You are on page 1of 28

Subscriber access provided by Eastern Michigan University | Bruce T.

Halle Library

Article
Selective Hydrogenation of Acetylene to Ethylene in the Presence of
a Carbonaceous Surface Layer on a Pd/Cu(111) Single-Atom Alloy
Christopher M Kruppe, Joel D Krooswyk, and Michael Trenary
ACS Catal., Just Accepted Manuscript DOI: 10.1021/acscatal.7b02862 Publication Date (Web): 16 Oct 2017
Downloaded from http://pubs.acs.org on October 16, 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 27 ACS Catalysis

1
2
3
4
5
Selective Hydrogenation of Acetylene to Ethylene in
6
7
8
9 the Presence of a Carbonaceous Surface Layer on a
10
11
12
13 Pd/Cu(111) Single-Atom Alloy
14
15
16
17
18
Christopher M. Kruppe, Joel D. Krooswyk, Michael Trenary*
19
20
21 Department of Chemistry, University of Illinois at Chicago, 845 West Taylor Street, Chicago,
22
23 Illinois 60607
24
25
26
27 *mtrenary@uic.edu
28
29
30 ABSTRACT
31
32
33
Reflection absorption infrared spectroscopy (RAIRS) was used to simultaneously monitor gas
34
35
36 phase and surface species in the presence of ambient pressures of acetylene and hydrogen over a
37
38 single-atom alloy (SAA). The alloy consisted of isolated Pd atoms at surface coverages in the
39
40
41
range 0.0028 to 0.085 ML in a Cu(111) surface. When C2H2(g) is present, but not H2(g), the
42
43 RAIR spectra are similar for Cu(111) with and without Pd, independent of C2H2(g) pressure for
44
45 temperatures between 180 and 500 K. The addition of H2(g) leads to different RAIR spectra
46
47
48 depending on the presence of Pd. With a C2H2:H2 ratio of 1:100 and a SAA-Pd/Cu(111) surface
49
50 with less than 1% Pd, C2H2(g) is converted to C2H4(g) at 400 K at total pressures up to 10 Torr.
51
52 From the rate of change in the gas phase IR peaks, a range of initial turnover frequencies was
53
54
55 estimated, which depend on which sites are assumed to be active for hydrogenation. Post-
56
57 reaction surface analysis with Auger electron spectroscopy (AES) showed a significant carbon
58
59
60 1
ACS Paragon Plus Environment
ACS Catalysis Page 2 of 27

1
2
3
coverage, which decreased with increasing Pd coverage. The combined RAIRS and AES results
4
5
6 suggest that by increasing reactivity for ethylene formation, Pd also limits the amount of carbon
7
8 that is deposited, while also changing the extent of oligomer formation.
9
10
11
12 Keywords: Single-atom alloy; acetylene selective hydrogenation; Pd/Cu(111); reflection
13
14 absorption infrared spectroscopy; turnover frequency
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 2
ACS Paragon Plus Environment
Page 3 of 27 ACS Catalysis

1
2
3
4
5
1. INTRODUCTION
6
7
8
9 The selective hydrogenation of alkynes to alkenes is an important area of heterogeneous
10
11 catalysis. For example, conversion of acetylene to ethylene is used to remove small quantities of
12
13
14
acetylene from the ethylene feedstocks used for the production of polyethylene.1-2 The challenge
15
16 is to achieve a high reactivity for alkyne hydrogenation in the presence of a large excess of
17
18 alkene without also hydrogenating the alkene to the alkane. This goal is usually achieved through
19
20
21 the use of bimetallic catalysts that contain Pd, with the exact composition designed to retain high
22
23 activity while increasing selectivity. Effective catalysts have even been reported for Pd combined
24
25 with non-metals such as sulfur.3 Among the many metals used to form Pd-containing bimetallic
26
27
28 catalysts, Cu is particularly well studied.2, 4 In general, in any catalyst it is desirable to minimize
29
30 the use of expensive metals, such as Pd, that are needed to achieve a particular catalytic
31
32
objective. An especially efficient use of Pd is in the form of single-atom alloys (SAAs), a term
33
34
35 first used by Sykes and coworkers to describe Cu(111) surfaces in which the Pd is present as
36
37 isolated atoms that replace Cu atoms in the topmost surface layer.5 Other studies have shown that
38
39
40
the Pd atoms are not distributed evenly on the surface but are concentrated near the top of
41
42 Cu{111} steps.6-8 It was found that the Pd promoted H2 dissociation with spillover of H atoms to
43
44 nearby Cu sites.5 Using DFT calculations of various configurations of surface and subsurface Pd
45
46
47 in Cu(111), Fu and Luo argued that the lowest activation energy for H2 dissociation was
48
49 achieved when a Pd surface atom was located above a subsurface Pd atom.9 Kyriakou et al. used
50
51 temperature programmed desorption (TPD) under ultrahigh vacuum conditions to show that a
52
53
54 Cu(111) surface with 0.01 monolayer (ML) of Pd could selectively hydrogenate acetylene to
55
56 ethylene.5 Ma et al. used DFT calculations to show that a Pd/Cu(111) SAA achieves selectivity
57
58
59
60 3
ACS Paragon Plus Environment
ACS Catalysis Page 4 of 27

1
2
3
in acetylene hydrogenation mainly by lowering the barrier to ethylene desorption.10 Supported
4
5
6 Pd/Cu SAA catalysts have been shown to be effective for the partial hydrogenation of
7
8 acetylene11-13 and propyne.14 It has also been found that the addition of CO to the feed gas can
9
10
11
enhance the selectivity of Pd/Cu catalysts.15 Other studies have shown that the addition of CO
12
13 can alter the structure of Cu catalysts that dont contain Pd and thereby enhance alkene
14
15 selectivity.16-17 Theoretical calculations of SAAs consisting of Pt, Pd, Ni, and Rh in a Cu host
16
17
18 indicate that Pd/Cu and Pt/Cu should be stable as acetylene hydrogenation catalysts, whereas
19
20 Ni/Cu and Rh/Cu are not.18 Here we have used reflection absorption infrared spectroscopy
21
22 (RAIRS) to monitor acetylene hydrogenation over a Pd/Cu(111) SAA under ambient pressures
23
24
25 of acetylene and hydrogen. The results reveal that while acetylene is selectively hydrogenated to
26
27 ethylene over Pd/Cu(111) but not over Pd-free Cu(111), the elevated acetylene pressures lead to
28
29
coupling products not observed with RAIRS under UHV conditions and that these products are
30
31
32 different on the Cu(111) and Pd/Cu(111) surfaces. The elevated pressure conditions more closely
33
34 mimic those of practical hydrogenation catalysis and add new insights into the surface species
35
36
37 present under actual acetylene hydrogenation conditions.
38
39
40 In addition to the reaction that is desired, namely hydrogenation of acetylene to ethylene,
41
42 acetylene is known to undergo other reactions on both Cu and Pd surfaces. In particular,
43
44
45 trimerization to form benzene is a structure sensitive reaction that occurs on several surfaces.19
46
47 On Cu(111), Kyriakou et al., used TPD to observe the desorption from Cu(111) of coupling
48
49
products including benzene and cyclooctatetraene under UHV conditions.20 They observed
50
51
52 similar products under atmospheric conditions for a catalyst consisting of Cu nanoparticles
53
54 supported on alumina.20 In their experiments they only observed desorption products, although
55
56
57
they inferred that non-desorbing oligomers also formed on the surface. It has been proposed that
58
59
60 4
ACS Paragon Plus Environment
Page 5 of 27 ACS Catalysis

