You are on page 1of 29

Subscriber access provided by UNIV OF TASMANIA

Article
Structure- and Temperature-Dependence of Pt-
Catalyzed Ammonia Oxidation Rates and Selectivities
Hanyu Ma, and William F. Schneider
ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.8b04251 • Publication Date (Web): 31 Jan 2019
Downloaded from http://pubs.acs.org on February 2, 2019

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides “Just Accepted” as a service to the research community to expedite the dissemination
of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in
full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully
peer reviewed, but should not be considered the official version of record. They are citable by the
Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore,
the “Just Accepted” Web site may not include all articles that will be published in the journal. After
a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web
site and published as an ASAP article. Note that technical editing may introduce minor changes
to the manuscript text and/or graphics which could affect content, and all legal disclaimers and
ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or
consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 28 ACS Catalysis

1
2
3
4
5
6
7
8 Structure- and Temperature-Dependence of
9
10
11
12
Pt-Catalyzed Ammonia Oxidation Rates and
13
14
15 Selectivities
16
17
18
19
Hanyu Ma and William F. Schneider∗
20
21
22
Department of Chemical and Biomolecular Engineering, University of Notre Dame, Notre
23
24
Dame, Indiana 46556, United States
25
26
E-mail: wschneider@nd.edu
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 1
60 ACS Paragon Plus Environment
ACS Catalysis Page 2 of 28

1
2
3
4
Abstract
5
6 Ammonia oxidation is operated at different temperatures over Pt catalysts of different struc-
7
8 tures to recover different products. In this work, we elucidate the dependency of ammonia
9
10 oxidation rates and selectivities on both Pt structure and temperature. We perform density
11
12 functional theory (DFT) computations to compare the reaction and activation energies of
13
14 elementary reactions on Pt(211) and Pt(111). We develop microkinetic model parameterized
15
16 with the DFT results. We show that barriers to product formation are lower on stepped
17
18 Pt than on terrace, leading to a much higher step rate at low temperature to selectively
19
20 oxidize ammonia to nitrogen. At high temperature, however, both step and terrace perform
21
22 comparably in rate to selectively produce nitric oxide. While N2 is always the thermody-
23
24 namic product, relative N and O coverages interact to make NO the kinetic product at high
25
26 temperature. The predicted rate and selectivity are consistent with experiments. We further
27
28 show rate-controlling steps on the two Pt surfaces are different at low temperature but are
29
30 the same at high temperature. The degrees of selectivity control for elementary reactions are
31
32 comparable for the two surfaces. Finally, we demonstrate the flows of elementary reactions
33
34 in the reaction network are also structure- and temperature-dependent.
35
36
37
38
39 Keywords
40
41
42 kinetic and thermodynamic control, density functional theory, microkinetic modeling, active
43
44 site, structure sensitivity, temperature-dependent catalysis, platinum, ammonia activation.
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 2
60 ACS Paragon Plus Environment
Page 3 of 28 ACS Catalysis

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 Introduction
22
23
24 Ammonia oxidation is a key step in the Ostwald process for production of nitric acid and, as
25
26 practiced in ammonia slip catalysis, important to environmental protection. 1 While both the
27
28 Ostwald process and ammonia slip are catalyzed by Pt-based materials, they are carried out
29
30 at different conditions, over materials of different structure, and with the goal of recovering
31
32 different products: 2,3
33
34
35
36 NH3 (g) + 1.25 O2 (g) −−→ NO(g) + 1.5 H2 O(g) (1a)
37
38 NH3 (g) + 0.75 O2 (g) −−→ 0.5 N2 (g) + 1.5 H2 O(g) (1b)
39
40 NH3 (g) + O2 (g) −−→ 0.5 N2 O(g) + 1.5 H2 O(g) (1c)
41
42
43
44 NO is the desired intermediate product in the Ostwald process, and to this end the
45
46 reaction is carried out at temperatures above 1000 K over a Pt gauze doped with Rh to
47
48 enhance mechanical strength and decrease Pt volatility. 2,4 In contrast, N2 is the desired
49
50 product in environmental protection, and applications to treating ammonia slip from diesel
51
52 engines and industrial emissions dictate that catalysts must be active at temperatures as low
53
54 as 450 K. 1,5 Here, the reaction is carried out over supported Pt nanoparticles that present a
55
56 larger fraction of edge and step atoms than does bulk Pt. 3 N2 O is an undesired by-product of
57
58
59 3
60 ACS Paragon Plus Environment
ACS Catalysis Page 4 of 28

1
2
3
both applications. 6 At standard pressures, free energy favors reaction 1b over reaction 1a and
4
5
1c from 400 to 1200 K, which leads to almost 100% N2 product selectivity at both Ostwald
6
7
and ammonia slip conditions (Figure S1). The question remains how kinetics contribute to
8
9 product differences, and the extent to which catalyst structure influences rate and selectivity.
10
11 Ultra-high vacuum (UHV) experiments suggest that stepped Pt surfaces are more ac-
12
13 tive than terraces for ammonia oxidation at low temperatures and pressures. A Pt(S)-
14
15 12(111)×(111) surface that presents a low density of step sites is reported to catalyze NH3
16
17 oxidation at nearly twice the rate per unit area as a Pt(111) terrace during steady-state re-
18
19 action from 373 to 900 K. 7 Zeng et al used temperature programmed reaction of Pt surfaces
20
21 pre-dosed with NH3 and O2 to conclude that surface-normalized product formation rates
22
23 are greater on stepped Pt surfaces and Pt foils than Pt(100) at temperatures below 800 K. 8
24
25 In both studies, the dominant product switches from N2 to NO around 700 K and stepped
26
27 surfaces are always more active than terraces. N2 O formation, however, is not observed un-
28
29 der UHV. During NH3 oxidation catalyzed by supported Pt in plug flow reactors, products
30
31 are similarly observed to transition from N2 at low to NO at high temperature. 9–12 N2 O is
32
33 often observed at low and intermediate temperatures over supported catalysts at ambient
34
35 pressure. Ostermaier et al report that NH3 oxidation turnover frequencies at temperatures
36
37 below 473 K increase with increasing particle size and produce N2 and N2 O as major and
38
39 minor products. The results are contradictory to UHV observations. The observed low ac-
40
41 tivity of small nanoparticles may not be intrinsic, because small nanoparticles show rapid
42
43 deactivation in the same study. 13
44
45 Density functional theory (DFT) models have been used to compare NH3 activation
46
47 pathways on Pt(111) and (100) 14,15 and Pt(211) surfaces. 16 Calculations suggest that N – H
48
49 bond cleaving steps are promoted by surface-bound O and/or OH on the three surfaces.
50
51 NH3 activation steps on O sites are computed to have comparable barriers on Pt(111) and
52
53 Pt(211). Pt(100) shows lower barriers than both Pt(111) and (211) for NH3 activation, but
54
55 Pt(100) is susceptible to reconstruction when exposed to oxygen. 17 DFT studies on Pt(111)
56
57
58
59 4
60 ACS Paragon Plus Environment
Page 5 of 28 ACS Catalysis

