37 49.
38
39 (26) Guzmán-Ramírez, G.; Aguilera-Granja, F.; Robles, J. Eur. Phys. J. D 2010, 57,
40
41 335.
42
43 (27) Yuan, X.; Liu, L.; Wang, X.; Yang, M.; Jackson, K. A. J. Phys. Chem. A 2011,
44
45 115, 8705.
46
47 (28) Padilla–Campos, L.; J. Molec. Struct. (Theochem.) 2007, 815, 63.
48
49 (29) White, J. A.; Bird, D. M.; Payne, M. C.; Stich, I. Phys. Rev. Lett. 1994, 73, 1404.
50
51
(30) Liu, W.; Zhao, Y. H.; Lavernia, E. J.; Jiang, Q. J. Phys. Chem. C 2008, 112, 7672.
52
(31) Lu, Z. H.; Xu, Q. Phys. Chem. Chem. Phys. 2010, 12, 7077.
53
54 (32) Zhao, S.; Ren, Y. L.; Wang, J. J.; Yin, W. P. J. Phys. Chem. A 2009, 113, 1075.
55
56 (33) Chen, L.; Zhang, Q.; Zhang, Y.; Li, W. Z.; Han, B.; Zhou, C.; Wu, J.; Forrey, R.
57
58 C.; Garg, D.; Cheng, H. Phys. Chem. Chem. Phys. 2010, 12, 9845.
59
60 21

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 22 of 25

1
2
3
(34) Liu, W.; Lian, J. S.; Jiang, Q. J. Phys. Chem. C 2007, 111, 18189.
4
5 (35) Park, P. W.; Ledford, J. S. Appl. Catal. B: Eevi. 1998, 15, 221.
6
7 (36) Boreskov, G. K. Heterogeneous Catalysis (Russian) Nauka, Moscow (1988).
8
9 (37) Pinchuk, V. M.; Kotlyarova, E. S.; Parkhomenko, N. V.; Tsybulev, P. N. J. Struct.
10
11 Chem. 1996, 37, 544.
12
13 (38) (a) Gao, Y.; Shao, N.; Pei, Y.; Zeng, X. C. Nano Lett. 2010, 10, 1055. (b)
14
15 Lopez-Acevedo, O.; Kacprzak, K. A.; Akola, J.; Häkkinen, H. Nat. Chem. 2010, 2,
16
17 329.
18
19
(39) (a) Beusch, H.; Fieguth, P.; Wicke, E. Chem. Ing. Tech. 1972, 44, 445. (b) Hori, G.
20 K.; Schmidt, L. D. J. Catal. 1975, 38, 335. (c) Hugo, P.; Jakubith, M. Chem. lng.
21
22 Tech. 1972, 44, 383. (d) Dauchot, J. P.; Cakenberghe, J. V. Nat. Phys. Sci. 1973, 249,
23
24 61.
25
26 (40) (a) Ivanov, V. P.; Boreskov, G. K.; Savchenko, V. I.; Egelhoff, W. F. Jr.; Weinberg,
27
28 W. H. J. Catal. 1977, 48, 269. (b) Taylor, J. L.; Ibbotson, D. E.; Weinberg, W. H. Surf.
29
30 Sci. 1979, 79, 349. (c) Taylor, J. L.; Ibbotson, D. E.; Weinberg, W. H. J. Catal. 1980,
31
32 80, 1. (d) Falconer, J. L.; Wentrick, P. R.; Wise, H. J. Catal. 1976, 45, 248.
33
34 (41) Xie, Y. P.; Gong, X. G. J. Chem. Phys. 2010, 132, 244302.
35
(42) Weng, M. H.; Ju, S. P. J. Phys. Chem. C 2012, 116, 18803.
36
37 (43) Song, E. H.; Wen, Z.; Jiang, Q. J. Phys. Chem. C 2011, 115, 3678.
38
39 (44) Sun, B. Z.; Chen, W. K.; Xu, Y. J. J. Chem. Phys. 2010, 133, 154502._
40
41 (45) (a) Cox, D. M.; Brickman, R. O.; Greegan, K.; Kaldor, A. Z. Z. Phys. D: At., Mol.
42
43 Clusters 1991, 19, 353. (b) Lee, T. H.; Ervin, K. M. J. Phys. Chem. 1994, 98, 10023.
44
45 (c) Salisbury, B. E.; Wallace, W. T.; Whetten, R. L. Chem. Phys. 2000, 262, 131
46
47 (2000).
48
49 (46) (a) Zemski, K. A.; Justes, D. R.; Castleman, A. W. Jr.; J. Phys. Chem. B 2002,
50
51
106, 6136. (b) Johnson, G. E.; Tyo, E. C.; Castleman, A. W. Jr.; Proc. Natl. Acad. Sci.
52
2008, 105, 18108. (c) Pal, R. Wang, L. M.; Pei, Y.; Wang, L. S.; Zeng, X. C. J. Am.
53
54 Chem. Soc. 2012, 134, 9438.
55
56 (47) DMOL is a density functional theory program distributed by Accelrys, Inc.;
57
58 Delley, B. J. Chem. Phys. 1990, 92, 508.
59
60 22