1
2
3
trimerization occurs via a C4H4 metallacyle intermediate.21 In addition to benzene formation on
4
5
6 the Pd(111) surface, when hydrogen is present acetylene is readily hydrogenated to ethylene.22-24
7
8 Under UHV conditions, the acetylene isomerizes to vinylidene and can be hydrogenated to the
9
10
11
spectator species ethylidyne, with the latter yielding a characteristic RAIR spectrum.25
12
13 Ethylidyne was readily detected on Pd(111) with RAIRS while benzene was being formed in the
14
15 presence of 5 Torr of acetylene.26 The formation of oligomers over practical Pd-based acetylene
16
17
18 hydrogenation catalysts is well-known.27 These oligomers are thought to be intermediates to the
19
20 formation of so-called green oil, a product that is associated with the deactivation of the
21
22 catalyst.28
23
24
25
26 As the coupling reactions to form carbonaceous surface species compete with acetylene
27
28 hydrogenation, it is clearly desirable to perform studies over model catalysts in which the
29
30
31
hydrogenation activity and the formation of carbonaceous species can be simultaneously
32
33 monitored. In a recent study of acetylene hydrogenation over a Pt(111) surface, we used
34
35 polarization-dependent RAIRS to monitor the conversion of gas-phase acetylene first to C2H4(g)
36
37
38 and then to C2H6(g) while also monitoring the formation of various surface species, the most
39
40 prominent of which was ethylidyne (CCH3).29 We use a similar approach here to observe that
41
42 C2H2(g) is converted to C2H4(g) but not to C2H6(g) while also observing the formation of
43
44
45 surface-bound carbonaceous species. Through comparisons with IR spectra in the literature, we
46
47 conclude that whereas polyacetylene forms on Pd-free Cu(111) under hydrogenation conditions,
48
49
the presence of Pd atoms in the Pd/Cu(111) SAA promotes the formation of a surface
50
51
52 carbonaceous species similar in structure to green oil.
53
54
55 2. EXPERIMENTAL SECTION
56
57
58
59
60 5
ACS Paragon Plus Environment
ACS Catalysis Page 6 of 27

1
2
3
All experiments were performed in a UHV system that houses an upper analysis chamber
4
5
6 and a lower ambient pressure IR cell. The upper chamber is equipped with a LK Technologies
7
8 RVL2000 Auger-LEED system, Pfeiffer Prisma quadrupole mass spectrometer, and a Perkin-
9
10
11
Elmer ion sputter gun. The sample is transferred into the IR cell through spring-loaded Teflon
12
13 seals that isolate the main chamber during the ambient pressure experiments. The upper analysis
14
15 chamber is kept at 110-9 Torr during the ambient pressure experiments with an ion pump. The
16
17
18 IR experiments were performed with a Bruker Vertex 70v FTIR. The polarization-dependent
19
20 reflection absorption infrared spectroscopy (PD-RAIRS) setup allows for acquisition of surface
21
22 and gas phase species through the use of a rotatable polarizer as described earlier.30
23
24
25
26 The Cu(111) crystal (99.9999%, Princeton Scientific) was cleaned by Ar ion sputtering (1
27
28 keV, 10 A) and annealing (950 K). The surface cleanliness was confirmed with AES. Pd was
29
30
31
deposited on the Cu(111) crystal at 380 K from a resistive evaporator consisting of a thin Pd wire
32
33 (99.99%, Alpha Aesar) wrapped around a W wire. The deposition rate was ~0.006 ML/min
34
35 allowing for low coverages of Pd to be prepared. Surface coverages of Pd have been determined
36
37
38 by PD-RAIRS measurements of CO as described earlier.31 Surface Pd coverages determined this
39
40 way were found to be lower than Pd coverages determined from AES by a factor of 0.28 due to
41
42 the detection of both surface and subsurface Pd with AES whereas CO only adsorbs on surface
43
44
45 Pd atoms. This relation was used to determine surface Pd coverages from post reaction AES
46
47 measurements to avoid using CO. The Pd deposition rate was reproducible so that it was not
48
49
necessary to verify the Pd coverage after each deposition with AES, which minimized carbon
50
51
52 contamination from electron-beam induced dissociation.
53
54
55 RAIRS experiments were performed with 1024 scans at 4 cm-1 resolution, unless
56
57
58
otherwise noted. For most experiments the sample was held at a given temperature, exposed to
59
60 6
ACS Paragon Plus Environment
Page 7 of 27 ACS Catalysis

