Subscriber access provided by Mount Allison University | Libraries and Archives
Remediation and Control Technologies
Balance between reducibility and N2O adsorption capacity for the N2O decomposition: CuxCoy catalysts as an example Shangchao Xiong, Jianjun Chen, Nan Huang, Shijian Yang, Yue Peng, and Junhua Li Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b02892 • Publication Date (Web): 05 Aug 2019 Downloaded from pubs.acs.org on August 5, 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 32
Environmental Science & Technology
1
Balance between reducibility and N2O adsorption capacity for
2
the N2O decomposition: CuxCoy catalysts as an example
3
Shangchao Xiong1, 2, Jianjun Chen*1, 2, Nan Huang1, 3, Shijian Yang4, Yue Peng1, 2, Junhua Li1, 2
4
1State
5
Environment, Tsinghua University, Beijing 100084, PR China
6
2National
7
Equipment, Beijing, 100084, China
8
3School
9
Technology, Nanjing, 210094 PR China
10
Key Joint Laboratory of Environment Simulation and Pollution Control, School of
4School
Engineering Laboratory for Multi Flue Gas Pollution Control Technology and
of Environmental and Biological Engineering, Nanjing University of Science and
of Environment and Civil Engineering, Jiangnan University, Wuxi 214122, PR China
11 12
*Corresponding author.
13
Phone: +86 010 62771093
14 15
Email address:
[email protected] (Jianjun Chen)
1
ACS Paragon Plus Environment
Environmental Science & Technology
16
Abstract:
17
CuxCoy (CuO−Co3O4 mixed oxides) catalysts were prepared via co−precipitation for the N2O
18
decomposition reaction. They exhibited a higher N2O decomposition activity than that of pure
19
CuO and Co3O4 due to the balance of redox property and N2O adsorption capacity. Co3O4
20
presented a large number of surface oxygen vacancies, increasing the N2O chemical adsorption as
21
“□−Co−ON2” on the catalyst surface, whereas CuO was dispersed around Co3O4 and presented
22
high reducibility on the interface of Co3O4−CuOx for the N−O break of N2O, healing oxygen
23
vacancies by leaving one oxygen atom in the vacancy. Based on kinetic studies, the rate constant
24
of N2O decomposition was related to the number of surface vacancy sites ([Mn+]) and the rate of
25
N−O break (k3), while the rate determining step is the N−O break. Therefore, the N2O
26
decomposition rate is first order to the N2O concentration. Overall, both the DFT calculations and
27
kinetic results indicate that the quantities of adsorption and activation sites derived from the
28
interaction between Co and Cu (dual−function mechanism) were accounted for the excellent N2O
29
decomposition performance of CuxCoy catalysts.
30
2
ACS Paragon Plus Environment
Page 2 of 32
Page 3 of 32
Environmental Science & Technology
31
Table of Contents
32 33
3
ACS Paragon Plus Environment
Environmental Science & Technology
34
1.
Introduction
35
Nitrous oxide (N2O) emitted from the production of adipic acid and nitric acid, as well as the
36
processes using nitric acid as an oxidant, contributes to the ozone hole and greenhouse effect.1, 2
37
Its global warming potential (GWP) is ~310 times and ~31 times higher than that of CO2 and CH4,
38
respectively, and the lifetime of N2O is ~114 years.3 Moreover, N2O can deplete the ozone layer
39
by a reaction pathway similar to that of chlorofluorocarbons (CFCs). Previous studies reported
40
that N2O would be the dominant ozone−depleting substance in the 21st century.4 Thus, the
41
reduction in anthropogenic N2O emissions is urgently required. Several techniques were proposed
42
to control anthropogenic N2O emissions, whereas the direct catalytic decomposition of N2O is
43
regarded as the most promising alternative technique.5,
44
treatment process to incorporate this technique is relatively convenient and can minimize the
45
economic demands.