1
2
3
and Pt(211) consider a limited number of reaction pathways and do not connect results to
4
5
macroscopic reaction rates and selectivities.
6
7
In this work, we seek to elucidate the relationship between Pt surface structure, reaction
8
9 conditions, and NH3 oxidation reactivity. We use DFT models to compute reaction and
10
11 activation energies for all relevant NH3 oxidation pathways over a stepped Pt(211) and
12
13 a terrace Pt(111) surface. We develop and parameterize microkinetic models to predict
14
15 product formation rates over a wide range of temperatures. We show that both N2 and NO
16
17 are produced more readily at Pt steps than terraces, and that product formation rates on
18
19 Pt(211) are orders of magnitude higher than Pt(111) below 700 K to produce N2 . At higher
20
21 temperatures, however, the rates converge, such that both terraces and steps contribute
22
23 comparably to NH3 oxidation to NO.
24
25
26
27
28 Computational details
29
30
31 DFT calculations were performed with the Vienna ab initio Simulation Package (VASP) 18,19
32
33 using the projector augmented wave (PAW) core model. 20,21 Electron exchange and correla-
34
35 tion were treated with the PBE generalized gradient approximation (GGA). 22 Plane waves
36
37 were included for the electronic wave functions up to a cutoff energy of 400 eV. Forces were
38
39 converged to 0.05 eV/Å. Transition states were located by the Climbing-Image Nudged Elas-
40
41 tic Band method (CINEB) 23 and were further confirmed by the existence of one imaginary
42
43 frequency along the reaction coordinate.
44
45 The lattice constants calculated using the Birch–Murnaghan equation of state 24,25 are
46
47 3.968 Å for Pt, in correspondence with the experimental values of 3.930 Å. 26 The (211)
48
49 surface was modeled with a 3 × 2 supercell, using a five-(111)-layer-thick slab separated from
50
51 its periodic image by a 14 Å vacuum gap. The top three (111) metal layers were allowed
52
53 to relax in all optimizations and bottom two (111) layers were fixed to mimic the bulk
54
55 metal. The first Brillouin zone was sampled with a Γ-centered 5 × 5 × 1 Monkhorst–Pack
56
57
58
59 5
60 ACS Paragon Plus Environment
ACS Catalysis Page 6 of 28

1
2
3
k-point mesh. 27 The (111) surface was modeled with a 2 × 2 supercell, using a five-layer-
4
5
thick slab separated by a 14 Å vacuum gap. The top three metal layers were allowed to relax
6
7
and bottom two layers were fixed. The first Brillouin zone was sampled with a Γ-centered
8
9 7 × 7 × 1 Monkhorst–Pack k-point mesh. 28
10
11 The rate constant for non-activated adsorption steps were calculated using the Hertz-
12
13 Knudsen equation (Eqn. 2), where α is a sticking coefficient, A is the area of one free site, m
14
15 is the molecular weight of a gas molecule, k B is the Boltzmann constant and T is the absolute
16
17 temperature. 29 We assume unity sticking coefficient as an upper limit on adsorption rates. 30
18
19
20
21 αA
kads = √ (2)
22 2πmkB T
23
24 For the desorption steps, the adsorption equilibrium constants were first computed (Eqn.
25
26 3). The enthalpies of adsorption are assumed to be the zero point energy corrected DFT
27
28 energies. The entropies of gas molecules are from NIST—JANAF thermochemical tables 31
29
30 and the entropies of adsorbates were estimated with harmonic oscillator assumption.
31
32
33 ∆EDFT + ∆EZP E − T ∆S ◦ (T )
!
34 K(T ) = exp − (3)
35
kB T
36
37
38
The rate constants of desorption processes can thus be written as follows, where P ◦ is the
39
40
standard pressure of 1 bar:
41 kads P ◦
kdes = (4)
42 K
43
44 Reactions rate constants (k) involving activated chemisorption and surface reactions were
45
46 calculated with harmonic Transition State Theory (TST):
47
48
49 kB T qTS

Ea

50 k= exp − (5)
51 h qIS kB T
52
53
54 Here h is the Planck constant and E a is the zero-point-corrected energy difference between
55
56 the transition state and the adsorbed reactants at infinite separation.
57
58
59 6
60 ACS Paragon Plus Environment
Page 7 of 28 ACS Catalysis

1
2
3
Partition functions, q TS for transition states and q IS for initial states are calculated by
4
5
taking the zero point energies as the reference. We consider vibrational degrees of freedom
6
7
for adsorbates, and they are calculated in the harmonic oscillator model in Eqn. 6 , where v
8
9 is the vibrational frequencies excluding the imaginary ones:
10
11
12 nY
real
1
13 q=   (6)
14 i=1 1 − exp − khv
BT
i

15
16
17 The surface coverages at steady state are solved with a mean-field microkinetic model. 32
18
19 The coverages are further used to calculate the net rates for product formation. The differ-
20
21 ential equation for the coverage of surface species i can be written as:
22
23
24 ∂θi X
25
= cij rj (7)
∂t j
26
27
28 with
29
−c −c c c
30 rj = k+j (8)
Y Y Y Y
θ+j ij P+j ij − k−j ij
θ−j P−jij
31 +j +j −j −j
32
33 where cij is the coefficient for the stoichiometry of species i in surface reaction j, rj is the
34
35 net rate of surface reaction j (forward and backward reactions are represented as +j and
36
37 −j), kj is the rate constant and θj and Pj are the coverages and pressures of different species
38
39 involved in reaction j. The site number is conserved for a surface:
40
41
42
θ∗ + θi = 1 (9)
X
43
44 i
45
46 The ordinary differential equations (ODE) are solved at steady state, where ∂θi
= 0. At
47 ∂t
48 steady state, the coverages can also be solved from a system of algebraic equations (
P
cij rj =
j
49
50 0). Using the coverages solved from ODE as initial guesses, the steady state coverage are
51
52 further confirmed by solving the system of algebraic equations.
53
54
55
56
57
58
59 7
60 ACS Paragon Plus Environment
ACS Catalysis Page 8 of 28

1
2
3
O* O* O* NO*
4
NH3* NH2* NH* N* NO
5
6
7 NH3 N2O
8
9 N2
10 OH* OH* OH*
11
12 Scheme 1: Schematic representation of reaction pathways relevant to ammonia oxidation.
13 Each bar represents an elementary step. Light pink bar denotes NH3 adsorption, light blue
14 bars denote reaction of NH3 fragments with adsorbed O*, and grey bars denote reaction of
15 NH3 fragments with adsorbed OH*. Orange, pink, and blue bars indicate pathways leading
16
17 to NO, N2 O, and N2 , respectively.
18
19
20 Results and Discussion
21
22
23 To model catalytic NH3 oxidation reactivity and selectivity, we computed the reaction and
24
25 activation energies of reactions relevant to ammonia oxidation on Pt(211) and Pt(111) using
26
27 plane wave, supercell DFT within the generalized gradient approximation. 19,20,22 We include
28
29
pathways evidenced by previous UHV and DFT results, as shown in Scheme 1. Detailed
30
31
reactions are shown in Table S2. In short, NH3 activation proceeds at adsorbed O (O*) and
32
33
OH (OH*) sites instead of metal sites and all product bond formation occurs after complete
34
35
NH3 reaction to N*. Consistent with this network, NH3 oxidation products are observed
36
37
after dosing NH3 and heating O-precovered Pt(111), while no such products are produced
38
39
from clean Pt(111). 33 Surface-bound NHx (x = 0, 1, 2, 3) and OHy (y = 0, 1) intermediates
40
41
are observed in these experiments over Pt 33 and Ir. 34 DFT-computed N – H dissociation
42
43
activation energies at O* and OH* sites are less than those for N – H dissociation at metal
44
45
sites over Pt, 14 Rh, 35 Au 36 and Cu. 37
46
47
We first computed the prefered binding sites of NH3 oxidation reactants, products, and
48
intermediates including NH3 , NH2 , NH, N, O, NO, N2 O, H2 O and OH on Pt(211) and
49
50
Pt(111) slabs. The most stable configurations and binding energies are presented in Figures
51
52
S2 and S3 and Table S1 and are consistent with previous reports. 27,38–40 Briefly, O, OH,
53
54
NH2 , NO and N2 O bind on bridge sites over the step of Pt(211), resulting in their enhanced
55
56
57
58
59 8
60 ACS Paragon Plus Environment
Page 9 of 28 ACS Catalysis