1
2
3
the reactant gases, annealed to a specific temperature, and cooled to the starting temperature
4
5
6 where a spectrum was obtained. For experiments performed at low temperatures, a UHV
7
8 exposure of C2H2(g) in fixed Langmuir (L) units (1 L = 110-6 Torr s) was made followed by
9
10
11
admitting a static pressure of H2(g) into the IR cell. The ambient pressure experiments starting at
12
13 300 K were done with gas mixtures of C2H2(g) and H2(g). The TOF experiment was done by
14
15 holding the crystal temperature at 400 K, admitting the gas phase reactants into the IR cell, and
16
17
18 taking IR scans as a function of time. Atomic absorption grade acetylene (99.6%) and hydrogen
19
20 (99.999%) was purchased from Praxair. The acetylene was further purified by the freeze-pump
21
22 thaw method and all gases were checked by mass spectrometry. The hydrogen was exposed
23
24
25 through a liquid N2 trap to remove any residual contaminants.
26
27
28 3. RESULTS and DISCUSSION
29
30
31
32
3.1 Low coverage, low temperature C2H2 adsorption and hydrogenation on Cu(111) and
33
34 Pd/Cu(111)
35
36
37 Infrared spectra of 2 L of C2H2 exposed to Pd-free Cu(111) and to 0.085 ML Pd/Cu(111)
38
39
40 at 150 and 180 K, respectively, are displayed in Figure 1. The spectra agree with earlier
41
42 vibrational studies of C2H2 exposed to Cu(111) at low temperatures and we assume that the
43
44 molecule bonds with the structure previously proposed.32-33 The identical peak positions in the
45
46
47 two cases indicate that the Pd atoms do not affect the internal bonding of C2H2 interacting with
48
49 the Cu(111) surface. This could be due to Pd atoms adsorbed near the step sites, while C2H2 is
50
51
52
mainly bound to the Cu(111) terraces. There was no significant change in the spectra for the two
53
54 surfaces after 2102 Torr of H2(g) was added to the cell with the crystal at 180 K. However,
55
56 annealing the surfaces to 280 K showed a significant decrease in the acetylene peaks for the
57
58
59
60 7
ACS Paragon Plus Environment
ACS Catalysis Page 8 of 27

1
2
3
Pd/Cu(111) surface whereas there was little change for the Cu(111) surface, even after annealing
4
5
6 to 360 K. In contrast, for the Pd/Cu(111) surface, the acetylene peaks are essentially gone after
7
8 annealing to 300 K. Other results, not shown, show little difference between Pd-free Cu(111) and
9
10
11
Pd/Cu(111) for annealing to 280 K in the absence of H2(g). These results are readily explained
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31 Figure 1. RAIR spectra obtained while monitoring C2H2 hydrogenation at low temperature. 2 L of
32 C2H2 was exposed to Cu(111) at 150 K (A) and to 0.085 ML Pd/Cu(111) at 180 K (B). 210-2 Torr of
33 H2 was then added to the IR cell and the samples annealed at the indicated temperatures for 30 s and a
34 spectrum was taken upon cooling to 180 K.
35
36 by removal of adsorbed acetylene through hydrogenation, which occurs to a much greater extent
37
38 in the presence of Pd atoms on the surface. When an ambient pressure of H2(g) is not present,
39
40
41
Kyriakou et al. used temperature programmed reaction (TPR) to show that acetylene adsorbed
42
43 under UHV conditions is removed from the surface by desorption and by coupling reactions to
44
45 form benzene and cyclooctatetraene, along with some hydrogenation to ethylene and butadiene,
46
47
48 all of which desorb in the range of 280 to 350 K.20 After the TPR sweeps, no carbon was
49
50 detected by XPS.20
51
52
53 3.2 C2H2 coupling at 110-2 Torr
54
55
56
57
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 27 ACS Catalysis

1
2
3
The peaks observed in the RAIR spectra in Figure 1 following exposure to C2H2(g) in UHV
4
5
6 are all assigned to adsorbed acetylene. As the surfaces were annealed, no new peaks were
7
8 observed that would be indicative of the formation of additional surface species. Distinctly
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 Figure 2. Comparison of ambient pressure C2H2 coupling over clean Cu(111) (blue dotted line),
37 0.0028 ML Pd/Cu(111) (green), and 0.080 ML Pd/Cu(111) (red). Spectra were acquired in the
38 presence of 110-2 Torr of C2H2 at 300 K then annealing to 480 K. Green numbers denote gas phase
39 peaks.
40
41
42
43 different results were observed when the surfaces were annealed in the presence of gas phase
44
45 C2H2(g). Figure 2 shows spectra for Pd-free Cu(111) as well as for 0.0028 and 0.080 ML
46
47
48 Pd/Cu(111) at 300 K and after annealing to 380 and 480 K in the presence of 110-2 Torr of
49
50 C2H2(g). At 300 K, the spectra for the three surfaces match almost perfectly and display the same
51
52 peaks as observed in Figure 1 under UHV conditions. However, upon annealing to 380 K, new
53
54
55 peaks appear with different intensities for the three surfaces. Gas phase acetylene peaks are also
56
57 present (identified with green numbers) but do not change with annealing temperature and have
58
59
60 9
ACS Paragon Plus Environment
ACS Catalysis Page 10 of 27

1
2
3
the same intensity for the three surfaces. For the 380 K anneal, the most prominent new peak is
4
5
6 at 1006 cm1, which is a good match to polyacetylene.34 In particular, for polyacetylene prepared
7
8 at 150 C, a peak at 1015 cm-1, assigned to an out-of-plane CH bending mode, is by far the
9
10
11
most intense peak in the IR spectrum.34 Also present in Figure 2 is a derivative-shaped peak with
12
13 a negative lobe at 1044 cm-1. Such shapes are typically observed with RAIRS when a peak
14
15 present in the background spectrum is shifted as more of the adsorbate responsible for the peak is
16
17
18 added to the surface. Since a peak at 1044 cm1 has previously been assigned to H atoms
19
20 adsorbed in the three-fold hollow sites of Cu(111),35-36 the results here imply that there was
21
22 already hydrogen inadvertently present when the background spectrum was taken. Whether
23
24
25 hydrogen was initially present or not depends on the exact conditions used so there is
26
27 considerable variability in this region from one experiment to the next. For example, this feature
28
29
is even more prominent in Figure 3. The other peaks observed in the 380 K spectrum in Figure 2
30
31
32 are not assigned to polyacetylene for the following reasons. First, the CH stretch peaks of
33
34 polyacetylene all occur at or above 3000 cm1, whereas CH stretch peaks are observed at 2865
35
36
37 and 2929 cm1 for the 480 K anneal. Second, these CH stretch peaks increase in intensity in the
38
39 480 K spectra relative to the 380 K spectra, whereas the 1006 cm1 peak decreases. Furthermore,
40
41 the 1006 cm1 peak behaves the same for all three surfaces, whereas the peaks at 2865 and 2929
42
43
44 cm1 are most intense for the Pd-free Cu(111) surface, and least intense for the highest Pd
45
46 coverage. These results indicate that the presence of Pd has little influence on the formation and
47
48 decomposition of polyacetylene on Cu(111). However, the presence of Pd slightly suppresses
49
50
51 formation of the second coupling product with characteristic peaks at 2929 and 2865 cm1, which
52
53 we assign to an oligomer.
54
55
56
57
3.3 C2H2 hydrogenation by SAA Pd/Cu(111) at elevated pressures
58
59
60 10
ACS Paragon Plus Environment
Page 11 of 27 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 Figure 3. A. Full RAIR spectra of C2H2 hydrogenation over 0.0055 ML Pd/Cu(111). B. RAIR spectra
25
of the CH2 wag region for gas phase C2H4. Hydrogenation is performed by admitting 110-2 Torr of
26
27
C2H2 followed by 1 Torr of H2 to the IR cell. The sample was then annealed to the listed temperatures
28 for 30 s prior to cooling to 300 K where spectra are obtained. Green numbers denote peaks due to gas
29 phase species.
30
31
32
33 Figure 3A shows the results of annealing a 0.0055 ML Pd/Cu(111) surface in the presence of
34
35 1 0-2 Torr of C2H2(g) and 1 Torr of H2(g). Comparison with the results in Figure 2 shows that
36
37 the addition of H2 had several effects. First, the gas phase acetylene peaks disappeared after the
38
39
40 480 K anneal, which was accompanied by the appearance of a gas-phase ethylene peak at 950
41
42 cm1. To see this peak more clearly, Figure 3B shows expanded spectra from 850 to 1100 cm1.
43
44
45
As previous work has shown, the 950 cm1 peak is the most intense one in the spectrum of gas
46
47 phase ethylene.29 In contrast to our earlier study of acetylene hydrogenation over a Pt(111)
48
49 surface, gas phase ethane is not observed.29 Figure 3 thus demonstrates that acetylene is
50
51
52 selectively hydrogenated to ethylene over this surface. Second, after annealing to 480 K, the
53
54 intensity of the oligomer peaks, such as the ones at 2929 and 2960 cm1, are more intense by a
55
56 factor of ~ 5 compared to the corresponding peaks in Figure 2. This result implies that hydrogen
57
58
59
60 11
ACS Paragon Plus Environment
ACS Catalysis Page 12 of 27

