Subscriber access provided by Angelo and Jennette | Volpe Library & Media Center
Article
Metal-porphyrin: A Potential Catalyst for N2O Direct Decomposition by Theoretical Reaction Mechanism Investigation Phornphimon Maitarad, Supawadee Namuangruk, Dengsong Zhang, Liyi Shi, Hongrui Li, Lei Huang, Bundet Boekfa, and Masahiro Ehara Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 23 May 2014 Downloaded from http://pubs.acs.org on May 25, 2014
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.
Environmental Science & Technology 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 37
Environmental Science & Technology
1
Metal–porphyrin: A Potential Catalyst for Direct
2
Decomposition of N2O by Theoretical Reaction
3
Mechanism Investigation
4
Phornphimon Maitarad,† Supawadee Namuangruk,‡ Dengsong Zhang,†* Liyi Shi,†
5
Hongrui Li,† Lei Huang,† Bundet Boekfa§ and Masahiro Ehara§* †
6
Research Center of Nano Science and Technology, Shanghai University, Shanghai
7
8
200444, P. R. China ‡
National Nanotechnology Center (NANOTEC), NSTDA, 111 Thailand Science Park,
9
10
Pahonyothin Road, Klong Luang, Pathum Thani 12120, Thailand §
Institute for Molecular Science and Research Center for Computational Science, 38
11
Nishigo-naka, Myodaiji, Okazaki, 444-8585, Japan
12
ABSTRACT
13
The adsorption of nitrous oxide (N2O) on metal–porphyrins (metal: Ti, Cr, Fe, Co, Ni,
14
Cu, or Zn), has been theoretically investigated using density functional theory with the
15
M06L functional to explore their use as potential catalysts for the direct decomposition of
16
N2O. Among these metal–porphyrins, Ti–porphyrin is the most active for N2O adsorption
1 ACS Paragon Plus Environment
Environmental Science & Technology
Page 2 of 37
17
in the triplet ground state with the strongest adsorption energy (–13.32 kcal/mol). Ti–
18
porphyrin was then assessed for the direct decomposition of N2O. For the overall reaction
19
mechanism of three N2O molecules on Ti–porphyrin, two plausible catalytic cycles are
20
proposed. Cycle 1 involves the consecutive decomposition of the first two N2O molecules,
21
while cycle 2 is the decomposition of the third N2O molecule. For cycle 1, the activation
22
energies of the first and second N2O decompositions are computed to be 3.77 and 49.99
23
kcal/mol, respectively. The activation energy for the third N2O decomposition in cycle 2
24
is 47.79 kcal/mol, which is slightly lower than that of the second activation energy of the
25
first cycle. O2 molecules are released in cycles 1 and 2 as the products of the reaction,
26
which requires endothermic energies of 102.96 and 3.63 kcal/mol, respectively.
27
Therefore, the O2 desorption is mainly released in catalytic cycle 2 of a TiO3–porphyrin
28
intermediate catalyst. In conclusion, regarding the O2 desorption step for the direct
29
decomposition of N2O, the findings would be very useful to guide the search for potential
30
N2O decomposition catalysts in new directions.
31
Keywords: Porphyrin; N2O decomposition; Density Functional Theory; Reaction
32
mechanism
33
2 ACS Paragon Plus Environment
Page 3 of 37
Environmental Science & Technology
34
1. INTRODUCTION
35
Nitrous oxide (N2O) emitted from industrial processes and vehicle engines is an
36
environmentally polluting gas. Moreover, it has been recognized as a strong greenhouse
37
gas contributing to the destruction of ozone in the stratosphere.1–3 Because of the
38
continuous increase in its concentration in the atmosphere, the discovery and
39
development of efficient catalysts for the reduction of N2O have become important issues
40
in the field of environmental research. A direct catalytic decomposition of N2O into N2
41
and O2 is thought to be the most convenient and economical option to reduce N2O
42
emissions.4 Therefore, several catalytic materials have been reported for direct
43
decomposition of N2O gas, such as, M-zeolites (M = Cu, Co, Fe, etc.),5–9 carbon
44
nanotubes,10, 11 perovskite-like mixed oxides,12–14 alumina-supported precious metals (Pd,
45
Rh, etc.),15, 16 and metal alloys.17 Among these catalysts, transition-metal ion-exchanged
46
ZSM-5 zeolites, especially for the Fe-, Co-, and Cu-exchanged zeolites, have been widely
47
used because they show high catalytic activity.9,
48
from oxygen inhibition and the low reaction rate of the N2O decomposition.
49 50
18–23
However, these catalysts suffer
In general, the full reaction mechanism of direct N2O decomposition has been reported as the following:24–28
51
[Cat] + N2O(g)
→
[Cat]…ON2
52
[Cat]…ON2
→
[Cat]O + N2(g)
53
[Cat]O + N2O(g) →
[Cat]O…ON2
54
[Cat]O…ON2
→
[Cat]O2 + N2(g)
55
[Cat]O2
→
[Cat] + O2(g)
3 ACS Paragon Plus Environment
Environmental Science & Technology
Page 4 of 37
56
Both theoretical and experimental studies have reported that the recombination of
57
two oxygen atoms into the O2 molecule or the O2 desorption is the rate-limiting step in
58
the overall direct N2O decomposition reaction.9,
59
catalysts for direct N2O decomposition will be accelerated by the preliminary study of the
60
energetics along the reaction pathway by using, for example, density functional theory
61
(DFT) calculations prior to the experiment.