1
2
3
4 NH3 + * + 3O* → N* + 3OH*
a NH3 + * + 3O* → N* + 3OH*
b
5 NH3 + * + 3OH* → N* + 3H2O* NH3 + * + 3OH* → N* + 3H2O*

NH2*+2O*
NH3*+3OH*

NH3*+3O*
7

NH2*+2OH*
NH3+* NH3+*
N*
8 NH*+OH* N*

9 NH*+O*

NH3*+3OH*
NH3*+3O*
10

NH2*+2OH*
NH*+O*
11 NH2*+2O* N* NH*+OH*
N*

12
13
14 Pt(211) c Pt(111) f
15 NH3-O* NH3-O*
16
17
18
19
20 Pt(211) d Pt(111) g
NH2-O* NH2-O*
21
22
23
24
25
26
Pt(211) e Pt(111) h
NH-O* NH-O*
27
28
29
30
31
32
Figure 1: Energy diagrams for O*- and OH*-assisted ammonia activation on (a) Pt(211)
33
34 and (b) Pt(111). (c), (d) and (e) are transition states of O*-assisted NH3 *, NH2 * and NH*
35 activation on Pt(211), respectively. (f), (g) and (h) are transition states of O*-assisted NH3 *,
36 NH2 * and NH* activation on Pt(111), respectively. Pt: silver; O: red; N: blue; H: white.
37
38
39 binding relative to Pt(111). Pt(211) binds NH2 , NO and OH more than 0.5 eV stronger than
40
41 Pt(111) and oxygen 0.26 eV stronger. The other species bind on atop or hcp sites on Pt(211)
42
43 and have comparable binding energies on Pt(111). We compute the energies of elementary
44
45 reactions with respect to the adsorbates on their most favorable binding sites at infinite
46
47 separation.
48
49
50
51 Ammonia activation
52
53
54
We identified transition states for reactions relevant to ammonia activation using the climb-
55
56
ing image nudged elastic band (CI-NEB) method (Table S2). 23 The transition states are
57
58
59 9
60 ACS Paragon Plus Environment
ACS Catalysis Page 10 of 28

1
2
3
confirmed by the existence of only one imaginary frequency along the reaction coordinate
4
5
(Table S4). Figure 1(a) and (b) compare potential energy surfaces for successive N – H cleav-
6
7
age steps on Pt(211) and Pt(111). Figure 1(a) and (b) include reaction (2) to (8) in Table
8
9 S2. The initial, transition and final states of each reaction are in Figure S4-S7. In general,
10
11 activation barriers are comparable on Pt(211) and Pt(111). The preferred pathways, how-
12
13 ever, are different for the two surfaces. On Pt(211), O*-assisted ammonia activation has
14
15 lower reaction and activation energies than the OH*-assisted route, except for the activation
16
17 of NH2 *. On Pt(111), however, OH*-assisted pathways have lower reaction and activation
18
19 energies than the O*-assisted ones, as also observed for (111) surfaces of Pt and Rh in other
20
21 DFT computations. 14,35 The difference of reaction energies between O*- and OH*-assisted
22
23 pathways (Figure 1) are related to reaction (9) in Table S2, i.e. OH* + OH* ←−→ O* +
24
25 H2 O*. For example, the energy difference between NH3 * + OH* ←−→ NH2 * + H2 O* and
26
27 NH3 * + O* ←−→ NH2 * + OH* is the reaction energy of reaction(9) in Table S2, which is
28
29 positive on Pt(211) and negative on Pt(111).
30
31 Figure 1(c)-(h) compare the transition states for ammonia activation by O* on Pt(211)
32
33 and (111), as examples. Results for ammonia activation by OH* are in Figures S6 and
34
35 S7. On Pt(211), reaction can happen on both step (Figure 1(c) and (e)) and terrace sites
36
37 (Figure 1(d)), while Pt(111) presents only terrace sites.
38
39 Reaction pathways are in general a combination of H-transfer and adsorbate migration
40
41 steps, driven by the different site preferences of the NHx species. In NH2 and NH oxidation
42
43 (Figure 1 (d), (e), (g) and (h)), the H-transfer process has the highest activation energy. In
44
45 NH3 oxidation (Figure 1 (c) and (f)), however, the migration steps determine the maximum
46
47 barrier. Similarly, the transition states of N – H cleavage steps over Rh(111), 35 Au(111) 36 and
48
49 Cu(111) 37 are correspond to transfer of H and migration of NHx . Activation energies and
50
51 transition states for NHx (x = 1,2,3) oxidation by O on Pt(211) computed here are similar
52
53 to those reported previously, 16 and corresponding results by OH have not been reported.
54
55 Previously reported transition states and activation energies for NHx (x = 1,2,3) activation
56
57
58
59 10
60 ACS Paragon Plus Environment
Page 11 of 28 ACS Catalysis

1
2
3
4
5
N* + N* → N2 + 2*
N* + O* → NO + 2*
a N* + N* → N2 + 2*
N* + O* → NO + 2*
b
N* + NO* → N2O + 2* N* + NO* → N2O + 2*
6 NO
7
N2O* N2O NO
8 N2O*
9 N*
N2*
N* NO*
N2O
10 N2
NO* N2*
11 N2
12
13
14 Pt(211) c Pt(111) f
15 N-N* N-N*
16
17
18
19
20 Pt(211) d Pt(111) g
N-O* N-O*
21
22
23
24
25 Pt(111) h
26
Pt(211) e
N-NO* N-NO*
27
28
29
30
31
32
Figure 2: Energy diagrams for N2 , NO and N2 O formation on (a) Pt(211) and (b) Pt(111).
33
34 (c), (d) and (e) are transition states of N2 , NO and N2 O formation on Pt(211), respectively.
35 (f), (g) and (h) are transition states of N2 , NO and N2 O formation on Pt(111), respectively.
36 Pt: silver; O: red; N: blue; H: white.
37
38
39 by O* and OH* on Pt(111) are similar to those reported here. The one exception is NH3 *
40
41 activation by O*, which has been reported to involve H transfer and to have an activation
42
43 energy of 0.27 eV. 14 We explored the complete pathway, find the highest energy point to
44
45 involve NH2 migration, and to have a barrier of 0.70 eV.
46
47
48
49 Product formation
50
51
52
We further calculate the energies and barriers for N2 , NO and N2 O formation with respect
53
54
to the adsorbates on their most favorable binding sites at infinite separation. Figure 2(a)
55
56
and (b) compare energy diagrams for N* reaction with N*, O* and NO* to form N2 , NO
57
58
59 11
60 ACS Paragon Plus Environment
ACS Catalysis Page 12 of 28