1
2
3
plays a significant role in the mechanism by which the oligomer forms. Third, the polyacetylene
4
5
6 peak at 1006 cm1 is almost entirely eliminated in the spectrum after annealing to 480 K,
7
8 whereas it is still present in Figure 2 when no H2(g) is present. The behavior of the 3014 cm1
9
10
11
peak is similar to the one at 1006 cm1, suggesting that the former can be assigned to a CH
12
13 stretch of polyacetylene. Although the 3014 cm-1 peak is close to the frequency of the
14
15 asymmetric CH stretch of gas phase methane, this possibility is excluded as it is not observed
16
17
18 when the experiment was performed with s-polarized light, which is sensitive only to gas-phase
19
20 molecules. The 3014 cm-1 peak also remains after the IR cell is evacuated providing further
21
22
23
24
Table 1 Comparison of peak position observed in Figure 3 with literature values for the upper phase
25 of green oil obtained from C2H2 hydrogenation over a Pd/Al2O3 catalyst.
26
27 Assignment Upper Phase Green Oil.a, Oligomer Peaks from Fig. 3b,
28 [cm1] [cm1]
29
30 CH3, asymm stretch 2960 2960
31
32
33
34 CH2, asymm stretch 2930 2929
35
36
37 CH2, symm stretch 2860 2857
38
39
40
41
CH2 def + CH3(as) def 1463 1463
42
43
44 CH3, symm def 1380 1378
45
46
47 Trans CH wagging 980 969
48
49
50
51 Vinyl CH2 wagging 920 909
52
53
54 a. Ref. 37
55
b. this work
56
57
58
59
60 12
ACS Paragon Plus Environment
Page 13 of 27 ACS Catalysis

1
2
3
evidence that it is not due to a gas phase molecule. Our assignment of the 3014 and 1006 cm1
4
5
6 peaks to polyacetylene is consistent with Shirakawa et al., where the IR spectrum of
7
8 polyacetylene produced at 150 C features only two prominent peaks, at 1015 and 3013 cm1.34
9
10
11
12 To more clearly distinguish between gas-phase and surface species, the s-polarized spectra
13
14 shown in Figure 4A were obtained under the same conditions as for Figure 3. With the 0.08 ML
15
16 Pd/Cu(111) surface at 300 K, after adding 1102 Torr of C2H2 and 1 Torr of H2 to the IR cell,
17
18
19 gas phase acetylene peaks are seen at 1300, 1352, 3265, and 3312 cm1. When the surface is
20
21 annealed to 340 K, the decrease in intensity of the C2H2(g) peaks is accompanied by the
22
23 appearance of the C2H4(g) peak at 950 cm1. The C2H2(g) peaks are absent in the 360 and 380 K
24
25
26 spectra and the disappearance of the 950 cm1 peak upon evacuation of the IR cell is consistent
27
28 with assigning it to a gas phase species. These results are in contrast to what we observed with s-
29
30
31
polarized spectra of acetylene hydrogenation over Pt(111) where the gas phase ethylene peak
32
33 was completely replaced upon annealing to 370 K by a peak centered at ~ 2980 cm1, which was
34
35 broadened due to unresolved rotational structure of the IR active CH stretches of ethane.29
36
37
38 There is some indication of IR absorption near 2980 cm1 for the 360 and 380 K spectra in
39
40 Figure 4A, but these features could be attributable entirely to a CH stretch of ethylene. For this
41
42 reason it is difficult to quantify the selectivity, but the results imply the amount of ethane
43
44
45 produced is at most only a few percent of the amount of the ethylene and could be much less.
46
47 Figure 4B shows the results of subtracting the s-polarized spectra from the p-polarized spectra
48
49
taken under the same conditions, which clearly reveals peaks attributable only to surface species.
50
51
52
53 The formation of oligomers during acetylene hydrogenation over Pd catalysts is well known
54
55 in the literature. In some cases a liquid byproduct has been separated from the catalyst and
56
57
58
characterized. This liquid has been termed green-oil, based on its color. As studied by Zhang et
59
60 13
ACS Paragon Plus Environment
ACS Catalysis Page 14 of 27