ACS Paragon Plus Environment


Page 23 of 25 The Journal of Physical Chemistry

1
2
3
(48) Delley, B. J. Chem. Phys. 2000, 113, 7756.
4
5 (49) Becke, A. D. J. Chem. Phys. 1993, 98, 5648.
6
7 (50) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B 1988, 37, 785.
8
9 (51) Xie, Z.; Ma, Q. M.; Liu, Y.; Li, Y. C. Phys. Lett. A 2005, 342, 459.
10
11 (52) Ma, Q. M.; Xie, Z.; Wang, J.; Liu, Y.; Li, Y. C. Solid State Commun. 2007, 142,
12
13 114.
14
15 (53) Ma, Q. M.; Liu, Y.; Xie, Z.; Wang, J. J. Phys. 2006, 29, 163.
16
17 (54) Zhang, D. B.; Shen, J. J. Chem. Phys. 2004, 120, 5104.
18
19
(55) Liu, N.; Ma, Q. M.; Xie, Z.; Liu, Y.; Li, Y. C. Chem. Phys. Lett. 2007, 436, 184.
20 (56) Delley, B. Phys. Rev. B 2002, 66, 155125.
21
22 (57) Dai, M.; Wang, Y.; Kwon, J.; Halls, M. D.; Chabal, Y. J. Nat. Mater. 2009, 8,
23
24 825.
25
26 (58) Ruscic, B.; Wagner, A. F.; Harding, L. B.; Asher, R. L.; Feller, D.; Dixon, D. A.;
27
28 Peterson, K. A.; Song, Y.; Qian, X.; Ng, J.; Liu, C. Y.; Chen, W.; Schwenke, D. W. J.
29
30 Phys. Chem. A 2002, 106, 2727.
31
32 (59) Feyereisen, M. W.; Feller, D.; Dixon, D. A. J. Phys. Chem. 1996, 100, 2993.
33
34 (60) Watson, L. R.; Thiem, T. L.; Dressler, R. A.; Salter, R. H.; Murad, E. J. Phys.
35
Chem. 1993, 97, 5577.
36
37 (61) Fisher, E. R.; Armentrout, P. B. J. Phys. Chem. 1990, 94, 1674.
38
39 (62) Feng, C. J.; Zhang, X. Y. Commun. Theor. Phys. 2009, 52, 675.
40
41 (63) NIST Standard Reference Database 69, March 2003 Release, NIST Chemistry
42
43 WebBook, http://webbook.nist.gov.
44
45 (64) Knight, H. T.; Rink, J. P. J. Chem. Phys. 1958, 29, 449.
46
47 (65) NIST Standard Reference Database 101, Sept. 2006 Release, Computational
48
49 Chemistry Comparison and Benchmark DataBase, http://cccbdb.nist.gov.
50
51
(66) Brix, P.; Herzberg, G. Canadian Journal of Physics, 1954, 32, 110.
52
(67) Merer, A. J. Annu. Rev. Phys. Chem. 1989, 40, 407.
53
54 (68) Pedley, J. B.; Marshall, E. M. J. Phys. Chem. Ref. Data 1983, 12, 967.
55
56 (69) K. P. Huberg, and G. Herzberg, Molecular Spectra and Molecular Structure–IV,
57
58 Van Nostrand–Reinhold: New York, 1989.
59
60 23

ACS Paragon Plus Environment


The Journal of Physical Chemistry Page 24 of 25

1
2
3 (70) Morse, M. D.; Hansen, G. P.; Langridge-Smith, P. R. R.; Zheng, L. S.; Geusic, M.
4
5 E.; Michalopoulos, D. L.; Smalley, R. E. J. Chem. Phys. 1984, 80, 5400.
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 24

ACS Paragon Plus Environment


Page 25 of 25 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 TABLE OF CONTENTS: Schematic diagram of competition between Eley-Rideal
22
23
24 and Langmuir-Hinshelwood pathways for CO oxidation reaction on Cun clusters.
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 25

ACS Paragon Plus Environment

You might also like