1
2
3
and N2 O, respectively, on Pt(211) and Pt(111). The initial, transition and final states of
4
5
each reaction are in Figure S8 and S9. N2 * and NO* formation have 1.34 and 0.74 eV lower
6
7
barriers on Pt(211) than Pt(111), respectively. The barriers for N2 O* formation on the two
8
9 slabs, however, are comparable. The ordering of reaction barriers is reversed on the two
10
11 surfaces: Pt(211) follows N2 < NO < N2 O while Pt(111) follows N2 > NO > N2 O. These
12
13 differences might suggest selectivity differences on the two facets. Figure 2(c)-(h) compare
14
15 the transition states for product formation on Pt(211) and Pt(111). N2 and NO association
16
17 reactions happen on step sites on Pt(211), resulting in lowered barriers compared to Pt(111).
18
19 N2 O formation proceeds through similar transition states on the terrace of Pt(211) and on
20
21 Pt(111), with similar barrier. Previously reported activation energies for N* reaction with
22
23 N*, O* and NO* to form N2 , NO and N2 O, respectively, on Pt(111) follow the same order
24
25 as reported here. 41 Step-enhanced association of N2 and NO and the similar barriers of N2 O
26
27 association on Pt(211) and Pt(111) are in agreement with a previous study. 40
28
29 To make the reaction network complete, we computed the reaction energies and barriers
30
31 for O2 dissociation and OH* + OH* ←−→ O* + H2 O* and show the energy diagrams in
32
33 Figure S10. O2 dissociative adsorption proceeds with molecular adsorption of O2 and O – O
34
35 bond cleavage. Because the transition state energy is below the O2 gas energy, we treat O2
36
37 dissociative adsorption as barrierless in the microkinetic model below. 42,43 The OH* + OH*
38
39 reactions are of small barrier and barrierless on Pt(211) and Pt(111), respectively. 44,45
40
41
42
43 Rate and selectivity temperature-dependence
44
45
46 To connect the energy diagrams of Pt(211) and Pt(111) to macroscopic reactivity and se-
47
48 lectivity, we developed mean-field microkinetic models to compare the steady-state reaction
49
50 rates and selectivities at fixed conditions relevant to both ammonia slip control and Ost-
51
52 wald process. 32 Following conventional approaches, we assume adsorbate diffusion is much
53
54 more rapid than surface reaction. 46 We compute rates at zero ammonia conversion to cap-
55
56 ture the intrinsic activities of the surfaces. 47 Activation energies of elementary reactions are
57
58
59 12
60 ACS Paragon Plus Environment
a b

NH3 Slip Ostwald


Control Process

c d
NH3 Slip Ostwald
Control Process

Figure 3: (a) Reaction rate against reciprocal temperature on Pt(211) and Pt(111) at
ammonia slip catalysis conditions and zero ammonia conversion. Pressures of NH3 , O2
and H2 O are 0.001, 0.02 and 0.05 bar, respectively. (b) Reaction rate against reciprocal
temperature on Pt (211) and Pt(111) at Ostwald process conditions and zero ammonia
conversion. Pressures of NH3 , air and H2 O are 0.9, 8.1 and 0.036 bar, respectively. (c)
Selectivities to reaction products on Pt(211) and Pt(111) at the same conditions as (a). (d)
Selectivities to reaction products on Pt(211) and Pt(111) at the same conditions as (b).

13
ACS Catalysis Page 14 of 28

1
2
3
from Figure 1 and 2 and detailed values are in Table S2. Prefactors are calculated with
4
5
Hertz–Knudsen equation for non-activated adsorptions and with harmonic transition state
6
7
theory for activated reactions (Table S2). 29 Surface coverages on Pt(211) and Pt(111) at
8
9 ammonia slip conditions are shown in Figure S11. From 400 to 1200 K, the most abundant
10
11 surface intermediate (MASI) changes from N* to NO* to O* with increasing T on Pt(211),
12
13 while NO is never the MASI on Pt(111). Above 1000 K, O* is the MASI on both surfaces.
14
15 Figure 3(a) reports the absolute turnover frequencies (TOFs) on Pt(211) and Pt(111) at
16
17 ammonia slip conditions in an Arrhenius plot. The TOF on Pt(211) is orders of magnitude
18
19 greater than that on Pt(111) at temperatures below 700 K, but TOFs converge on the two
20
21 surfaces with increasing temperature. From linear fits to the Arrhenius plots, the apparent
22
23 activation energies (Ea,app ) on Pt(211) and Pt(111) are 51 and 239 kJ/mol, respectively, so
24
25 that the TOF difference between (211) and (111) decreases with increasing T . Figure 3(b)
26
27 shows the absolute TOFs at Ostwald conditions, where the pressures of ammonia and oxygen
28
29 are much higher. Similarly, we observe Pt(211) to be much more active at low T while the
30
31 TOF difference between two surfaces diminishes with increasing temperature. The TOF
32
33 convergence shifts to higher T with increasing pressure.
34
35 Figure 3(c) compares product selectivities at ammonia slip catalysis conditions. Selec-
36
37 tivities are defined as the relative rates of products formation, s:
38
39
40 ri
41 si = P (10)
42 j rj
43
44
45 where i represents a product and j indexes over the three products, N2 , NO and N2 O. On
46
47
Pt(211), the N2 selectivity decreases and NO selectivity increases with T , so that prod-
48
49
ucts become kinetically controlled at high temperatures. The increasing selectivity to-
50
51
wards NO with increasing T has both rate-constant and N* vs O* coverage contributions.
52
53
kNO,forward /kN2 ,forward increases from 2.3×10−4 at 400 K to 0.046 at 1200 K, while N* coverage
54
55
decreases relative to O* coverage across the same temperature range (Figure S11).
56
57
58
59 14
60 ACS Paragon Plus Environment
Page 15 of 28 ACS Catalysis

1
2
3
The transition from N2 to NO product with T occurs at a lower temperature on Pt(111)
4
5
than Pt(211). The lower barrier for NO formation than N2 formation (Figure 2(b)), however,
6
7
leads to kNO,forward /kN2 ,forward decreasing from 9.5×103 at 400 K to 18 at 1200 K. Despite this
8
9 weaker elementary step selectivity towards NO, diminishing N* and increasing O* coverage
10
11 still leads to the same temperature-dependent selectivity as Pt(211). The results indicate
12
13 that surface coverage plays a more significant role than products formation barriers in de-
14
15 termining product selectivity. Similar factors are at play at Ostwald process conditions,
16
17 but the transition from N2 to NO selectivity shifts upward by about 300 K, as shown in
18
19 Figure 3(d). To test the robustness of these conclusions to finite conversions and product
20
21 pressures, we modeled a continuous stirred-tank reactor (CSTR) at ammonia slip conditions
22
23 (SI, P15). Structure- and temperature-dependent rates and selectivities are consistent with
24
25 the zero-conversion model. These results indicate that Pt steps are more active at low T
26
27 for ammonia slip control and is selective to N2 to a higher T than terrace Pt. Terrace Pt,
28
29 however, is as active as step Pt and is more selective to NO at the elevated T appropriate
30
31 to the Ostwald process.
32
33 The computed selectivity transitions at about 700 and 600 K on Pt(211) and Pt(111),
34
35 respectively, are consistent with available experiment. Imbihl et al measured N2 and NO
36
37 production rates over a Pt gauze at PO2 = PNH3 = 0.03 bar in an adiabatic tubular reactor
38
39 and observed a product transition between 700 and 750 K. 16 Similarly, Baerns et al measured
40
41 product formation rates over a Pt foil at atmospheric pressure in a continuous-flow micro-
42
43 structured reactor and found a transition at about 730 K when PO2 = 0.045 bar and PNH3 =
44
45 0.03 bar. 48
46
47 Baerns et al have reported NH3 oxidation rates at 10−7 bar over Pt(533). 48 We reran
48
49 the Pt(211) model at the same conditions; results are reported in Figure S13. Consistent
50
51 with observation, we find that overall rates are non-Arrhenius and much less sensitive to
52
53 temperature than predicted at ambient pressure. N2 is the primary oxidation product up to
54
55 600 K, above which the N2 production rate declines and NO becomes the primary product.
56
57
58
59 15
60 ACS Paragon Plus Environment
ACS Catalysis Page 16 of 28