1
2
3
al., the consensus from various studies is that green oil is formed from acetylene through
4
5
6 coupling of two C2H2 molecules to form 1,3-butadiene, followed by formation of an oligomer
7
8 intermediate, from which the green oil forms.28 Srkny et al. have observed that the green oil
9
10
11
separates upon standing into two liquid phases and have presented IR spectra for both the upper
12
13 and lower phases.37 As indicated by the comparison in Table 1, there is a reasonable match
14
15 between their IR spectra of the upper green oil phase and the peaks of the product observed in
16
17
18 Figure 3. Furthermore, the intensity patterns are also similar, with the CH stretch peaks most
19
20 intense in the spectrum of Srkny et al. 37
21
22
23 Srkny et al.37 also observed intense peaks at 1750 and 1720 cm1 due to C=O stretches of
24
25
26 carbonyl groups that are absent in our spectra. However, they also note that CO is often added to
27
28 the gas stream to improve the selectivity of the hydrogenation reaction, and presumably they also
29
30
31
added CO in their experiment. Although they dont explicitly indicate this in their IR study, they
32
33 mention CO addition in a companion publication.38 In another study, it was suggested that the
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52 Figure 4. A. s-polarized spectra after 1102 Torr of C2H2 and 1 Torr of H2 were added to the IR cell
53 with the 0.08 ML Pd/Cu(111) crystal at 300 K followed by annealing the surface to the indicated
54 temperatures for 30 s. B. Subtraction of the s-polarized spectra from the p-polarized spectra to yield
55 spectra of the surface species.
56
57
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 27 ACS Catalysis

1
2
3
presence of ketone and carboxylic acids in NMR spectra of green oil indicated that it is formed
4
5
6 through a reaction between C2H2 and CO.39 Given the generally high intensities of the C=O
7
8 stretch peaks of carbonyls, the carbonyl-containing compounds in the green oil obtained by
9
10
11
Srkny et al. may constitute only a small amount in a mixture of compounds. Nevertheless, the
12
13 strong similarity between our spectra and the remaining peaks in their IR spectra strongly
14
15 suggests that the same coupling product is produced. What is not clear is whether green oil can
16
17
18 be distinguished by infrared spectroscopy from its oligomer precursor. Therefore we will refer to
19
20 the product with RAIRS peaks listed in Table 1 as the oligomer. The conclusion from Table 1
21
22 that a similar product is formed in the two cases demonstrates that the small amount of Pd in our
23
24
25 Pd/Cu(111) SAA is able to produce similar
26
27 surface chemistry to that of a Cu-free
28
29
Pd/Al2O3 catalyst.
30
31
32
33 The most prominent peaks of the oligomer
34
35 in the CH stretch region at 2929 and 2960
36
37
38 cm1 are assigned to asymmetric stretches of
39
40 CH2 and CH3 groups, respectively. It has been
41
42 proposed that the intensity ratio of these peaks
43
44
45 can be used to quantify the ratio of the number
46
47 of CH2 groups to the number of CH3 groups
48
49 Figure 5. Spectra after annealing to the indicated
and hence the chain length in the oligomers.40- temperatures a Pd-free Cu(111) surface (blue) and a
50
51 42
0.08 ML Pd/Cu(111) surface (red) in a background
52 Using the extinction coefficient ratio of 0.36 of 110-2 Torr of C2H2 and 1 Torr of H2. For the
53 annealing experiments, the crystal was heated to the
54 established by McNab et al.,40-41 we calculate target temperatures, held there for 30 s, then cooled
55 to 300 K where the spectra were obtained.
56
57
that the CH2:CH3 ratio of the oligomers in our
58
59
60 15
ACS Paragon Plus Environment
ACS Catalysis Page 16 of 27

1
2
3
samples is in the range of 8 to 11. However, with RAIRS there can be a definite orientation of
4
5
6 the surface species with respect to the electric field of the infrared radiation, which is restricted
7
8 to being perpendicular to the surface, whereas in a transmission IR experiment on high-surface
9
10
11
area powders a random orientation of the molecules with respect to the electric field of the light
12
13 can be assumed. Therefore, the extinction coefficient ratios determined by McNab et al.40-42 may
14
15 not be directly applicable to our results.
16
17
18
19 The role of Pd on the Cu(111) surface under acetylene hydrogenation conditions is
20
21 highlighted by the results in Figure 5 in which spectra over Pd-free Cu(111) are directly
22
23 compared with spectra obtained over a 0.08 ML Pd/Cu(111) surface. At 300 K, in the presence
24
25
26 of 110-2 Torr of C2H2 and 1 Torr of H2, the spectra from the two surfaces are almost identical.
27
28 After annealing to 320 K, the oligomer peaks become prominent over the Pd/Cu(111) surface
29
30
31
and continue to develop as the surface is annealed to temperatures up to 380 K. The peaks at
32
33 1008 and 3011 cm1 are clearly seen for both surfaces after annealing to 340 K and are again
34
35 attributed to polyacetylene. However, once the surface is annealed to 360 and 380 K, the
36
37
38 polyacetylene peaks are still quite prominent for the Cu(111) surface, but have disappeared from
39
40 the spectra for the Pd/Cu(111) surface with the oligomer peaks increasing. This suggests that
41
42 polyacetylene is hydrogenated to a green oil-like oligomer at these temperatures if and only if Pd
43
44
45 is present on the surface. Annealing to 360 K also marks the disappearance of the gas phase
46
47 acetylene peaks at 3265 and 3313 cm1 for Pd/Cu(111) but they are still present in the Cu(111)
48
49
spectra for the 360 and 380 K anneals. For Pd/Cu(111), the disappearance of the gas phase
50
51
52 acetylene peaks is accompanied by the simultaneous appearance of the gas phase ethylene peak
53
54 at 950 cm1. It is only after annealing to 380 K that some evidence of the characteristic CH
55
56
57
stretch peaks associated with the oligomer are seen for the Cu(111) surface. The results of Figure
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 27 ACS Catalysis