6
Retrofitting the existing flue gas
46
A series of noble metals and nonnoble metals were used to catalyze the decomposition of N2O.6
47
Noble metals (e.g., Rh and Ru) show a satisfactory N2O decomposition performance at low
48
temperature, but their high cost and poor tolerance to various influential factors (e.g., oxygen and
49
water vapor) extremely restrict their widespread applications.7-9 Iron−based zeolites (especially
50
Fe−ZSM−5) are another type of N2O decomposition catalyst, which attracted great interest
51
because of their tolerance to O2 and H2O.10, 11 The N2O decomposition activity of Fe−ZSM−5 is
52
even promoted by the presence of NO in flue gas.12 However, the reaction temperature of
53
iron−based zeolites is quite high, and it is difficult to meet the actual flue gas conditions.
54
Metal oxides, especially transition metal oxides, are widely researched and employed in the
55
N2O decomposition reaction, which are consequences of their low price, excellent reducibility and
56
adequate catalytic characteristics.13, 14 Particularly, metal oxides exhibiting the spinel structure are
57
efficient catalysts to decompose N2O.15 The metal cations in the spinel structure are in the mixed
58
valence state, which frequently consists of divalent and trivalent states. The divalent and trivalent
59
cations in the spinel structure are located in tetrahedral and octahedral coordination centers and are
60
represented as AIIBIII2O4. Since the key step in the N2O decomposition reaction is generally
61
regarded as the charge transfer from the active sites to the antibonding orbital of N2O, spinels can
62
decompose N2O at a relatively low temperature due to their excellent redox property attributed to 4
ACS Paragon Plus Environment
Page 4 of 32
Page 5 of 32
Environmental Science & Technology
63
the divalent and trivalent cations in the spinel structure.16
64
with the spinel structure were systematically investigated in the decomposition of N2O.6 Russo et
65
al. investigated several spinel−type catalysts and found that Co−based spinels can provide the
66
most efficient N2O decomposition performance.18 However, the redox properties of Co−based
67
spinels are not the best among those of spinel catalysts. Consequently, there must exist another
68
crucial property that significantly affects the catalytic performance of N2O decomposition. Many
69
researchers used DFT methods to calculate the reaction pathway of N2O decomposition and
70
proposed that N2O adsorption is the first step in N2O decomposition.15, 19 The chemical adsorption
71
of N2O generally follows “N−N−O−□”.20,
72
vacancies can contribute to the chemical adsorption of N2O, and this is probably the main reason
73
for the superior N2O decomposition performance of the Co3O4 spinel. Given this perspective,
74
improving the reducibility without blocking oxygen vacancies is the most efficient way to
75
improve the N2O decomposition performance of the Co3O4 spinel. Cu−based catalysts are another
76
type of N2O decomposition catalysts that possess a superior redox property22. Combining the
77
advantages of both the Cu−based catalysts and Co3O4 spinel can certainly improve the N2O
78
decomposition performance. Therefore, in this work, a series of CuxCoy (CuO−Co3O4 spinel
79
mixed oxides) catalysts were synthesized to decompose N2O.
21
17
Given this perspective, metal oxides
This result suggests that abundant surface oxygen
80
Kinetic study is an important approach to investigate the key factors of the N2O decomposition
81
reaction. Freek Kapteijn et al. accomplished a comparative kinetic analysis over Co−, Fe−, and
82
Cu−ZSM−5 and found that the effects of O2, NO and CO were influenced by their partial
83
pressure.23 L. Obalova and V. Fıla established a novel kinetic model over hydrotalcites, which
84
proposes that N2O chemisorption determines the rate of N2O decomposition at low O2 partial
85
pressures, whereas the reaction between active O atoms and N2O is the rate−determine step at high
86
O2 partial pressures.24 However, kinetic studies of N2O decomposition are rarely reported among
87
recent studies and are even absent from recent reviews.6,
88
between the kinetic model and physicochemical properties exists and greatly limits the design of
89
efficient catalysts for N2O decomposition.
15
Furthermore, a lack of connection
90
Herein, the N2O decomposition mechanism and the key roles of CuO and Co3O4 spinel in
91
CuxCoy mixed oxides were systematically investigated by a kinetic study combined with DFT, in 5
ACS Paragon Plus Environment
Environmental Science & Technology
92
situ DRIFTs, N2O−TPD, H2−TPR and XPS studies. The crucial properties of CuxCoy catalysts and
93
the connection between the physicochemical properties and the kinetic study were proposed.