1
2
3
N2 O formation rates are always at least an order of magnitude less than the rates to form
4
5
the primary products, consistent with the absence of N2 O product in low-pressure oxidation.
6
7
Rafti et al subsequently parameterized a microkinetic model for this system using a
8
9 combination of experimental observation-fitted and DFT-reported parameters. 49 The model
10
11 considers the same reaction network reported here except to neglect backward reactions
12
13 involving OH*. Model parameters for O*-assisted NH3 activation are taken from DFT
14
15 results for steps, computed on Pt(211), model parameters for OH*-assisted NH3 activation
16
17 taken from DFT results computed on Pt(111), and product formation steps are fitted to
18
19 match experimental observations. DFT parameters are comparable to those reported here,
20
21 while fitted reactions have lower barriers than computed here on either steps or terraces.
22
23 The primary difference between this semi-empirical model and the model here is that the
24
25 former predicts a higher coverage of N and NHx . Both models predict a transition from N2
26
27 to NO as the primary reaction product across the same temperature range. OH* was argued
28
29 to be unimportant for NH3 oxidation based on its low surface coverage. 49,50 As we show
30
31 below, despite that low coverage, the OH*-assisted NHx activation pathways do contribute
32
33 significantly to overall NH3 oxidation, especially at low temperature.
34
35 Rebrov et al parameterized a mean-field microkinetic model for ambient-pressure NH3
36
37 oxidation using rate constants fitted to experimental observation. 51 The mechanism ignores
38
39 OH*-assisted ammonia activation. Fitted ammonia activation energies by O* are greater
40
41 and for product formation are lower than the Pt(211) parameters computed here. When
42
43 solved at atmospheric pressure and in excess oxygen, the model predicts the most abundant
44
45 surface species to switch from N* to NO* and the product to switch from N2 to N2 O and
46
47 NO from 473 to 648 K. Both predictions are in agreement with the Pt(211) results here.
48
49 Results above assume ammonia decomposition pathways are irrelevant and the sticking
50
51 coefficients are unity. For completeness, we incorporated direct NHx + * dissociation events
52
53 into the reaction network (Table S3). 27,45,52 Direct dissociation rates are much less than
54
55 the net rates of other NH3 -consuming pathways. Overall rates and selectivities on both
56
57
58
59 16
60 ACS Paragon Plus Environment
Page 17 of 28 ACS Catalysis

1
2
3
4 Pt(211) Pt(111)

5 a N*+O* b
NH3*+O*
6 N*+N*
NH3*+O*
7 N*+N* NO*=NO+*
8 N*+NO*
9 O2+2*
10
O2+2*
11
12 N*+O*
13
14
15
16 c d
NH3*+O*
17 NH3*+O*

18 N*+N* N*+N*

19
20 O2+2* O2+2*
21
22 N*+O* N*+O*
23
24
25
26
27
28 Figure 4: Degree of rate control on (a) Pt(211) and (b) Pt(111). Degree of N2 selectivity
29 control on (c) Pt(211) and (d) Pt(111). The partial pressures of NH3 , O2 and H2 O are 0.001,
30 0.02 and 0.05 bar, respectively.
31
32
33 Pt(211) and Pt(111) are unchanged (Figure S14), as are the degree of rate control values
34
35 for all reactions. We tested the sensitivity of the microkinetic model results to the assumed
36
37 unity molecular and dissociative sticking coefficients. We find molecular adsorption to be
38
39 equilibrated at all conditions and thus to be insensitive to sticking coefficient. O2 dissociation
40
41 is not always equilibrated. With decreasing O2 sticking coefficient, the O coverage decreases,
42
43 the overall rate increases and the NO selectivity decreases on both Pt(211) and Pt(111).
44
45 Nonetheless, predicted conclusions regarding trends in coverages, rates, and selectivities are
46
47 insensitive to O2 sticking coefficients reduced by 0.1 and 0.01, as shown in Figure S15 and
48
49 S16. We further analyze the results at unit sticking coefficient below.
50
51
52
53
54
55
56
57
58
59 17
60 ACS Paragon Plus Environment
ACS Catalysis Page 18 of 28

1
2
3
4
Degree of rate and selectivity control and reaction flow pathways
5
6 The temperature-dependence of the overall rate of ammonia oxidation, as indicated by the
7
8 apparent activation energy, is different on steps and terraces. To identify the elementary
9
10 steps most responsible for controlling the reaction rates on steps and terraces, we performed
11
12 degree of rate control (DRC) across temperatures. 53 We define rate sensitivity to specific
13
14 reactions in Eqn. 11. XRC,i is calculated by changing the forward (k+i ) and reverse (k−i )
15
16 rate constants of step i simultaneously by 10%, and monitoring the change of the overall rate
17
18 (r) with respect to the reference rate, r0 . DRC results using a 1% change in rate constant
19
20 differ by less than 10 %.
21 ∂lnr r − r0
22 XRC,i = ≈ (11)
23 ∂lnki 0.1 × r0
24
25 Figure 4(a) and (b) show elementary reactions on steps and terraces controlling the overall
26
27 rate. The DRC values sum to 1, as expected. 53 We compare the DRC values for different
28
29 reactions vs T to understand the dependence of rate on T and surface structure. When
30
31 T < 600 K, terrace TOF is controlled by N2 association, which corresponds with a barrier
32
33 of 2.53 eV. On Pt(211), TOF is controlled by multiple reactions, with barriers lower than
34
35 1.62 eV. The rate-controlling steps on Pt(211) show lower barriers than the one on Pt(111),
36
37 leading to its higher TOF (Figure 3). When 600 < T < 1000 K, the overall rates are
38
39 controlled by NO formation on Pt(111), whose barrier is higher than the elementary steps
40
41 controlling the rate on Pt(211) by at least 1.5 eV. We therefore observe higher TOF on
42
43 Pt(211). When T > 1000 K, TOF is controlled positively by NH3 * + O* on both Pt(211)
44
45 and Pt(111). The similar barriers of this step, i.e. 0.58 and 0.70 eV on Pt(211) and (111)
46
47 respectively, are responsible for their similar TOFs at high temperatures.
48
49 As with activity, ammonia oxidation selectivity has a temperature dependence on both
50
51 surfaces. The temperature dependence is similar at steps and on terraces, but selectivity
52
53 transitions from N2 to NO at lower temperature on the terrace than step. To identify the
54
55 elementary steps responsible for this selectivity behavior, we computed degree of selectivity
56
57
58
59 18
60 ACS Paragon Plus Environment
Page 19 of 28 ACS Catalysis