25, 28
Thus, the development of new
62
Porphyrin readily forms ordered monolayers by self-assembly and possesses two
63
axial coordination sites that are available as centers of catalytic activity or sensor
64
functionality. Metal–porphyrins are well suited for anchoring on solid substrates as
65
assemblies in various types of applications, such as photovoltaic materials, field-
66
responsive materials, catalytic materials, etc.29–33 It has been generally recognized that the
67
metal–porphyrins are highly active toward oxygen, nitric oxide, carbon monoxide, etc.34–
68
39
69
which the coordination of gas molecules to the metal center causes measurable changes
70
of electronic properties, color, etc.30,
71
synthesized as microporous solid frameworks that have a selective sorption of small
72
molecules and size- or shape-selective heterogeneous catalysis.44–48 Metal–porphyrins
73
have been considered as potential catalysts for many commercially important reactions
74
such as the reduction of carbon dioxide49–51 and nitrogen oxides,52–54 and the oxidation of
75
hydrocarbons and alcohols.55–58
Therefore, one important application of metal–porphyrins is the sensing of gases, in
40–43
In addition, the metal–porphyrins can be
76
Although there are some reports on the ability of metal–porphyrins with respect to
77
N2O adsorption,35, 59 their catalytic activity for direct N2O decomposition has not yet been
78
demonstrated. Therefore, this motivated us to apply DFT calculations to the N2O
4 ACS Paragon Plus Environment
Page 5 of 37
Environmental Science & Technology
79
adsorption ability of metal–porphyrins, where the metal is Ti, Cr, Fe, Co, Ni, Cu, or Zn,
80
to find potential catalysts for the N2O decomposition. The DFT calculations suggest that
81
among these metal–porphyrins, Ti–porphyrin is the most active to N2O adsorption in its
82
triplet ground state. Therefore, the possible direct N2O decomposition over a Ti–
83
porphyrin catalyst has been examined and three N2O decomposition mechanisms have
84
been proposed. O2 desorption, which is a key step for the N2O decomposition, has also
85
been considered to determine a favorable pathway. The characteristics and performance
86
of the present Ti–porphyrin are compared with those of the potential catalysts of Fe-, Co-,
87
and Cu-ZSM-5 zeolites.9, 19, 28
88 89
2. METHODS
90
As the model of catalysts, single metal–porphyrins in which the metal is Ti, Cr, Fe, Co,
91
Ni, Cu, or Zn, were examined without a support or assembly. Neutral metal–porphyrins,
92
core system of the catalyst, were assessed in some spin states without considering the
93
charge transfer from the support or surroundings when locating the energetically lowest
94
geometric structure. The most stable spin state of each metal–porphyrin was selected to
95
examine the N2O adsorption ability. The spin state of the adsorption complex (N2O–
96
metal–porphyrin) was set to the same spin state as the metal–porphyrin without
97
considering spin crossing. For the initial structure of the geometry optimization, the N2O
98
molecule was laid along the perpendicular axis to the plane of the porphyrin ring with the
99
oxygen atom of N2O pointed to the metal atom of the metal–porphyrin (Figure 1). The
100 101
adsorption energy (Ead) of N2O was calculated by Ead = Ecomplex – (EM–por + EN2O),
(1)
5 ACS Paragon Plus Environment
Environmental Science & Technology
Page 6 of 37
102
where Ecomplex, EM–por, and EN2O are the total energies of the metal–porphyrin···N2O
103
complex, metal–porphyrin, and N2O molecule, respectively.
104
The direct N2O decomposition over Ti–porphyrin was examined. During the
105
geometry optimizations, all atoms of the Ti–porphyrin complexes were fully relaxed. The
106
transition states were confirmed as the real saddle points with only one imaginary
107
frequency by vibrational analysis. The reaction energy profiles of each step were
108
presented in the relative energy, which is defined as,
109
∆E = Ecomplex – (Ecatalyst + Eabsorbate),
(2)
110
where Ecomplex, Ecatalyst, and Eabsorbate are the energies of the Ti–porphyrin–gas complexes,
111
the Ti–porphyrin at each step, and the small gas molecules, e.g., N2O, O2, and N2. For the
112
reaction pathway, we examined the possibility of the spin crossing or intersystem
113
crossing and found that the singlet state of some intermediates and transition states were
114
more stable than their triplet counterparts.
115
All of the electronic structure calculations and the geometry optimizations were
116
performed using the DFT method with the M06L functional60 without any restriction on
117
the symmetry. The 6-31G* basis set61 was used for the C, O, N, and H atoms and the
118
scalar relativistic effective core potential of LANL2DZ62 was adopted for the transition
119
metal elements. All calculations in the present work were carried out using the
120
Gaussian09 suite of programs, revision B01.63 The charge analyses of the systems were
121
performed using the natural bond orbital (NBO) analysis.64
122 123
3. RESULTS AND DISCUSSION
6 ACS Paragon Plus Environment
Page 7 of 37
Environmental Science & Technology
124
3.1. N2O adsorption over M (Ti, Cr, Fe, Co, Ni, Cu, or Zn) – porphyrins. The
125
adsorption energy of N2O, partial charges, and selected structural parameters for the Ti–,
126
Cr–, Fe–, Co–, Ni–, Cu–, and Zn–porphyrins are summarized in Table 1. The relative
127
energies of the metal–porphyrins with various spin multiplicities are compared in Table
128
S1. Based on the unrestricted DFT calculations, the Ni– and Zn–porphyrins are found to
129
have a closed-shell singlet ground state, and the Co– and Cu–porphyrins prefer the low-
130
spin doublet state. The Ti– and Fe–porphyrins, on the other hand, result in the triplet
131
ground state, whereas the high-spin quintet state is the lowest state for the Cr–porphyrin.
132
The metal–nitrogen (M–Npor) bond distance of the metal–porphyrin is about 2 Å and the
133
spin state is correlated to the M–Npor bond distance, reflecting the coordination field of
134
porphyrin. All metal–porphyrins were found to be planar as represented by a ∠Npor–M–
135
Npor angle of 180°. The present results are consistent with the previous works.65, 66
136
Comparing the structures of the metal–porphyrins and the N2O-adsorbed complexes,
137
the adsorption of N2O over the metal–porphyrins does not affect the M–Npor distances in
138
almost all systems except for Ti–porphyrin, which shows slight structure reorganization
139
(∠Npor–M–Npor = 172°). Ti–porphyrin provided a considerable adsorption energy of –
140
13.32 kcal/mol, while other metal porphyrins exhibited low adsorption energies in the
141
range of –4.6 to –7.8 kcal/mol. We also examined the adsorption of the N-binding mode
142
of N2O (see Table S2) and obtained the adsorption energy of –22.27 kcal/mol, which
143
indicates that N-binding adsorption also competes with O-binding over Ti–porphyrin. In
144
the present work, we focus on the O-binding adsorption, as in other works,9, 19, 28 because
145
TiO–porphyrin forms a very stable intermediate, as shown later. In addition, an adsorbed
146
distance between Ti and the O1 atom of the N2O molecule was 2.25 Å, which is the
7 ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 37
147
shortest metal–oxygen (M–O1) distance among the complexes. The adsorption M–O1
148
distance depends on both the M–O1 interaction and the metal radius. Because of its
149
relatively strong interaction, the N2O molecule over the Ti–porphyrin catalyst resulted in
150
the most elongated O1–N2 bond and the shortest N2–N3 bond, as compared with N2O
151
over the other porphyrins.
152
The NBO partial charges of the metal–porphyrins and N2O···metal–porphyrin
153
complexes are also listed in Table 1. The charge difference (qdiff) between N2O gas (q =
154
0) and adsorbed N2O shows the amount of charge reorganization or transfer from N2O to
155
the metal–porphyrins upon N2O adsorption. We found that qdiff increases in the order Ni–
156
(0.02) < Cu– (0.03) < Zn–, Cr–, Fe– (0.04) < Co– (0.05) < Ti–porphyrin (0.12). This
157
charge reorganization contributes to the adsorption strength. Therefore, the Ti–porphyrin
158
was chosen to examine its catalytic activity of the direct N2O decomposition in the
159
following sections.