94
2.
95
2.1 Catalyst preparation
Experimental
96
Co3O4, CuxCoy and CuO catalysts were prepared via the coprecipitation method. CuxCoy
97
represents Cu1Co2, Cu1.5Co1.5 and Cu2Co1 catalysts, corresponding to molar ratios of Cu to Co of
98
1:2, 1:1 and 2:1, respectively. Suitable amounts of cupric nitrate and cobaltous sulfate were added
99
to a solution with an excess of sodium hydroxide followed by continuous stirring for 3 h. The
100
suspension was separated by centrifugation and washed with deionized water. The process of
101
centrifugation and washing was repeated 5 times to remove any residual substances. The obtained
102
particles were dried at 105 °C for 12 h and then calcinated at 500 °C for 3 h in air.
103
2.2 Characterization
104
The BET surface area and X−ray diffraction (XRD) data were determined on a physisorption
105
analyzer (Blesorp max II) and an X−ray diffractometer (Rigaku D/max−2500). The surface
106
analyses (XPS and AES) were carried out on an X−ray photoelectron spectroscopy microprobe
107
(EscaLab 250 Xi). Temperature program desorption (TPD) of O2 and N2O were both conducted
108
on a chemisorption apparatus (Autochem II 2920), and N2O−TPD were further analyzed by a
109
mass spectrum (MS, HPR−20 R&D). H2−temperature program reduction (H2−TPR) was also
110
performed on the chemisorption apparatus. After H2−TPR studies, the coefficient between H2
111
consumption rate and peak intensity was gotten by the test results of standard sample (standard
112
CuO). Therefore, the dependences of H2 consumption rates versus 1/T (T=130~165 oC) were
113
obtained as the initial H2 consumption rates.
114
2.3 DFT calculation details
115
Previous experimental data indicated that the Co3O4 spinel mainly exposes the (100) and (111)
116
planes, with only a minor exposure of the (110) plane.25 Additionally, the (100) plane is more
117
stable than the (110) and (111) planes in a wide range of temperatures.26 Thus, the (100) plane of
118
the Co3O4 spinel was reconstructed by a [2×2] supercell to generate the slab model. For CuO, the
119
(111) plane is regarded as the most stable plane.27 Therefore, a slab model of the CuO (111) plane 6
ACS Paragon Plus Environment
Page 6 of 32
Page 7 of 32
120
Environmental Science & Technology
was also reconstructed by a [2×2] supercell.
121
All calculations were conducted by the Vienna ab initio simulation package (VASP 5.4.4). The
122
Perdew, Burke, and Ernzerhof (PBE) functional within the generalized gradient approximation
123
plus Hubbard model (GGA+U) was used to calculate the electronic exchange and correlation. The
124
Ueff of Cu and Co in this study were 7.0 eV and 3.5 eV, respectively.28, 29 The cutoff energy was
125
500 eV, and a Monkhorst−Pack grid of 2×2×1 k−points were employed due to the large size of the
126
slab (~12 Å×12 Å). The thickness of the slab was ~8 Å, with a 15 Å vacuum gap. Moreover, all
127
the slabs were relaxed until the atomic forces were reduced below 0.05 eV/Å.
128
The adsorption energies of N2O (Ead) were estimated by the following equation:
129
Ead= Esurf+N2O−Esurf−EN2O
130
where Esurf represents the energy of the clean surface, EN2O denotes the energy of a free N2O
131
molecule in the vacuum, and Esurf+N2O is the energy of N2O adsorbed on the surface. It is
132
noteworthy that a negative value for Ead indicates a stable adsorption.