1
2
3
control (DSC) towards N2 (Eqn. 12) by replacing rate with selectivity in Equation 11. 53
4
5
6 ∂lns s − s0
7 XSC,i = ≈ (12)
8 ∂lnki 0.1 × s0
9
10
11 The DSC values sum to 0, as expected. 53 The same four reactions control N2 selectivity on
12
13 both surfaces, as shown in Figure 4(c) and (d). NH3 * activation by O*, which contributes
14
15 to the formation of N*, and N2 formation itself contribute positively to N2 selectivity. O2
16
17 adsorption and NO* formation contribute negatively because they increase the formation of
18
19 both NO and N2 O. Furthermore, DSC values are comparable on Pt surfaces below 500 K and
20
21 above 1000 K. DSC is similar on Pt(211) and Pt(111), so that selectivity is not as sensitive
22
23 to Pt structure as T , consistent with Figure 3.
24
25 While degree of rate and selectivity control reveal the reactions most relevant to observed
26
27 rate and selectivity, they do not show the flow of each elementary reaction within the reaction
28
29 network. To compare the relative contribution of O*- and OH*-assisted NHx * activation to
30
31 the overall rate, we report the net rate of each reaction step normalized to the overall TOF
32
33 on Pt(211) and Pt(111) at 500 and 1000 K in the form of Sankey diagrams (Figure 5). For
34
35 example, the relative segment widths connecting NH3 * activation to NH2 * in the first plot in
36
37 Figure 5, computed at 500 K on Pt(211), indicates that reactions with O* contribute much
38
39 more to overall NH3 oxidation than does reaction with OH*. Comparison of diagrams at
40
41 500 and 1000 K reveal the transition from N2 to NO production.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 19
60 ACS Paragon Plus Environment
ACS Catalysis Page 20 of 28

1
2
3
4 Pt (211) Pt (111) Pt (211) Pt (111)
500 K NH3* 500 K NH3* 1000 K NH3* 1000 K NH3*
5
6
OH* O* OH* O* OH* O* OH* O*
7 NH2* NH2* NH2* NH2*
8
9 OH* O* OH* O* OH* O* OH* O*
10 NH* NH* NH* NH*
11
12 OH* O* OH* O* OH* O* OH* O*
13 N* N* N* N*
14 NO* NO* NO* NO*
15
16
17 N2 N2O NO N2 N2O NO N2 N2O NO N2 N2O NO
18
19
20 Figure 5: Steady-state net rates of surface reactions on Pt(211) and (111) at 500 and 1000 K,
21 presented as a Sankey diagram, in which each segment indicates an elementary step and the
22 corresponding segment width indicates net rate normalized to the overall ammonia oxidation
23 rate. The partial pressures of NH3 , O2 and H2 O are 0.001, 0.02 and 0.05 bar, respectively.
24
25
26
27 The Sankey diagrams show that path from reactant to products is sensitive both to
28
29 catalyst structure and temperature. Furthermore, these results show that the relative con-
30
31 tributions of pathways to overall rates cannot be inferred from relative activation energies.
32
33 Figure 1 shows that O*-assisted pathways always have lower barrier on Pt(211) and OH*-
34
35 assisted pathway lower on Pt(111). Nonetheless, at 500 K, where the product is exclusively
36
37 N2 , OH*-assisted steps contribute more to the net rate of NH2 * activation on Pt(211) than
38
39 does the O*-assisted pathway. On Pt(111), O*-assisted activation contributes more than
40
41 OH*-assisted pathway for activating NH3 *. At 1000 K, where the product is NO, the rel-
42
43 ative contributions of O*- and OH*-assisted NHx * activation to the rate are different than
44
45 those at 500 K. The contribution of O* increases with T , likely resulted from the high
46
47 abundance of surface O and negligible surface OH at elevated T (Figure S11), so that the
48
49 way a reaction traverses the reaction network is a function both of intrinsic rates, as reflected
50
51 in reaction barriers, and in the statistical availability of reactants, as determined by surface
52
53 coverages.
54
55
56
57
58
59 20
60 ACS Paragon Plus Environment
Page 21 of 28 ACS Catalysis

1
2
3
4
Conclusions
5
6 Selectivity is central to catalytic function. The catalytic oxidation of ammonia illustrates
7
8 selectivity in multiple guises: the conventional selectivity between competing products, the
9
10 “selectivity” of reaction networks for some sites over others (a component of structure sen-
11
12 sitivity), and the microscopic selectivity between parallel reaction paths. Here we analyze
13
14 how these various selectivites interconnect with one another.
15
16 With DFT computation and microkinetic modeling, we compare the reactivity and selec-
17
18 tivity of step and terrace Pt for ammonia oxidation over a wide range of conditions relevant
19
20 to ammonia slip control and the Ostwald process. We conclude that the rate of ammonia
21
22 oxidation is structure-sensitive at low T , but the sensitivity diminishes at high T . Maxi-
23
24 mizing step sites will improve ammonia slip activity at low T . The DRC analysis suggests
25
26 that rates are controlled by only a few elementary reactions. At high T , ammonia activation
27
28 by O* limits the overall rate on both step and terrace Pt. Catalysts with low barrier for
29
30 ammonia activation will likely increase the overall rate at high T .
31
32 Reaction selectivity is kinetically controlled. We show the temperature corresponding to
33
34 product switch from N2 to NO is higher on Pt(211) than (111), suggesting more step sites
35
36 will likely improve N2 selectivity in ammonia slip catalysis and more terrace sites improve
37
38 the NO selectivity in the Ostwald process. In terms of by-product N2 O, Pt(111) produces
39
40 negligible amount of N2 O at both ammonia slip catalysis and Ostwald conditions but Pt(211)
41
42 produces small amount of N2 O at intermediate temperatures. The selectivity to N2 O is low
43
44 (1.5-2.5%) in industrial Ostwald process with Pt gauze, but nitric acid plants are still the
45
46 largest chemical source of N2 O, 54 which is both a greenhouse gas and a stratospheric ozone
47
48 depletor. 55,56 Pt with more terrace sites may be beneficial to reduce the reaction selectivity
49
50 to N2 O in the Ostwald process.
51
52
53
54
55
56
57
58
59 21
60 ACS Paragon Plus Environment
ACS Catalysis Page 22 of 28

1
2
3
4
Supporting Information Available
5
6 The following files are available free of charge on the ACS Publications website.
7
8
9 • Free energies of ammonia oxidation, binding strength of reaction intermediates, initial,
10
11 transition and final states of reaction steps, energy diagrams for O2 dissociative adsorp-
12
13 tion and OH* reaction, reaction energies and barriers of ammonia oxidation on Pt (211)
14
15 and Pt(111), imaginary frequencies of transition states and prefactors of the reactions,
16
17 surface coverages of Pt(211) and Pt(111), rate and selectivity temperature-dependence
18
19 in a CSTR, rate at low pressure, rate and selectivity temperature-dependence in with
20
21 *-assisted dehydrogenation of NH3 , and rate and selectivity sensitivity to sticking co-
22
23 efficient. (PDF)
24
25
26
27 Acknowledgments
28
29
30 This research was supported by the U.S. Department of Energy, Office of Science, Office
31
32 of Basic Energy Sciences, under Award DE-FG02-06ER15830. The computing resources
33
34 and technical support for this work were provided by the Notre Dame Center for Research
35
36 Computing.
37
38
39
40 References
41
42
43 (1) Chen, J. G.; Crooks, R. M.; Seefeldt, L. C.; Bren, K. L.; Bullock, R. M.; Darens-
44
45 bourg, M. Y.; Holland, P. L.; Hoffman, B.; Janik, M. J.; Jones, A. K.; Kanatzidis, M. G.;
46
47 King, P.; Lancaster, K. M.; Lymar, S. V.; Pfromm, P.; Schneider, W. F.; Schrock, R. R.
48
49 Beyond Fossil Fuel–Driven Nitrogen Transformations. Science 2018, 360, eaar6611.
50
51
52 (2) Bartholomew, C. H.; Farrauto, R. J. Fundamentals of Industrial Catalytic Processes:
53
54 2nd Ed; John Wiley & Sons: Hoboken, 2006; pp 560–634.
55
56
57
58
59 22
60 ACS Paragon Plus Environment
Page 23 of 28 ACS Catalysis