160 161
3.2. N2O decomposition over Ti–porphyrin. For the direct N2O decomposition, the
162
reaction mechanism was investigated over Ti–porphyrin, which exhibited the strongest
163
adsorption of N2O. The assessment of the reaction pathways follows the structures and
164
the spin states of the N2O···Ti–porphyrin complex. The reaction mechanism of direct
165
decomposition of three N2O molecules can be decomposed into four elementary reaction
166
steps as follows:
167
Step 1 [Ti] + 1st N2O(g) → [Ti]···ON2 → [Ti]O + N2(g),
168
Step 2 [Ti]O + 2nd N2O(g) → [Ti]O···ON2 → [Ti]O2 + N2(g) → [Ti] + O2(g),
169
Step 3 [Ti]O2 + 3rd N2O(g) → [Ti]O2···ON2 → [Ti]O3 + N2(g),
8 ACS Paragon Plus Environment
Page 9 of 37
170
Environmental Science & Technology
Step 4 [Ti]O3 → [Ti]O···O2 → [Ti]O + O2(g).
171
Following the previous reports on the full direct N2O decomposition mechanism,9, 24–28
172
the present system, namely, N2O decomposition over Ti–porphyrin, is schematized in
173
Figure 2. The products of the N2O decomposition are N2 and O2 molecules. We assume
174
that the overall reaction mechanism consists of two cycles. Cycle 1 has two steps: steps 1
175
and 2 (see equations above), which involve the consecutive decomposition of the first and
176
second N2O molecules, respectively. Cycle 2 consists of the third N2O decomposition
177
(step 3) and the depletion of O2 from the TiO3–porphyrin intermediate (step 4). The
178
calculated total energy of all of the intermediates and transition states are listed in Table
179
S3, and the coordinates of the transition states are given in SI.
180 181
3.2.1. The first N2O decomposition over Ti–porphyrin (Step 1). In step 1, the reaction
182
energy profile of the first N2O molecule over Ti–porphyrin consists of N2O adsorption,
183
N–O bond scission, and N2 desorption (Figure 3). At the adsorption complex (AD1), the
184
N2O molecule is located on the Ti atom in a tilted structure with a ∠Ti–O–N angle of
185
about 120°. The Ti···O distance is 2.25 Å and the O–N1 bond slightly elongates from
186
1.19 to 1.20 Å. The calculated adsorption energy (Ead1) is –13.32 kcal/mol. At the
187
transition state (TS1), the ∠Ti–O–N angle changes from 120° to 156° and the Ti–O bond
188
is contracted to 2.06 Å. The first transition state (TS1) has a relative energy of –9.55
189
kcal/mol with one imaginary frequency of 418i cm–1, which corresponds to the reaction
190
coordinate. The Ti active site abstracts the O1 atom from the absorbed N2O to form the
191
TiO–porphyrin intermediate and the N2 molecule (IM1). Spin crossing occurs between
192
TS1 and IM1, assuming that the spin–orbit interaction is large enough. Otherwise, the
9 ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 37
193
reaction may proceed in the triplet state, as shown in the supporting information (Figure
194
S1). This step proceeds in a highly exothermic manner with the relative energy of IM1
195
being –116.53 kcal/mol, indicating that IM1 is very stable. Finally, the N2 molecule is
196
easily desorbed from the TiO–porphyrin intermediate (IM2), which requires an energy of
197
only 3.73 kcal/mol. Therefore, in the first N2O decomposition process, the N2O molecule
198
readily adsorbs onto the Ti–porphyrin, the N–O bond scission has a low activation energy
199
barrier (Ea1) of 3.77 kcal/mol, and the TiO–porphyrin is a thermodynamically stable
200
intermediate.
201 202
3.2.2. The second N2O decomposition over TiO–porphyrin (Step 2). For the second N2O
203
decomposition (Figure 4), although all steps are assumed to be similar to the first N2O
204
decomposition, the active site is based on the TiO–porphyrin intermediate instead of Ti–
205
porphyrin. The second N2O molecule adsorbs over the TiO–porphyrin with the
206
adsorption energy (Ead2) of –2.36 kcal/mol. It is seen that the second N2O molecule
207
weakly adsorbs on the TiO–porphyrin relative to the first one. Regarding the adsorption
208
geometry (AD2), N2O retains a linear structure with unchanged bond distances. The
209
transition state (TS2) was confirmed with an imaginary frequency of 838i cm–1
210
representing the reaction coordinate. At TS2, the O2 atom of N2O approaches the TiO–
211
porphyrin, with the Ti···O2 distance being 2.05 Å. The adsorbed N2O shows a bent
212
structure with ∠O2–N1–N2 of 148° and undergoes the predominant dissociation of the
213
N1–O2 bond. This step requires the activation barrier energy (Ea2) of 49.99 kcal/mol,
214
which is much higher than the first N2O decomposition (Ea1). In the N2 generation
215
(IM21), the O atom from the N2O molecule is abstracted by the TiO–porphyrin to form
10 ACS Paragon Plus Environment
Page 11 of 37
Environmental Science & Technology
216
the TiO2–porphyrin intermediate. The Ti–O bond and O1···O2 distance are 1.83 Å and
217
1.44 Å, respectively. The calculated relative energy is –3.65 kcal/mol. The N2 molecule
218
simultaneously desorbs from the TiO2–porphyrin intermediate (IM22) with desorption
219
energy of 3.70 kcal/mol. It is possible that the O2 molecule is generated in this step.
220
However, it was found that the O2 desorption energy (Ede1) is very high (102.96 kcal/mol)
221
(IM23), implying that the TiO2–porphyrin would not easily release the O2 molecule.
222
Thus, an alternative route for the TiO2–porphyrin catalytic intermediate is considered,
223
and it could decompose another N2O molecule.
224 225
3.2.3. The third N2O decomposition over TiO2–porphyrin (Step 3). The energetic profile
226
for the third N2O decomposition over the TiO2–porphyrin intermediate (IM22) was
227
calculated (Figure 5). The adsorption energy (Ead3) of the third N2O is –5.22 kcal/mol
228
(AD3), which is slightly stronger than that of the second N2O adsorption (–2.36
229
kcal/mol). Reflecting the weak adsorption, the bond distance of the TiO2–porphyrin
230
catalyst and N2O molecule is nearly unaltered at the adsorption step. Thus, to activate an
231
N2O molecule over the TiO2–porphyrin via TS3, it needs the relatively high activation
232
energy (Ea3) of 47.79 kcal/mol. At TS3 with an imaginary frequency of 747i cm–1, the
233
oxygen atom (O3) of the N2O molecule points to the Ti active site with the interaction
234
distance of Ti···O3 = 2.08 Å. The O1···O2 interaction distance is shortened from 1.44 Å
235
to 1.38 Å, while the Ti–O1 and Ti–O2 bond distances are elongated. At TS3, the N2O
236
molecule is distorted from linear to a bent structure of about ∠O3–N1–N2 = 142°. In
237
addition, the O3–N2 bond is significantly lengthened, which implies that the O3–N2
238
bond dissociates and then O3–Ti or O3–O2 may form a new bond on the catalyst surface.