133
2.4 Activity test
134
The catalytic decomposition of N2O was performed in a fixed−bed reactor with 100 ml min−1 of
135
flue gas containing 1000 ppm N2O, 2% O2 (when used), 200 ppm NO (when used), 0.5% H2O
136
(when used) and the balance as N2. The catalyst mass was 100 mg, and the corresponding GHSV
137
was 60,000 cm3 g−1 h−1. The N2O concentration at the outlet was monitored online by a
138
MultiGas™ 2030 FTIR continuous gas analyzer.
139
The steady-state kinetic study of N2O decomposition was also performed in the fixed−bed
140
reactor. The flue gas contained 500−1500 ppm N2O with the balance as N2. An extremely high
141
GHSV of 60,000−6,000,000 cm3 g−1 h−1 was used to ensure that the N2O decomposition was less
142
than 20%, thus, overcoming the diffusion limitation.30-32
143
3.
144
3.1 Performance of N2O decomposition
Results
145
The N2O decomposition performance of Co3O4, CuxCoy and CuO catalysts is shown in Figure
146
1a. The N2O decomposition activity of Co3O4 was superior to that of CuO. Furthermore, all the
147
CuxCoy catalysts showed superior activity to that of Co3O4 and CuO. Interestingly, although CuO
148
presented the lowest N2O decomposition activity, high amounts of Cu promoted the performance 7
ACS Paragon Plus Environment
Environmental Science & Technology
149
more significantly. For Cu2Co1, the reaction temperature for full N2O decomposition was ~375 °C
150
under idealized reaction conditions, which was significantly lower than the ~450 °C for Co3O4 and
151
~500 °C for CuO.
152
Generally, O2, NO and water vapor exist in flue gas and often interfere with N2O
153
decomposition.33 The effects of O2, NO and water vapor on N2O decomposition over Cu2Co1 were
154
investigated (Figure 1b). O2, NO and water vapor all interfered with the N2O decomposition
155
performance of Cu2Co1 at low temperatures, and the influencing degree increased according to the
156
following sequence: NO < O2 < H2O. These results were also observed in Co3O4, Cu1Co2 and
157
CuO catalysts (shown in Figure S1). Previous studies have shown that NO and H2O have
158
completely different inhibiting mechanisms in the N2O decomposition. H2O prefers to bind with
159
oxygen vacancy sites and then blocks the oxygen transfer, whereas NO shows a competitive
160
oxidation effect, which consumes labile oxygen and decelerates the regeneration of active sites.34
161
Further, O2 possibly can inhibit the recombination of residual O during N2O decomposition
162
reaction, leading to the inhibition of the regeneration of active sites. On the whole, N2O
163
decomposition dramatically decreased from ~90% to ~6% at 350 °C when O2, NO and water
164
vapor coexisted; however, their influencing degrees decreased with increasing temperature. The
165
N2O decomposition of Cu2Co1 in the presence of O2, NO and H2O increased form ~6% to ~95%
166
when the temperature increased from 350 °C to 450 °C. This result suggests that Cu2Co1 still
167
showed a superior N2O decomposition performance in the simulated flue gas.
168
3.2 Characterization
169
3.2.1 XRD and BET surface area
170
As shown in Figure S2, the XRD pattern of Co3O4 corresponded well to that of the cubic spinel
171
(JCPDS: #43−1003), and the XRD pattern of CuO was assigned to tenorite (JCPDS: #48−1548).
172
The XRD patterns of the CuxCoy catalysts showed characteristic peaks corresponding to both the
173
Co3O4 and CuO, and the peak positions were nearly unchanged. The crystal sizes and crystal
174
parameters of CuO and Co3O4 clusters in the Co3O4, CuxCoy and CuO catalysts were calculated on
175
the basis of the XRD patterns, and the results are shown in Table 1. The crystal parameters of
176
Co3O4 clusters of CuxCoy catalysts (a=b=c=~0.8097) were all slightly higher than those of pure
177
Co3O4 (a=b=c=0.8090), meanwhile the crystal parameters of CuO clusters of CuxCoy catalysts 8
ACS Paragon Plus Environment
Page 8 of 32
Page 9 of 32
Environmental Science & Technology
178
were slightly different with those of pure CuO. These results indicate that a small amount of
179
Cu−Co solid solution was generated, whereas most of CuO and Co3O4 existed in crystal form. The
180
Co3O4 and CuO catalysts exhibited the maximum crystal sizes, whereas the crystal size of Cu2Co1
181
was the smallest. These results are in accordance with the results of BET surface area. The BET
182
surface area of Cu2Co1 was higher than those of Co3O4, CuO and other CuxCoy catalysts.