1
2
3
(3) Chmielarz, L.; Jabłońska, M. Advances in Selective Catalytic Oxidation of Ammonia
4
5
to Dinitrogen: A Review. RSC Adv. 2015, 5, 43408–43431.
6
7
8 (4) Handforth, S.; Tilley, J. Catalysts for Oxidation of Ammonia to Oxides of Nitrogen.
9
10 Ind. Eng. Chem. 1934, 26, 1287–1292.
11
12
13 (5) Jabłońska, M.; Palkovits, R. Copper Based Catalysts for the Selective Ammonia Ox-
14
15 idation into Nitrogen and Water Vapour-Recent Trends and Open Challenges. Appl.
16
17 Catal., B 2016, 181 .
18
19
20 (6) Shrestha, S.; Harold, M. P.; Kamasamudram, K.; Kumar, A.; Olsson, L.; Leist-
21
22 ner, K. Selective Oxidation of Ammonia to Nitrogen on Bi-Functional Cu-SSZ-13 and
23
24 Pt/Al2O3 Monolith Catalyst. Catal. Today 2016, 267, 130–144.
25
26
27 (7) Gland, J. L.; Woodard, G. C.; Korchak, V. N. Ammonia Oxidation on the Pt(111) and
28
29 Pt(S)-12(111) × (111) Surfaces. J. Catal. 1980, 61, 543–546.
30
31
(8) Zeng, Y. F.; Imbihl, R. Structure Sensitivity of Ammonia Oxidation over Platinum. J.
32
33
Catal. 2009, 261, 129–136.
34
35
36 (9) Chmielarz, L.; Jabłońska, M.; Strumiński, A.; Piwowarska, Z.; Wegrzyn, A.;
37
38 Witkowski, S.; Michalik, M. Selective Catalytic Oxidation of Ammonia to Nitrogen
39
40 over Mg-Al, Cu-Mg-Al and Fe-Mg-Al Mixed Metal Oxides Doped with Noble Metals.
41
42 Appl. Catal., B 2013, 130-131, 152–162.
43
44
45 (10) Gang, L.; Anderson, B.; van Grondelle, J.; van Santen, R. NH3 Oxidation to Nitrogen
46
47 and Water at Low Temperatures Using Supported Transition Metal Catalysts. Catal.
48
49 Today 2000, 61, 179–185.
50
51
52 (11) Li, Y.; Armor, J. N. Selective NH3 Oxidation to N2 in a Wet Stream. Appl. Catal., B
53
54 1997, 13, 131–139.
55
56
57
58
59 23
60 ACS Paragon Plus Environment
ACS Catalysis Page 24 of 28

1
2
3
(12) Il’chenko, N. I. Catalytic Oxidation of Ammonia. Russ. Chem. Rev. 1976, 45, 1119–
4
5
1134.
6
7
8 (13) Ostermaier, J. J.; Katzer, J. R.; Manogue, W. H. Crystallite Size Effects in the Low-
9
10 Temperature Oxidation of Ammonia over Supported Platinum. J. Catal. 1974, 33,
11
12 457–473.
13
14
15 (14) Offermans, W.; Jansen, A.; van Santen, R. Ammonia Activation on Platinum {111}:
16
17 A Density Functional Theory Study. Surf. Sci. 2006, 600, 1714–1734.
18
19
20 (15) Novell-Leruth, G.; Ricart, J. M.; Pérez-Ramírez, J. Pt(100)-Catalyzed Ammonia Oxi-
21
22 dation Studied by DFT: Mechanism and Microkinetics. J. Phys. Chem. C 2008, 112,
23
24 13554–13562.
25
26
27
(16) Imbihl, R.; Scheibe, A.; Zeng, Y.; Günther, S.; Kraehnert, R.; Kondratenko, V.;
28
29
Baerns, M.; Offermans, W.; Jansen, A.; Van Santen, R. Catalytic Ammonia Oxida-
30
31
tion on Platinum: Mechanism and Catalyst Restructuring at High and Low Pressure.
32
33
Phys. Chem. Chem. Phys. 2007, 9, 3522–3540.
34
35 (17) Rafti, M.; Lovis, F.; Zeng, Y.; Imbihl, R. Homogeneous and Front-Induced Surface
36
37 Transformations during Catalytic Oxidation of Ammonia over Pt (1 0 0). Chem. Phys.
38
39 Lett. 2007, 446, 323–328.
40
41
42 (18) Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals
43
44 and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci 1996, 6, 15–50.
45
46
47 (19) Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Cal-
48
49 culations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186.
50
51
52 (20) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979.
53
54
(21) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-
55
56
Wave Method. Phys. Rev. B 1999, 59, 1758–1775.
57
58
59 24
60 ACS Paragon Plus Environment
Page 25 of 28 ACS Catalysis

1
2
3
(22) Perdew, J.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Sim-
4
5
ple. Phys. Rev. Lett. 1996, 77, 3865–3868.
6
7
8 (23) Henkelman, G.; Uberuaga, B. P.; Jónsson, H. Climbing Image Nudged Elastic Band
9
10 Method for Finding Saddle Points and Minimum Energy Paths. J. Chem. Phys. 2000,
11
12 113, 9901–9904.
13
14
15 (24) Birch, F. Finite Elastic Strain of Cubic Crystals. Phys. Rev. 1947, 71, 809–824.
16
17
18 (25) Murnaghan, F. D. The Compressibility of Media under Extreme Pressures. Proc. Natl.
19
20 Acad. Sci. U.S.A. 1944, 30, 244–247.
21
22
23 (26) Hull, A. W. X-Ray Crystal Analysis of Thirteen Common Metals. Phys. Rev. 1921,
24
25 17, 571–588.
26
27
28
(27) Offermans, W.; Jansen, A.; Van Santen, R.; Novell-Leruth, G.; Ricart, J.; Perez-
29
Ramirez, J. Ammonia Dissociation on Pt {100}, Pt {111}, and Pt {211}: A Com-
30
31
parative Density Functional Theory Study. J. Phys. Chem. C 2007, 111, 17551–17557.
32
33
34 (28) Monkhorst, H.; Pack, J. Special Points for Brillouin Zone Integrations. Phys. Rev. B
35
36 1976, 13, 5188–5192.
37
38
39 (29) Foppa, L.; Silaghi, M.-C.; Larmier, K.; Comas-Vives, A. Intrinsic Reactivity of Ni,
40
41 Pd and Pt Surfaces in Dry Reforming and Competitive Reactions: Insights from First
42
43 Principles Calculations and Microkinetic Modeling Simulations. J. Catal. 2016, 343,
44
45 196–207.
46
47
48 (30) Dumesic, J. The Microkinetics of Heterogeneous Catalysis; ACS Professional Reference
49
50 Book; American Chemical Society: Washington, 1993; pp 23–54.
51
52
53 (31) Chase Jr, M. W. NIST-JANAF Thermochemical Tables. J. Phys. Chem. Ref. Data,
54
55 Monograph 1998, 9 .
56
57
58
59 25
60 ACS Paragon Plus Environment
ACS Catalysis Page 26 of 28