11 ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 37
239
Thus, the TiO2–porphyrin abstracts the oxygen atom from the N2O molecule, then the N2
240
molecule and TiO3–porphyrin intermediate are produced (IM31). In step 3, the TiO2–
241
porphyrin intermediate shows a stronger N2O adsorption ability than that of the TiO–
242
porphyrin of step 2. The activation energy barrier of the third N2O insertion on the TiO2–
243
porphyrin is also slightly lower than that of the second N2O insertion by about 2
244
kcal/mol. These can imply that the third N2O desorption over the TiO2–porphyrin is a
245
prominent reaction pathway as compared with O2 desorption of the TiO2–porphyrin.
246
Desorption of the N2 molecule is calculated to be 4.72 kcal/mol (IM32), which is slightly
247
higher than those of the first and second N2 molecules. This is because of the less stable
248
nature of the TiO3–porphyrin intermediate.
249 250
3.2.4. Oxygen depletion from TiO3–porphyrin (Step 4). The intermediate obtained via the
251
three N2O decompositions is the TiO3–porphyrin (IM32). The four-membered-ring
252
structure (Ti–O1–O2–O3) of IM32 is very strained; the O2 molecule is therefore easily
253
lost to release the strain (Figure 6). The optimized structure of TiO3–porphyrin shows that
254
the Ti–O1 and Ti–O2 have equal bond lengths of 1.91 Å. At TS4 (imaginary frequency =
255
188i cm–1), the obtained geometry is very similar to the IM32, therefore, it is not
256
surprising that the activation energy (Ea4) of the O2 depletion from IM32 is only 5.97
257
kcal/mol because of the unstable four-membered-ring structure in the TiO3–porphyrin
258
intermediate. Thus, this observation can imply that the TiO3–porphyrin catalytic
259
intermediate would spontaneously generate an O2 molecule. Spin crossing again occurs to
260
produce the stable triplet TiO···O2 intermediate (IM41) located at –59.69 kcal/mol. The
12 ACS Paragon Plus Environment
Page 13 of 37
Environmental Science & Technology
261
product of this step is the TiO–porphyrin (IM2), which is a key intermediate catalyst for
262
the N2O decomposition of cycle 2.
263
The O2 desorption is an important process for the direct N2O decomposition reaction
264
because it is usually the rate-limiting step.9, 25, 26 In the present case, considering the O2
265
desorption (Figure 2), there are two possible intermediates to recombine two oxygen
266
atoms into an O2 molecule. The first one is the TiO2–porphyrin (IM22), an intermediate
267
from the second N2O decomposition. The TiO2–porphyrin needs a very high desorption
268
energy (Ede1) of about 103 kcal/mol to release the O2 molecule; therefore, this process
269
hardly occurs. Another possible route for the O2 desorption exists in step 4. Herein, the
270
O2 desorption energy (Ede2) was calculated to be only 3.63 kcal/mol, which is much lower
271
than that of the O2 desorption in step 2. Therefore, the O2 molecule desorbs from the
272
TiO–porphyrin···O2 intermediate in a spontaneous route that requires a very low
273
endothermic energy.
274
A full reaction mechanism for the direct N2O decomposition over the Ti–porphyrin
275
catalyst is summarized in Figure 7. As mentioned above, there are two catalytic cycles for
276
the three N2O decompositions. Cycle 1 consists of steps 1 and 2 based on Ti–porphyrin
277
recycling catalyst form; this step shows an O2 desorption barrier of about 103 kcal/mol.
278
For steps 2, 3, and 4 in Cycle 2, the TiO–porphyrin is a recycling catalyst for this cycle,
279
and the O2 desorption barrier requires only 3.63 kcal/mol. Therefore, based on the key
280
step of O2 desorption, Cycle 2 is the main catalytic pathway for N2O decomposition over
281
Ti–porphyrin.
282
13 ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 37
283
3.3. Comparison of Ti–porphyrin and zeolites. The comparison of theoretical
284
activation energies and O2 desorption energies of three direct N2O decompositions over
285
the Fe-, Cu-, and Co-ZSM5 zeolites9,
286
summarized in Table 2. For the first N2O decomposition, the Ti–porphyrin and Fe/Co-
287
ZSM-5 zeolites show similar activation energy barriers in the range of 4–6 kcal/mol,
288
which is much lower than that of the Cu-ZSM-5 zeolite. This suggests that the first N2O
289
decomposition over Ti–porphyrin and Fe/Co-ZSM-5 zeolites is easier than that of the Cu-
290
ZSM-5 zeolite. On the other hand, the second N2O decomposition is feasible over the Cu-
291
ZSM-5 zeolite. It is noted that the third N2O decomposition over all catalysts has similar
292
energy barriers, and it is important to mention that Ti–porphyrin results in a slightly
293
higher activation energy. In particular, the activation energy for the third N2O
294
decomposition is higher than that for the first and second decompositions in all catalysts
295
because of the greater steric effect of oxygen atoms bonded to the Ti active site.
296
Considering the rate-limiting step of the O2 desorption in the second and third N2O
297
decompositions, the Cu-ZSM-5 zeolite prefers to desorb the O2 molecule from the second
298
N2O decomposition or in the form of [Cu]-O2-ZSM-5 intermediate catalyst. In contrast,
299
O2 desorption over Ti–porphyrin prefers to release from the third N2O decomposition or
300
in the form of [TiO]O2–porphyrin intermediate catalyst, which corresponds well with the
301
Fe/Co-ZSM-5 zeolite. Significantly, for the O2 desorption from the third N2O
302
decomposition, the present Ti–porphyrin catalyst shows the lowest desorption energy
303
barrier compared with those of O2 desorption over the Fe-, Cu-, and Co-ZSM5 zeolites,
304
which implies that the O2 desorption seems to be a spontaneous pathway over the
305
[TiO]O2–porphyrin intermediate catalyst. Therefore, based on the energy comparison, we
19, 28
and the present Ti–porphyrin catalyst are
14 ACS Paragon Plus Environment
Page 15 of 37
Environmental Science & Technology
306
propose that the Ti–porphyrin catalyst is one of the candidates for the direct
307
decomposition of N2O.