1
2
3
5 indicate that the Pd promotes oligomer formation in the presence of H2. Since the most obvious
4
5
6 role of Pd atoms in Cu(111) is to provide sites for H2 dissociation, the results suggest that the
7
8 availably of atomic hydrogen not only allows C2H2 hydrogenation over the Pd/Cu(111) surface,
9
10
11
but also promotes conversion of polyacetylene to the oligomer.
12
13
14 In general, C-H stretch values can be correlated with the hybridization on the carbon atom
15
16 and the sp2 carbon atoms of the (HC=CH) units of polyacetylene are consistent with C-H
17
18
19 stretch values above 3000 cm1. Similarly, the CH stretch vibrations of the oligomer below
20
21 3000 cm1 are consistent with sp3 hybridized carbon, as would be expected as C=C bonds are
22
23 replaced by C-C bonds by hydrogenation. Although the contrasting spectra in Figure 5 are
24
25
26 simply explained by assuming that the presence of Pd, in a background of H2(g), allows
27
28 polyacetylene that forms on Cu(111) sites to be hydrogenated to the oligomer, the results of
29
30
31
Figure 2 indicate that in the absence of H2(g), the oligomer can form on both Pd/Cu(111) and Pd-
32
33 free Cu(111), but that it requires higher temperatures. At the higher temperatures of Figure 2, the
34
35 influence of Pd is to actually decrease the amount of oligomer formed.
36
37
38
39 3.4 Trends with Pd coverage
40
41
42 The hydrogenation activity for two different Pd coverages compared to the behavior of the
43
44 Pd-free Cu(111) surface is clearly revealed by monitoring the CH stretch region for gas phase
45
46
47 acetylene as shown by the spectra in Figure 6. The black spectrum was obtained after adding
48
49 110-2 Torr of C2H2 and 1 Torr of H2 to the cell with the Cu(111) surface at 300 K. The surface
50
51
52
was then annealed to 500 K, cooled back to 300 K and the orange spectrum obtained. This shows
53
54 a slight decrease in the intensity of the peaks. A further decrease is observed for 0.0028 ML of
55
56 Pd, while for 0.0055 ML of Pd, the gas phase acetylene peaks are below the noise level. The
57
58
59
60 17
ACS Paragon Plus Environment
ACS Catalysis Page 18 of 27

1
2
3
disappearance of the gas phase acetylene peaks in Figure 5 is accompanied by the appearance of
4
5
6 a gas phase ethylene peak, confirming that the change seen in Figure 6 is associated with
7
8 acetylene hydrogenation to ethylene.
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 Figure 6. RAIR spectra for the C2H2 gas phase region after annealing surfaces with different Pd
34 coverages in a background of 1 102 Torr of C2H2 and 1 Torr of H2. Black, after adding C2H2 and H2 at
35
300 K to the Pd-free Cu(111) surface. Orange, after annealing the surface for the black spectrum to 500
36
37
K. Blue and green are for annealing a Cu(111) surface with 0.0028 and 0.055 ML of Pd to 480 K for 30 s.
38
39
40
41
42 An increased Pd coverage not only increases the extent of hydrogenation but also leads to a
43
44 decrease in carbon deposited on the surface as determined with AES. Figure 7 shows the
45
46
47 correlation between gas phase ethylene, oligomer, post-reaction carbon coverage, and Pd
48
49 coverage. In panels A, B, and C, each color corresponds to the amount of Pd on the surface. The
50
51 RAIR spectra shown for each surface were taken after gas phase C2H2 had disappeared from the
52
53
54 spectra. The temperature where this occurred was lower for higher Pd coverages. The only
55
56 changes with increasing annealing temperature are slight differences in the CH stretch
57
58
59
60 18
ACS Paragon Plus Environment
Page 19 of 27 ACS Catalysis

1
2
3
intensities. The results were obtained after annealing for 30 s to the temperatures (380-580 K)
4
5
6 needed to eliminate the gas phase acetylene peaks in a background of 1102 Torr C2H2 and 1
7
8 Torr of H2. In Figure 7A, RAIR spectra in the region of the gas phase ethylene peak are shown,
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
Figure 7. Summary of C2H2 hydrogenation results for increasing Pd coverages on Cu(111) after
52 annealing with different Pd coverages in a background of 1102 Torr of C2H2 and 1 Torr of H2. RAIR
53 spectra in the CH2 wag region of gas phase ethylene (A), and in the CH stretch region (B). Auger
54 electron spectra (C), plots of RAIRS peak data, and carbon coverage from AES versus Pd coverage to
55 show trends (D). In (A), (B), and (C), the colors correspond to surface Pd coverages of 0.0055 ML (blue),
56 0.08 ML (red), and 0.2 ML (black).
57
58
59
60 19
ACS Paragon Plus Environment
ACS Catalysis Page 20 of 27

1
2
3
with the blue spectrum corresponding to 0.0055 ML of Pd. In this case, the C2H4(g) peak at 950
4
5
6 cm1 is accompanied by peaks at 909 and 969 cm1 assigned to the oligomer. The ethylene peak
7
8 increases with Pd coverage but the oligomer peaks are absent at the higher Pd coverages in
9
10
11
Figure 7A. As shown in Figure 7B, the CH stretch peaks assigned to the oligomer are most
12
13 intense for the lowest Pd coverage, but dont seem to decrease monotonically for the higher Pd
14
15 coverages. The AES results in Figure 7C show a correlation between the amount of Pd and the
16
17
18 post reaction carbon coverage, with the lowest Pd coverage giving the highest carbon signal. The
19
20 results of Figure 7A-C are summarized in the plots in Figure 7D. For a Pd coverage of 0.08 ML,
21
22 it was also found that first admitting 110-2 Torr of H2 to the cell before adding C2H2(g) caused
23
24
25 an increase in the amount of C2H4(g) and a decrease in the amount of post-reaction carbon
26
27 detected with AES. However, the C-H stretch peaks of the oligomer were largely unaffected by
28
29
pre-exposure to hydrogen. This suggests that the pathway involved in converting C2H2 to C2H4 is
30
31
32 distinct from the pathway that leads to oligomer formation and that C2H4 is formed from
33
34 intermediates that can also decompose to deposit surface carbon. The first step of the
35
36
37 hydrogenation reaction must be to form a C2H3 species, the most likely of which is vinyl, while
38
39 the first step in the coupling reactions that lead to polyacetylene and the oligomer is presumably
40
41 the formation of a C4H4 metallacycle.
42
43
44
45 The hydrogenation reaction that we observe over the Pd/Cu(111) SAA converts a static
46
47 volume of gas phase acetylene into ethylene. Figure 8 shows a plot of the areas (in integrated
48
49
absorbance) of the C2H2(g) peaks in the 3200 to 3350 cm1 region (blue) and the C2H4(g) peak at
50
51
52 950 cm1 (green) as a function of time in the presence of a gas phase mixture of 0.1 Torr of C2H2
53
54 and 10 Torr of H2 for a 0.0055 ML Pd/Cu(111) surface at 400 K. The corresponding spectra
55
56
57
were obtained with 128 scans with 4 cm1 resolution, which required about 30 s per spectrum.
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 27 ACS Catalysis