183
Moreover, The BET surface area of CuO was only 8.1 m2 g−1, which was significantly lower than
184
those of the Co3O4 and CuxCoy catalysts. This difference might have been one of the reasons
185
responsible for the poor N2O decomposition performance of CuO.
186
3.2.2 Redox properties
187
The reducibility of active sites (e.g., Cu,21 Ni,35 Co,17 and Fe36) would strongly affect the N2O
188
decomposition performance. H2−TPR studies were performed to investigate the enhancement in
189
reducibility of CuxCoy catalysts (Figure 2a). The H2 reduction peaks of the CuxCoy catalysts were
190
similar to those of CuO, which are situated at lower temperatures than those of Co3O4. That result
191
means the reductions of CuO and Co3O4 in CuxCoy catalysts occurred at approximately the same
192
time. Consequently, the data indicate the Cu species in CuxCoy catalysts played a dominant role in
193
the redox reaction, which could promote the reduction of Co species by a charge interaction in the
194
CuxCoy catalysts.37 Moreover, the H2 reduction peaks and the initial H2 consumption temperature
195
of all the Cu-containing catalysts seem identical (~158 °C). To further identify the reducibility of
196
Cu-containing catalysts, the initial H2 consumption rates were determined and are shown in Figure
197
2b. The initial H2 consumption rate of Cu2Co1 was clearly faster than those of the other
198
Cu-containing materials. This result indicates that Cu2Co1 represented the optimal reducibility,
199
which could facilitate the N2O decomposition of the Cu2Co1 mixed oxide.
200
3.2.3 Surface analysis
201
Generally, the redox cycles between the Cu2+/Cu+ and Co3+/Co2+ play important roles in N2O
202
decomposition.38 Thus, the ratios of Cu+/(Cu++Cu2+) and Co2+/(Co2++Co3+) are crucial to N2O
203
decomposition.15 The AES and XPS studies were used to determine the surface components of the
204
Co3O4, CuxCoy, and CuO catalysts, and the surface chemical compositions are listed in Table 2. In
205
Figure 3a, the AES spectra of Cu-containing catalysts over the spectral region of the Cu LMM
206
contained features mainly centered at ~916.0 eV and ~918.0 eV, which were assigned to Cu+ and 9
ACS Paragon Plus Environment
Environmental Science & Technology
207
Cu2+, respectively.39, 40 Among the Cu-containing samples, the ratios of Cu+/(Cu++Cu2+) for all the
208
CuxCoy catalysts were higher than those for CuO. This result was further confirmed by the XPS
209
spectra for the Cu 2p3/2 spectral region (shown in Figure S3). Additionally, the Cu 2p3/2 spectral
210
region of pure CuO was situated at higher binding energies, whereas the peaks of Cu 2p3/2 for the
211
CuxCoy catalysts were shifted to lower binding energies. These results imply the electron cloud of
212
Cu species in the CuxCoy catalysts were altered due to the charge interaction between Cu and Co.
213
Corresponding with the shift in Cu 2p3/2, Figure 3b shows that the Co 2p3/2 spectral region of
214
pristine Co3O4 was also located at higher binding energies, and the peaks moved to lower binding
215
energies as Cu was added, further confirming the existence of the charge interaction between Cu
216
and Co. These results are in accordance with those of the H2−TPR study. The XPS spectra of
217
Co-containing samples for the spectral region of the Co 2p3/2 contained peaks mainly centered at
218
780.5 eV and 779.3 eV, which were attributed to Co2+ and Co3+, respectively. The ratios of
219
Co2+/(Co2++Co3+) for the CuxCoy catalysts were also higher than that for Co3O4. Overall, part of
220
the metal elements on the surface of the CuxCoy catalysts transformed from high valence states to
221
relatively low valence states, which was mainly due to the charge interaction between Co and
222
Cu.41 The facilitation of redox cycles between Cu2+/Cu+ and Co3+/Co2+ could play an important
223
role in the N2O decomposition performance of the CuxCoy catalysts.