308
Hence, the present theoretical study demonstrates the potential use of Ti–porphyrin
309
as a catalyst for a direct N2O decomposition that comparably enhances the N2O
310
decomposition and effectively produces O2 depletion and desorption compared with the
311
reaction pathways over conventional catalysts such as zeolites. Therefore, the synthesis
312
of this catalyst and reaction kinetics are very interesting, and it is underway in our
313
laboratory. Theoretical analysis of the extended model system of this catalyst is also
314
underway taking account of the surroundings.
315 316
ASSOCIATED CONTENT
317
Supporting Information
318
(1) Total energies of isolated metal–porphyrins based on M06L/6-31G* (C, N, O, and H)
319
LANL2DZ (Ti, Cr, Fe, Co, Ni, Cu, and Zn). (2) The N2O adsorption energies for O- and
320
N-binding modes over the metal–porphyrins. (3) Total energies of the adsorption
321
complexes (AD1, AD2, AD3), intermediates (IM1, IM2, IM21, IM22, IM31, IM32,
322
IM41, IM42), and transition states (TS1, TS2, TS3, TS4). (4) Cartesian coordinates of
323
the transition states. (5) Reaction pathway of N2O decomposition in the singlet and triplet
324
states of Ti–porphyrin. This information is available free of charge via the Internet at
325
http://pubs.acs.org.
326 327
AUTHOR INFORMATION
15 ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 37
328
Corresponding Author
329
*Ph: +86- 21-66136079. E-mail:
[email protected] (D.Z.)
330
*Ph: +81-564-55-7461. E-mail:
[email protected] (M.E.)
331
Notes
332
The authors declare no competing financial interest.
333
ACKNOWLEDGEMENTS
334
The
335
(12R21413300) and National Natural Science Foundation of China (51108258). The
336
work was also supported by Nanotechnology Platform Program (Molecule and Material
337
Synthesis) of the Ministry of Education, Culture, Sports, Science and Technology
338
(MEXT), Japan. We thank Research Center for Computational Science in Okazaki, Japan
339
and National Nanotechnology Center (NANOTEC) in Thailand for computing resources.
authors
acknowledge
the
supports
of
STCSM
postdoctoral
foundation
340 341
REFERENCES
342
1.
343
dominant ozone-depleting substance emitted in the 21st century. Science 2009, 326,
344
(5949), 123-125.
345
2.
346
tropical forests. Nature 1999, 400, 152-155.
Ravishankara, A. R.; Daniel, J. S.; Portmann, R. W., Nitrous oxide (N2O): the
Hall, S. J.; Matson, P. A., Nitrogen oxide emissions after nitrogen additions in
16 ACS Paragon Plus Environment
Page 17 of 37
Environmental Science & Technology
347
3.
Becker, K. H.; Lörzer, J. C.; Kurtenbach, R.; Wiesen, P., Nitrous oxide (N2O)
348
emissions from vehicles. Environ. Sci. Technol. 1999, 33, 4134-4139.
349
4.
350
decomposition of nitrous oxide. Appl. Catal. B-Environ. 1996, 9, 25-64.
351
5.
352
transition metal exchanged ZSM-5 zeolites prepared by the solid-state ion-exchange
353
method. Appl. Catal. B-Environ. 2008, 84, (1-2), 277-288.
354
6.
355
catalyst for effective decomposition of nitrous oxide. Environ. Sci. Technol. 2007, 41,
356
7901-7906.
357
7.
358
catalysts prepared by different methods for the decomposition of N2O. Appl. Catal. B-
359
Environ. 2004, 51, (4), 215-228.
360
8.
361
decomposition over Fe-based catalysts: effects of gas-phase composition and catalyst
362
constitution. J. Catal. 2002, 208, (1), 211-223.
363
9.
364
mechanism of N2O decomposition over Cu-ZSM-5 zeolites. J. Phys. Chem. C 2012, 116,
365
(38), 20262-20268.
Kapteijn, F.; Rodriguez-Mirasol, J.; Moulijn, J. A., Heterogeneous catalytic
Abu-Zied, B. M.; Schwieger, W.; Unger, A., Nitrous oxide decomposition over
Li, L. D.; Shen, Q. Y., J. J.; Hao, Z. P.; Xu, Z. P.; Lu, G. Q. M., Fe-USY zeolite
Pieterse, J. A. Z.; Booneveld, S.; van den Brink, R. W., Evaluation of Fe-zeolite
Pérez-Ramírez, J.; Kapteijn, F.; Mul, G.; Moulijn, J. A., NO-assisted N2O
Liu, X.; Yang, Z.; Zhang, R.; Li, Q.; Li, Y., Density functional theory study of
17 ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 37
366
10.
Namuangruk, S.; Khongpracha, P.; Tantirungrotechai, Y.; Limtrakul, J.,
367
Decomposition of nitrous oxide on carbon nanotubes. J. Mol. Graph. Model. 2007, 26,
368
(1), 179-186.
369
11.
370
Limtrakul, J., Structures, energetics and reaction mechanisms of nitrous oxide on
371
transition-metal-doped and -undoped single-wall carbon nanotubes. Chem. Phys. Chem
372
2012, 13, (2), 583-587.
373
12.
374
various spinel-type oxides. Catal. Today 2007, 119, (1-4), 228-232.
375
13.
376
decomposition of nitrous oxide to nitrogen by in situ oxygen removal with a perovskite
377
membrane. Angew. Chem. Int. Ed. 2009, 48, (16), 2983-2986.
378
14.
379
of alkali metals and alkaline earth metals on cobalt-cerium composite oxide catalysts for
380
N2O decomposition. Environ. Sci. Technol. 2009, 43, 890-895.
381
15.
382
On the importance of the catalyst redox properties in the N2O decomposition over
383
alumina and ceria supported Rh, Pd and Pt. Appl. Catal. B-Environ. 2010, 96, (3-4), 370-
384
378.
Pannopard, P.; Khongpracha, P.; Warakulwit, C.; Namuangruk, S.; Probst, M.;
Russo, N.; Fino, D.; Saracco, G.; Specchia, V., N2O catalytic decomposition over
Jiang, H.; Wang, H.; Liang, F.; Werth, S.; Schiestel, T.; Caro, J., Direct
Xue, L.; He, H.; Liu, C.; Zhang, C.; Zhang, B., Promotion effects and mechanism
Parres-Esclapez, S.; Illán-Gómez, M. J.; de Lecea, C. S.-M.; Bueno-López, A.,
18 ACS Paragon Plus Environment
Page 19 of 37
Environmental Science & Technology
385
16.