1
2
3
(32) Grabow, L. C. In Computational Catalysis; Asthagiri, A., Janik, M. J., Eds.; Royal
4
5
Society of Chemistry: London, 2013; pp 1–58.
6
7
8 (33) Mieher, W. D.; Ho, W. Thermally Activated Oxidation of NHs on Pt(111): Intermediate
9
10 Species and Reaction Mechanisms. Surf. Sci. 1995, 322, 151–167.
11
12
13 (34) Weststrate, C. J.; Bakker, J. W.; Gluhoi, A. C.; Ludwig, W.; Nieuwenhuys, B. E.
14
15 Ammonia Oxidation on Ir (111): Why Ir is More Selective to N2 than Pt. Catal. Today
16
17 2010, 154, 46–52.
18
19
20 (35) Popa, C.; Van Santen, R. A.; Jansen, A. P. J. Density-Functional Theory Study of NHx
21
22 Oxidation and Reverse Reactions on the Rh(111) Surface. J. Phys. Chem. C 2007, 111,
23
24 9839–9852.
25
26
27
(36) López, N.; García-Mota, M.; Gómez-Díaz, J. NH3 Oxidation on Oxygen-Precovered
28
29
Au(111): A Density Functional Theory Study on Selectivity. J. Phys. Chem. C 2008,
30
31
112, 247–252.
32
33 (37) Jiang, Z.; Qin, P.; Fang, T. Mechanism of Ammonia Decomposition on Clean and
34
35 Oxygen-Covered Cu (1 1 1) Surface: A DFT Study. Chem. Phys. 2014, 445, 59–67.
36
37
38 (38) Clay, J. P.; Greeley, J. P.; Ribeiro, F. H.; Delgass, W. N.; Schneider, W. F. DFT
39
40 Comparison of Intrinsic WGS Kinetics over Pd and Pt. J. Catal. 2014, 320, 106–117.
41
42
43 (39) Getman, R. B.; Schneider, W. F. DFT-Based Characterization of the Multiple Adsorp-
44
45 tion Modes of Nitrogen Oxides on Pt (111). J. Phys. Chem. C 2007, 111, 389–397.
46
47
48 (40) Liu, Z.-P.; Jenkins, S. J.; King, D. A. Step-Enhanced Selectivity of NO Reduction on
49
50 Platinum-Group Metals. J. Am. Chem. Soc. 2003, 125, 14660–14661.
51
52
(41) Gonzalez, J. D.; Shojaee, K.; Haynes, B. S.; Montoya, A. The Effect of Surface Coverage
53
54
on N2, NO and N2O Formation over Pt (111). Phys. Chem. Chem. Phys. 2018, 20,
55
56
25314–25323.
57
58
59 26
60 ACS Paragon Plus Environment
Page 27 of 28 ACS Catalysis

1
2
3
(42) Bray, J.; Schneider, W. First-Principles Analysis of Structure Sensitivity in NO Oxida-
4
5
tion on Pt. ACS Catal. 2015, 5, 1087–1099.
6
7
8 (43) Gambardella, P.; Šljivančanin, Ž.; Hammer, B.; Blanc, M.; Kuhnke, K.; Kern, K. Oxy-
9
10 gen Dissociation at Pt Steps. Phys. Rev. Lett. 2001, 87, 056103.
11
12
13 (44) Grabow, L. C.; Gokhale, A. A.; Evans, S. T.; Dumesic, J. A.; Mavrikakis, M. Mechanism
14
15 of the Water Gas Shift Reaction on Pt: First Principles, Experiments, and Microkinetic
16
17 Modeling. J. Phys. Chem. C 2008, 112, 4608–4617.
18
19
20 (45) Stamatakis, M.; Chen, Y.; Vlachos, D. G. First-Principles-Based Kinetic Monte Carlo
21
22 Simulation of the Structure Sensitivity of the Water–Gas Shift Reaction on Platinum
23
24 Surfaces. J. Phys. Chem. C 2011, 115, 24750–24762.
25
26
27 (46) Nørskov, J. K.; Studt, F.; Abild-Pedersen, F.; Bligaard, T. Fundamental Concepts in
28
29 Heterogeneous Catalysis; John Wiley & Sons, 2014.
30
31
(47) Yoo, J. S.; Schumann, J.; Studt, F.; Abild-Pedersen, F.; Nørskov, J. K. Theoretical
32
33
Investigation of Methane Oxidation on Pd (111) and Other Metallic Surfaces. J. Phys.
34
35
Chem. C 2018, 122, 16023–16032.
36
37
38 (48) Baerns, M.; Imbihl, R.; Kondratenko, V. A.; Kraehnert, R.; Offermans, W. K.; Van
39
40 Santen, R. A.; Scheibe, A. Bridging the Pressure and Material Gap in the Catalytic Am-
41
42 monia Oxidation: Structural and Catalytic Properties of Different Platinum Catalysts.
43
44 J. Catal. 2005, 232, 226–238.
45
46
47 (49) Rafti, M.; Vicente, J. L.; Albesa, A.; Scheibe, A.; Imbihl, R. Modeling Ammonia Oxi-
48
49 dation over a Pt (533) Surface. Surf. Sci. 2012, 606, 12–20.
50
51
52 (50) Scheibe, A.; Hinz, M.; Imbihl, R. Kinetics of Ammonia Oxidation on Stepped Platinum
53
54 Surfaces: II. Simulation Results. Surf. Sci. 2005, 576, 131–144.
55
56
57
58
59 27
60 ACS Paragon Plus Environment
ACS Catalysis Page 28 of 28

1
2
3
(51) Rebrov, E.; De Croon, M.; Schouten, J. Development of the Kinetic Model of Platinum
4
5
Catalyzed Ammonia Oxidation in a Microreactor. Chem. Eng. J. 2002, 90, 61–76.
6
7
8 (52) Phatak, A. A.; Delgass, W. N.; Ribeiro, F. H.; Schneider, W. F. Density Functional
9
10 Theory Comparison of Water Dissociation Steps on Cu, Au, Ni, Pd, and Pt. J. Phys.
11
12 Chem. C 2009, 113, 7269–7276.
13
14
15 (53) Campbell, C. T. The Degree of Rate Control: A Powerful Tool for Catalysis Research.
16
17 ACS Catal. 2017, 7, 2770–2779.
18
19
20 (54) Pérez-Ramırez, J.; Kapteijn, F.; Schöffel, K.; Moulijn, J. Formation and Control of
21
22 N2O in Nitric Acid Production: Where Do We Stand Today? Appl. Catal., B 2003,
23
24 44, 117–151.
25
26
27 (55) Leuenberger, M.; Siegenthaler, U. Ice-Age Atmospheric Concentration of Nitrous Oxide
28
29 from An Antarctic Ice Core. Nature 1992, 360, 449.
30
31
(56) Ravishankara, A.; Daniel, J. S.; Portmann, R. W. Nitrous Oxide (N2O): the Dominant
32
33
Ozone-Depleting Substance Emitted in the 21st Century. Science 2009, 326, 123–125.
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 28
60 ACS Paragon Plus Environment

You might also like