1
2
3
There is a monotonic decrease in the acetylene
4
5
6 until it becomes undetectable after about 1500
7
8 s. The initial signal decreases linearly with one
9
10
11
slope until about 900 s, where there is a change
12
13 in slope, followed by another linear decrease.
14
15 The time dependence of the ethylene
16
17
18 concentration mirrors that of the acetylene,
19
20 offering further proof that acetylene disappears
21 Figure 8. Time dependence of the IR peak areas
22 through hydrogenation to ethylene. Some of the for C2H2(g) and C2H4(g) for 0.0055 ML
23
Pd/Cu(111) exposed to 0.1 Torr C2H2 and 10 Torr
24
25 acetylene is also lost through surface coupling H2 at 400 K. The peak areas were normalized to
26 the largest peak area.
27 reactions, but the amount should be negligible
28
29
compared to the total amount of gas phase acetylene present. From the volume of the reaction
30
31
32 cell, the initial partial pressure of acetylene, the Pd coverage, and the size of the Cu(111) crystal,
33
34 the time dependence of the acetylene disappearance can be used to estimate a turn over
35
36
37 frequency (TOF), i.e., the number of acetylene molecules converted to ethylene per second per
38
39 active surface atom. Division of the initial rate of acetylene decrease by the number of surface Pd
40
41 atoms yielded a TOF of 95 s-1. This, of course, assumes that the reaction takes place only at the
42
43
44 Pd sites. If we instead assume that all surface atoms, Cu and Pd, are active sites for
45
46 hydrogenation, then the TOF would be only 0.52 s-1. As STM images show that spill over does
47
48 not produce a uniform distribution of H atoms on the Pd/Cu(111) SAA,8 the TOF based on the
49
50
51 actual active sites must be between the extremes of 95 and 0.52 s1. The TOFs for acetylene
52
53 hydrogenation to ethylene over Pd catalysts are generally in the range of ~0.03 35 s-1.43-51
54
55
56
When Pd concentrations are lower, such as with a bimetallic catalyst, TOFs tend to be higher.
57
58
59
60 21
ACS Paragon Plus Environment
ACS Catalysis Page 22 of 27

1
2
3
Experimental measurements such as these of TOFs for a well-characterized single crystal SAA
4
5
6 provide the most direct way to compare results obtained on model catalysts to results for more
7
8 practical catalysts. As Boudart has noted, TOFs from single crystals do not suffer from mass-
9
10
11
and heat-transport limitations or from as much ambiguity as to the number of surface sites and
12
13 therefore set a standard by which the quality of TOFs from supported catalysts can be judged.52
14
15
16 4. CONCLUSIONS
17
18
19 By simultaneously monitoring gas phase and surface species when a Pd/Cu(111) single-atom
20
21 alloy was exposed to an ambient pressure of acetylene and hydrogen, it was found that acetylene
22
23 hydrogenation proceeds in the presence of a carbonaceous layer on the surface. The turn over
24
25
26 frequency obtained is in the same range as reported for practical acetylene hydrogenation
27
28 catalysts. Under the same conditions, hydrogenation of acetylene to ethylene does not occur on
29
30
31
the Pd-free Cu(111) surface. The presence of Pd also alters the nature of the carbonaceous
32
33 species by promoting the hydrogenation of the C=C bonds of polyacetylene that forms from
34
35 acetylene coupling reactions on Cu(111). The RAIR spectrum of the oligomer that forms on the
36
37
38 Pd/Cu(111) surface matches that of the green oil product formed during selective acetylene
39
40 hydrogenation over supported Pd catalysts. The results demonstrate that elevated acetylene
41
42 pressures over a model single-crystal Pd/Cu(111) SAA catalyst produce similar surface chemical
43
44
45 reactions as observed for practical Pd catalysts.
46
47 AUTHOR INFORMATION
48
49
Corresponding Author
50
51
52 * E-mail: mtrenary@uic.edu
53
54
55
56
57
58
59
60 22
ACS Paragon Plus Environment
Page 23 of 27 ACS Catalysis

1
2
3
ORCID
4
5
6 Michael Trenary: 0000-0003-1419-9252
7
8 Notes
9
10
11
12 The authors declare no competing financial interest.
13
14
15 ACKNOWLEDGEMENTS
16
17
18 This work was supported by a grant from the National Science Foundation (CHE-1464816).
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 23
ACS Paragon Plus Environment
ACS Catalysis Page 24 of 27

1
2
3
4
5
6
7 References
8
9 1. Bond, G. C.; Dowden, D. A.; Mackenzie, N., Trans. Faraday Soc. 1958, 54, 1537-1546.
10
11 2. McCue, A. J.; Anderson, J. A., Front. Chem. Sci. Eng. 2015, 9, 142-153.
12 3. McCue, A. J.; Guerrero-Ruiz, A.; Rodrguez-Ramos, I.; Anderson, J. A., J. Catal. 2016,
13
14 340, 10-16.
15 4. Friedrich, M.; Villaseca, S.; Szentmiklsi, L.; Teschner, D.; Armbrster, M., Materials
16
17 2013, 6, 2958.
18
19
5. Kyriakou, G.; Boucher, M. B.; Jewell, A. D.; Lewis, E. A.; Lawton, T. J.; Baber, A. E.;
20 Tierney, H. L.; Flytzani-Stephanopoulos, M.; Sykes, E. C. H., Science 2012, 335, 1209-
21
22 1212.
23 6. Aaen, A. B.; Laegsgaard, E.; Ruban, A. V.; Stensgaard, I., Surf. Sci. 1998, 408, 43-56.
24
25 7. Baber, A. E.; Tierney, H. L.; Lawton, T. J.; Sykes, E. C. H., ChemCatChem 2011, 3, 607-
26 614.
27
28 8. Tierney, H. L.; Baber, A. E.; Sykes, E. C. H., J. Phys. Chem. C 2009, 113, 7246-7250.
29 9. Fu, Q.; Luo, Y., ACS Catal. 2013, 3, 1245-1252.
30
31 10. Ma, L.-L.; Lv, C.-Q.; Wang, G.-C., Appl. Surf. Sci. 2017, 410, 154-165.
32 11. Cao, X. X.; Mirjalili, A.; Wheeler, J.; Xie, W.; Jang, B. W. L., Front. Chem. Sci. Eng.
33
34 2015, 9, 442-449.
35
12. Pei, G. X.; Liu, X. Y.; Yang, X. F.; Zhang, L. L.; Wang, A. Q.; Li, L.; Wang, H.; Wang,
36
37 X. D.; Zhang, T., ACS Catal. 2017, 7, 1491-1500.
38
39
13. McCue, A. J.; McRitchie, C. J.; Shepherd, A. M.; Anderson, J. A., J. Catal. 2014, 319,
40 127-135.
41
42 14. McCue, A. J.; Gibson, A.; Anderson, J. A., Chem. Eng. J. 2016, 285, 384-391.
43 15. McCue, A. J.; Anderson, J. A., J. Catal. 2015, 329, 538-546.
44
45 16. Bridier, B.; Lpez, N.; Prez-Ramrez, J., J. Catal. 2010, 269, 80-92.
46 17. Bridier, B.; Hevia, M. A. G.; Lpez, N.; Prez-Ramrez, J., J. Catal. 2011, 278, 167-172.
47
48 18. Cao, X. R.; Fu, Q.; Luo, Y., Phys. Chem. Chem. Phys. 2014, 16, 8367-8375.
49 19. Abdelrehim, I. M.; Pelhos, K.; Madey, T. E.; Eng, J.; Chen, J. G., J. Mol. Catal. A Chem.
50
51 1998, 131, 107-120.
52
20. Kyriakou, G.; Kim, J.; Tikhov, M. S.; Macleod, N.; Lambert, R. M., J. Phys. Chem. B
53
54 2005, 109, 10952-10956.
55
56
21. Ormerod, R. M.; Lambert, R. M.; Hoffmann, H.; Zaera, F.; Yao, J. M.; Saldin, D. K.;
57 Wang, L. P.; Bennett, D. W.; Tysoe, W. T., Surf. Sci. 1993, 295, 277-286.
58
59
60 24
ACS Paragon Plus Environment
Page 25 of 27 ACS Catalysis