Yuzaki, K.; Yarimizu, T.; Ito, S.; Kunimori, K., Catalytic decomposition of N2O
386
over supported rhodium catalysts: high activities ofRh/USY and Rh/Al2O3 and the effect
387
of Rh precursors. Catal. Lett. 1997, 47, 173-175.
388
17.
389
J., Bimetallic Au–Pd Alloy catalysts for N2O decomposition: effects of surface structures
390
on catalytic activity. J. Phys. Chem. C 2012, 116, (10), 6222-6232.
391
18.
392
Fe/ZSM-5 with low iron loading. J. Phys. Chem. B 2005, 109, 2295-2301.
393
19.
394
nitrous oxide decomposition over Fe- and Co-ZSM-5. J. Phys. Chem. B 2002, 106, 7059-
395
7064.
396
20.
397
Schoonheydt, R. A., Bis(µ-oxo)dicopper in Cu-ZSM-5 and its role in the decomposition
398
of NO: a combined in situ XAFS, UV-Vis-Near-IR, and kinetic study. J. Am. Chem. Soc.
399
2003, 125, 7629-7640.
400
21.
401
oxide formation during the interaction of nitrogen oxides with Cu-ZSM-5 at low
402
temperature. Appl. Catal. A-Gen. 2012, 413-414, 117-131.
403
22.
404
decomposition and surface oxygen formation on Fe-ZSM-5. J. Catal. 2004, 224, (1), 148-
405
155.
Wei, X.; Yang, X.-F.; Wang, A.-Q.; Li, L.; Liu, X.-Y.; Zhang, T.; Mou, C.-Y.; Li,
Sang, C.; Kim, B. H.; Lund, C. R. F., Effect of NO upon N2O decomposiiton over
Ryder, J. A.; Chakraborty, A. K.; Bell, A. T., Density functional theory study of
Groothaert, M. H.; van Bokhoven, J. A.; Battiston, A. A.; Weckhuysen, B. M.;
Lisi, L.; Pirone, R.; Russo, G.; Santamaria, N.; Stanzione, V., Nitrates and nitrous
Wood, B. R.; Reimer, J. A.; Bell, A. T.; Janicke, M. T.; Ott, K. C., Nitrous oxide
19 ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 37
406
23.
Pantu, P.; Boekfa, B.; Sunpetch, B.; Limtrakul, J., Nanocavity effects on N2O
407
decomposition on different types of Fe-Zeolites (Fe-FER, Fe-BEA, Fe-ZSM-5 and Fe-
408
FAU): a combined theoretical and experimental study. Chem. Eng. Commun. 2008, 195,
409
(11), 1477-1485.
410
24.
411
decomposition of nitrous oxide. Appl. Catal. B-Environ. 1996, 9, 25-64.
412
25.
413
decomposition of nitrous oxide over M (Fe, Co, Cu)–BEA zeolites. J. Catal. 2012, 294,
414
99-112.
415
26.
416
nitrous oxide decomposition over Fe-ZSM-5. J. Phys. Chem. B 2005, 109, 1857-1873.
417
27.
418
N2O to N2 conversion during the catalytic decomposition of NO by Cu-zeolites. Catal.
419
Lett. 2001, 74, 193-199.
420
28.
421
functional study. Catal. Today 2008, 137, (2-4), 410-417.
422
29.
423
of porphyrins and metalloporphyrins. J. Porphyrins Phthalocyanines 2000, 4, 407-413.
424
30.
425
K. C.; Schmidt-Mende, L.; Nazeeruddin, M. K.; Wang, Q.; Grätzel, M.; Officer, D. L.,
Kapteijn, F.; Rodriguez-Mirasol, J.; Moulijn, J. A., Heterogeneous catalytic
Liu, N.; Zhang, R.; Chen, B.; Li, Y.; Li, Y., Comparative study on the direct
Heyden, A.; Peters, B.; Bell, A. T.; Keil, F. J., Comprehensive DFT study of
Sengupta, D.; Adams, J. B.; Schneider, W. F.; Hass, K. C., Theoretical analysis of
Fellah, M. F.; Onal, I., N2O decomposition on Fe- and Co-ZSM-5: A density
Suslick, K. S.; Rakow, N. A.; Kosal, M. E.; Chou, J.-H., The materials chemistry
Campbell, W. M.; Jolley, K. W.; Wagner, P.; Wagner, K.; Walsh, P. J.; Gordon,
20 ACS Paragon Plus Environment
Page 21 of 37
Environmental Science & Technology
426
Highly efficient porphyrin sensitizers for dye-sensitized solar cells. J. Phys. Chem. C.
427
Lett. 2007, 111, 11760-11762.
428
31.
429
Coordin. Chem. Rev. 2003, 247, (1-2), 1-19.
430
32.
431
properties of metal-mediated assemblies of porphyrins. Coordin. Chem. Rev. 2006, 250,
432
(11-12), 1471-1496.
433
33.
434
reaction by porphyrin-based catalysts for fuel cells. Electrocatal. 2012, 3, (3-4), 238-251.
435
34.
436
porphyrin anions in the gas phase. Angew. Chem. Int. Ed. 2013, 52, (39), 10374-10377.
437
35.
438
with small ligands X (X: O, CO, NO, O2, N2, H2O, N2O, CO2). J. Phys. Chem. A 2009,
439
113, 9202-9206.
440
36.
441
porphyrins: radiative association, blackbody infrared radiative dissociation, and gas-
442
phase association equilibrium. J. Am. Chem. Soc. 1999, 121, 11910-11911.
443
37.
444
selectivity among NO, CO, and O2 by heme protein sensors. Biochemistry 2012, 51, (1),
445
172-186.
Harvey, J., Developments in the metal chemistry of N-confused porphyrin.
Scandola, F.; Chiorboli, C.; Prodi, A.; Iengo, E.; Alessio, E., Photophysical
Cheng, N.; Kemna, C.; Goubert-Renaudin, S.; Wieckowski, A., Reduction
Karpuschkin, T.; Kappes, M. M.; Hampe, O., Binding of O2 and CO to metal
Abdurahman, A.; Renger, T., Density functional studies of iron-porphyrin cation
Chen, O.; Groh, S.; Liechty, A.; Ridge, D. P., Binding of nitric oxide to iron(ii)
Tsai, A. L.; Berka, V.; Martin, E.; Olson, J. S., A "sliding scale rule" for
21 ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 37
446
38.
Kano, K.; Itoh, Y.; Kitagishi, H.; Hayashi, T.; Hirota, S., A supramolecular
447
receptor of diatomic molecules (O2, CO, NO) in aqueous Solution. J. Am. Chem. Soc.
448
2008, 130, 8006-8015.
449
39.
450
di-iron(ii) complexes and O2, CO, and NO adducts as reduced and substrate-bound
451
models for the active site of bacterial nitric oxide reductase. J. Am. Chem. Soc. 2005, 127,
452
3310-3320.