1
2
3
4 22. Tysoe, W. T.; Nyberg, G. L.; Lambert, R. M., J. Chem. Soc., Chem. Comm. 1983, 623-
5 625.
6
7 23. Tysoe, W. T.; Nyberg, G. L.; Lambert, R. M., Surf. Sci. 1983, 135, 128-146.
8
24. Stacchiola, D.; Molero, H.; Tysoe, W. T., Catal. Today 2001, 65, 3-11.
9
10 25. Azad, S.; Kaltchev, M.; Stacchiola, D.; Wu, G.; Tysoe, W. T., J. Phys. Chem. B 2000,
11
12
104, 3107-3115.
13 26. Stacchiola, D.; Wu, G.; Molero, H.; Tysoe, W. T., Catal. Lett. 2001, 71, 1-4.
14
15 27. McCue, A. J.; Shepherd, A. M.; Anderson, J. A., Catal. Sci. Technol. 2015, 5, 2880-2890.
16 28. Zhang, J.; Sui, Z.; Zhu, Y.-A.; Chen, D.; Zhou, X.; Yuan, W., Chem. Eng. Technol. 2016,
17
18 39, 865-873.
19 29. Krooswyk, J. D.; Waluyo, I.; Trenary, M., ACS Catal. 2015, 5, 4725-4733
20
21 30. Krooswyk, J. D.; Yin, J.; Asunskis, A. L.; Hu, X.; Trenary, M., Chem. Phys. Lett. 2014,
22 593, 204-208.
23
24 31. Kruppe, C. M.; Krooswyk, J. D.; Trenary, M., J. Phys. Chem. C 2017, 121, 9361-9369.
25 32. Bandy, B. J.; Chesters, M. A.; Pemble, M. E.; McDougall, G. S.; Sheppard, N., Surf. Sci.
26
27 1984, 139, 87-97.
28
29
33. Chesters, M. A.; McCash, E. M., J. Electron. Spec. Rel. Phen. 1987, 44, 99-108.
30 34. Shirakawa, H.; Ikeda, S., Polymer J. 1971, 2, 231-244.
31
32 35. McCash, E. M.; Parker, S. F.; Pritchard, J.; Chesters, M. A., Surf. Sci. 1989, 215, 363-
33 377.
34
35 36. Mudiyanselage, K.; Yang, Y.; Hoffmann, F. M.; Furlong, O. J.; Hrbek, J.; White, M. G.;
36 Liu, P.; Stacchiola, D. J., J. Chem. Phys. 2013, 139, 044712.
37
38 37. Srkny, A.; Weiss, A. H.; Szilgyi, T.; Sndor, P.; Guczi, L., Appl. Catal. 1984, 12, 373-
39 379.
40
41 38. Srkny, A.; Guczi, L.; Weiss, A. H., Appl. Catal. 1984, 10, 369-388.
42 39. Borodziski, A.; Bond, G. C., Catal. Rev. 2006, 48, 91-144.
43
44 40. McNab, A. I.; McCue, A. J.; Dionisi, D.; Anderson, J. A., J. Catal. 2017, 353, 295-304.
45
41. McNab, A. I.; McCue, A. J.; Dionisi, D.; Anderson, J. A., J. Catal. 2017, 353, 286-294.
46
47 42. McNab, A. I.; Heinze, T.; McCue, A. J.; Dionisi, D.; Anderson, J. A., Spectrochim. Acta
48
49
Part A Mol. Biomol. Spectrosc. 2017, 181, 65-72.
50 43. Ruta, M.; Semagina, N.; Kiwi-Minsker, L., J. Phys. Chem. C 2008, 112, 13635-13641.
51
52 44. Borodziski, A., Catal. Lett. 2001, 71, 169-175.
53 45. Feng, J. T.; Liu, Y. N.; Yin, M.; He, Y. F.; Zhao, J. Y.; Sun, J. H.; Li, D. Q., J. Catal.
54
55 2016, 344, 854-864.
56 46. Duca, D.; Frusteri, F.; Parmaliana, A.; Deganello, G., Appl. Catal. A: Gen. 1996, 146,
57
58 269-284.
59
60 25
ACS Paragon Plus Environment
ACS Catalysis Page 26 of 27

1
2
3
4 47. Borodzinski, A.; Bond, G. C., Cat. Rev. Sci. Eng. 2008, 50, 379-469.
5 48. Kim, W. J.; Kang, J. H.; Ahn, I. Y.; Moon, S. H., J. Catal. 2004, 226, 226-229.
6
7 49. Zhang, Y. Y.; Diao, W. J.; Williams, C. T.; Monnier, J. R., Appl. Catal. A: Gen. 2014,
8
469, 419-426.
9
10 50. Li, Y. N.; Jang, B. W. L., Appl. Catal. A: Gen. 2011, 392, 173-179.
11
12
51. Asplund, S., J. Catal. 1996, 158, 267-278.
13 52. Boudart, M., Chem. Rev. 1995, 95, 661-666.
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 26
ACS Paragon Plus Environment
Page 27 of 27 ACS Catalysis

1
2
3
4
5
6 TOC Figure
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 27
ACS Paragon Plus Environment

You might also like