453
40.
454
chromophores for metal oxide semiconductor sensitization: effect of the spacer length
455
and anchoring group position. J. Am. Chem. Soc. 2007, 129, 4655-4665.
456
41.
457
(1), 291-304.
458
42.
459
Di Natale, C., The influence of gas adsorption on photovoltage in porphyrin coated ZnO
460
nanorods. J. Mater.Chem. 2012, 22, (37), 20032-20037.
461
43.
462
chemical sensors. Adv. Colloid Interfac. Sci. 2012, 171-172, 17-35.
463
44.
464
organization of supramolecular self-assembled porphyrin hollow hexagonal nanoprisms.
465
J. Am. Chem. Soc. 2005, 127, 17090-17095.
Wasser, I. M.; Huang, H.-W.; Moënne-Loccoz, P.; Karlin, K. D., Heme/non-heme
Rochford, J.; Chu, D.; Hagfeldt, A.; Galoppini, E., Tetrachelate porphyrin
Li, L. L.; Diau, E. W., Porphyrin-sensitized solar cells. Chem. Soc. Rev. 2013, 42,
Sivalingam, Y.; Magna, G.; Pomarico, G.; Catini, A.; Martinelli, E.; Paolesse, R.;
Giancane, G.; Valli, L., State of art in porphyrin Langmuir-Blodgett films as
Hu, J.-S.; Guo, Y.-G.; Liang, H.-P.; Wan, L.-J.; Jiang, L., Three-dimensional self-
22 ACS Paragon Plus Environment
Page 23 of 37
Environmental Science & Technology
466
45.
Lee, S. J.; Hupp, J. T., Porphyrin-containing molecular squares: design and
467
applications. Coordin. Chem. Rev. 2006, 250, (13-14), 1710-1723.
468
46.
469
Porous metalloporphyrinic frameworks constructed from metal 5,10,15,20-tetrakis(3,5-
470
biscarboxylphenyl)porphyrin for highly efficient and selective catalytic oxidation of
471
alkylbenzenes. J. Am. Chem. Soc. 2012, 134, (25), 10638-10645.
472
47.
473
organic framework: from theory to computational design. J. Phys. Chem. C 2012, 116,
474
(44), 23494-23502.
475
48.
476
D. W.; Wilson, S. R., Microporous porphyrin solids. Acc. Chem. Res. 2005, 38, 283-291.
477
49.
478
the electrochemical reduction of carbon dioxide. Chem. Commun. 2012, 48, (10), 1392-
479
1399.
480
50.
481
porphyrin catalyzed electrochemical reduction of carbon dioxide in water. 2. mechanism
482
from first principles. J. Phys. Chem. A 2010, 114, 10174-10184.
483
51.
484
reduction of carbon dioxide for solar fuels. Acc. Chem. Res. 2009, 42, 1983-1994.
485
52.
486
Am. Chem. Soc. 1995, 117, 5594-5595.
Yang, X. L.; Xie, M. H.; Zou, C.; He, Y.; Chen, B.; O'Keeffe, M.; Wu, C. D.,
Roy, S.; George, C. B.; Ratner, M. A., Catalysis by a zinc-porphyrin-based metal–
Suslick, K. S.; Bhyrappa, P.; Chou, J.-H.; Kosal, M. E.; Nakagaki, S.; Smithenry,
Finn, C.; Schnittger, S.; Yellowlees, L. J.; Love, J. B., Molecular approaches to
Leung, K.; Nielsen, I. M. B.; Sai, N.; Medforth, C.; Shelnutt, J. A., Cobalt-
Morris, A. J.; Meyer, G. J.; Fujita, E., Molecular approaches to the photocatalytic
Groves, J. T.; Roman, J. S., Nitrous oxide activation by a ruthenium porphyrin. J.
23 ACS Paragon Plus Environment
Environmental Science & Technology
Page 24 of 37
487
53.
Saito, S.; Ohtake, H.; Umezawa, N.; Kobayashi, Y.; Kato, N.; Hirobe, M.;
488
Higuchi, T., Nitrous oxide reduction-coupled alkene-alkene coupling catalysed by
489
metalloporphyrins. Chem. Commun. 2013, 49, (79), 8979-8981.
490
54.
491
as electrocatalysts for commercially important reacyions. Transit. Metal Chem. 2003, 28,
492
838-847.
493
55.
494
heterogeneous catalysts. Res. Chem. Intermed. 2006, 32, 235-251.
495
56.
496
oxidative DNA cleavage. Chem. Rev. 1992, 92, 1411-1456.
497
57.
498
alkanes with molecular oxygen in the presence of acetaldehyde. Tetrahedron Lett. 1995,
499
36, 8059-8062.
500
58.
501
by rhodium(iii) porphyrin complexes in water: reactivity and mechanistic studies. Chem.
502
Commun. 2010, 46, (34), 6353.
503
59.
504
oxide reduction comparison of several porphyrins and binary "face-to-face" porphyrins.
505
J. Electroanal. Chem. 1981, 124, 113-131.
506
60.
507
thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and
Phougat, N.; Vasudevan, P.; Jha, N. K.; Bandhopadhyay, D. K., Metal porphyrins
Wang, Y., Selective oxidation of hydrocarbons catalyzed by iron-containing
Meunier, B., Metalloporphyrins as versatile catalysts for oxidation reactions and
Murahashi, S.-I.; Naota, T.; Komiya, N., Metalloporphyrin-catalyzed oxidation of
Liu, L.; Yu, M.; Wayland, B. B.; Fu, X., Aerobic oxidation of alcohols catalyzed
Collman, J. P.; Marrocco, M.; Elliott, C. M.; L'Her, M., Electrocatalysis of nitrous
Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group
24 ACS Paragon Plus Environment
Page 25 of 37
Environmental Science & Technology
508
transition elements: two new functionals and systematic testing of four M06-class
509
functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215-41.
510
61.
511
methods. 9. Extended Gaussian-type basis for molecular-orbital studies of organic
512
molecules. J. Chem. Phys. 1971, 54, 724.
513
62.
514
III, 1976, 3 (Plenum, New York), 1-28.
515
63.
516
Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji,
517
H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.;
518
Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida,
519
M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.;
520
Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.;
521
Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J.
522
C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.;
523
Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.;
524
Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;
525
Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.;
526
Dapprich, S.; Daniels, A. D.; Farkas, O .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.;
527
Fox, D. J., Gaussian 09, Revision B.01; Gaussian, Inc.: Wallingford, CT, 2010.
528
64.
529
7211-7218.
Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-consistent molecular orbital
Dunning Jr., T. H.; Hay, P. J. Modern Theoretical Chemistry, Ed. H. F. Schaefer
Frisch, M. J. T., G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;
Foster, J. P.; Weinhold, F. Natural hybrid orbitals. J. Am. Chem. Soc., 1980, 102,
25 ACS Paragon Plus Environment
Environmental Science & Technology
Page 26 of 37
530
65.
Feixas, F.; Sola, M.; Swart, M. Chemical bonding and aromaticity in
531
metalloporphyrin. Can. J. Chem. 2009, 87, 1063-1073.
532
66.
533
binding specificity of porphyrin: a conceptual density functional theory study. J. Phys.
534
Chem. C 2009, 113, 13381-13389.
Feng, X-T.; Yu, J-G.; Lei, M.; Fang, W-H.; Liu, S., Toward understanding metal-
535
26 ACS Paragon Plus Environment
Page 27 of 37
Environmental Science & Technology
536
Table 1. Structural parameters (bond lengths in Å, angles in degrees), adsorption energy
537
(kcal/mol), and partial charges (e) on the selected atoms of the metal–porphyrin and
538
N2O…metal–porphyrin complex systems, where the metal is Ti, Cr, Fe, Co, Ni, Cu, or Zn. N3
539
O1
N2
540 541
M
542 543 N2O
Ti
Cr
Fe
Co
Ni
Cu
Zn
Structural Parameters M-por system Spin multiplicity M–Npor ∠Npor–M–Npor
3 2.06 180
5 2.04 180
3 2.00 180
2 1.99 180
1 1.97 180
2 2.02 180
1 2.06 180
N2O…M-por system N2O Adsorption Energy M–Npor ∠Npor–M–Npor M–O1 O1–N2 N2–N3 ∠M–O1–N2 ∠O1–N2–N3
-13.32 2.06 172.4 2.25 1.20 1.14 120.3 178.3
-6.81 2.04 178.5 2.74 1.19 1.14 111.1 179.6
-7.81 2.00 179.4 2.74 1.19 1.14 108.4 179.9
-6.84 1.99 179.2 2.64 1.19 1.14 108.7 179.9
-4.65 1.97 179.7 3.14 1.19 1.14 92.1 179.9
-5.48 2.02 179.7 2.81 1.19 1.14 106.4 179.8
-6.92 2.06 176.9 2.54 1.19 1.14 114.3 179.4
Partial Charges M-por system M Npor
1.46 -0.70
0.96 -0.62
0.79 -0.59
0.82 -0.58
0.71 -0.56
0.99 -0.63
1.39 -0.72
N2O…M-por system M Npor O1 N2 N3 Total charge of N2O (qdiff) Mdiff (MN2O-Mpor) Npor-diff (NN2O-Npor)
1.29 -0.67 -0.29 0.40 0.02 0.12 -0.18 0.03
0.92 -0.61 -0.30 0.39 -0.06 0.04 -0.04 0.00
0.84 -0.59 -0.29 0.39 -0.06 0.04 0.05 0.00
0.78 -0.58 -0.29 0.39 -0.06 0.05 -0.04 0.00
0.69 -0.57 -0.29 0.39 -0.08 0.02 -0.02 -0.00
0.97 -0.63 -0.29 0.39 -0.07 0.03 -0.03 0.00
1.36 -0.71 -0.31 0.39 -0.04 0.04 -0.03 0.00
1.19 1.14 180.0
-0.30 0.38 -0.08 0.00
544 27 ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 37
545
Table 2. Comparison of the adsorption energy, activation energy, and desorption energy
546
for the direct decomposition of N2O over potential zeolites and Ti–porphyrin catalyst.
st
1 N2O 2nd N2O
3rd N2O
Activation barrier Activation barrier O2 desorption
Ti–porphyrin 3.77
Energy (kcal/mol) Cu-ZSM-5 Fe-ZSM-5 a 35.18 4.41b
49.99
28.07a
102.96
39.48a
Activation 47.79 barrier 3.63 O2 desorption a 9b 28 c Liu, et al. Fellah et al. Ryder et al. 19
Co-ZSM-5 6.28b
42.10a
58.13b 37.6c 67.30b 94.9c 44.6c
48.56b 32.9c 64.95b 85.6c 40.2c
63.42a
51.9c
52.8c
547 548
28 ACS Paragon Plus Environment
Page 29 of 37
Environmental Science & Technology
549
Figure Captions
550
Figure 1. Model systems in this work.
551
Figure 2. Mechanistic cycles of direct decomposition of N2O over Ti–porphyrin while
552
the inner cycle shows the recycling of the active TiO–porphyrin catalytic intermediate for
553
the third N2O decomposition.
554
Figure 3. The first N2O decomposition over Ti–porphyrin (Step 1).
555
Figure 4. The second N2O decomposition over TiO–porphyrin (Step 2).
556
Figure 5. The third N2O decomposition over TiO2–porphyrin (Step 3).
557
Figure 6. O2 depletion and desorption processes from the TiO3–porphyrin intermediate
558
catalyst (Step 4).
559
Figure 7. Energy profile of the full reaction pathway for the direct decomposition of
560
three N2O molecules on the Ti–porphyrin catalyst.
561
29 ACS Paragon Plus Environment
Environmental Science & Technology
Page 30 of 37
562
563 564
Figure 1. Model systems in this work.
565
30 ACS Paragon Plus Environment
Page 31 of 37
Environmental Science & Technology
566
567 568
Figure 2. Mechanistic cycles of direct decomposition of N2O over Ti–porphyrin while
569
the inner cycle shows the recycling of the active TiO–porphyrin catalytic intermediate for
570
the third N2O decomposition.
571
31 ACS Paragon Plus Environment
Environmental Science & Technology
Page 32 of 37
572 573
Figure 3. The first N2O decomposition over Ti–porphyrin (Step 1).
574
32 ACS Paragon Plus Environment
Page 33 of 37
Environmental Science & Technology
575 576
Figure 4. The second N2O decomposition over TiO–porphyrin (Step 2).
577
33 ACS Paragon Plus Environment
Environmental Science & Technology
Page 34 of 37
578 579
Figure 5. The third N2O decomposition over TiO2–porphyrin (Step 3).
580
34 ACS Paragon Plus Environment
Page 35 of 37
Environmental Science & Technology
581 582
Figure 6. O2 depletion and desorption processes from the TiO3–porphyrin intermediate
583
catalyst (Step 4).
584
35 ACS Paragon Plus Environment
Environmental Science & Technology
Page 36 of 37
585 586
Figure 7. Energy profile of the full reaction pathway for the direct decomposition of
587
three N2O molecules on the Ti–porphyrin catalyst.
588
36 ACS Paragon Plus Environment
Page 37 of 37
589
Environmental Science & Technology
Table of Contents (TOC)
590
37 ACS Paragon Plus Environment