224
3.2.4 N2O adsorption capacities
225
DFT calculations were employed to identify the characteristics of N2O adsorption on the
226
CuxCoy catalysts. Considering that the CuxCoy catalysts were mixed oxides of CuO and Co3O4, the
227
N2O adsorption configurations on their slab models were calculated. A negative value for Ead
228
indicates a stable adsorption in this study. As shown in Figure 4a, N2O could be weakly adsorbed
229
on Cu2+ to form a Cu−ON2 species. The bond length of Cu−O in the Cu−ON2 species was 2.87 Å,
230
and the corresponding N2O adsorption energy (Ead) was only −0.1 eV. N2O could also be weakly
231
adsorbed on a CuO surface with an oxygen vacancy to form a □−Cu−ON2 species (Figure 4b).
232
The bond length of Cu−O in the □−Cu−ON2 species was quite short (2.14 Å), whereas the
233
corresponding Ead was slightly lower than that of the Cu−ON2 species. These results suggest that
234
oxygen vacancies slightly promoted N2O adsorption on CuO, but these adsorption configurations
235
remained very unstable. For Co3O4, N2O could hardly be adsorbed on the complete surface 10
ACS Paragon Plus Environment
Page 10 of 32
Page 11 of 32
Environmental Science & Technology
236
structure (Figure 4c). The Ead of Co−ON2 was even higher than 0. However, the Ead of N2O
237
adsorbed on the Co3O4 surface with an oxygen vacancy (Figure 4d−f) was lower than those of
238
N2O adsorbed on Co3O4 and CuO, which suggests that N2O could be strongly adsorbed on the
239
Co3O4 surface with oxygen vacancies to form the □−Co−ON2 species. Consequently, N2O is more
240
likely to be adsorbed on a Co3O4 surface with oxygen vacancies than on a CuO or CuO−□ surface
241
in the CuxCoy catalysts.
242
N2O−TPD was performed to determine the capacity of N2O adsorption at 50 oC. In the
243
N2O−TPD study, all the samples were first treated under He atmosphere at 400 °C for 1 h and
244
then cooled to 50 °C to adsorb 2% N2O/He for 30 min. Finally, the original N2O−TPD profiles
245
(Figure S4a) were recorded by a TCD detector at a heating rate of 10 °C/min, and the detailed
246
desorption species were analyzed by MS (Figures S4b−4f). Therefore, the desorption amounts of
247
N2O (Figure 4g), N2 (Figure S4g) and NO (Figure S4h, a by-product of N2O decomposition) were
248
obtained by the integration of MS spectra. It is worth mentioning that the desorption amounts of
249
O2 generated by N2O decomposition during N2O−TPD could not be obtained, due to the
250
influences of adsorbed oxygen and/or the crystal oxygen on/in the catalysts.
251
Interestingly, although the N2O decomposition performance of Co3O4 was weaker than those of
252
the CuxCoy catalysts, the catalyst showed a relatively high capacity for N2O adsorption. This
253
property is mainly originated from the abundant surface oxygen vacancies of Co3O4 (shown in
254
Figure S5), which could promote the N2O adsorption through a □−Co−ON2 style (shown in Figure
255
4d−f). The high N2O adsorption capacity was probably responsible for the excellent N2O
256
decomposition performance of Co3O4 and the other Co−based spinels. In contrast, CuO exhibited
257
the worst N2O adsorption capacity, which was mainly due to the unstable Cu−ON2 and
258
□−Cu−ON2 species (shown in Figure 4a−b) and the lowest BET surface area, and therefore, this
259
catalyst showed the poorest N2O decomposition performance. The N2O adsorption capacities of
260
the CuxCoy catalysts increased with the Cu doping amount, which is in excellent accordance with
261
their N2O decomposition performance and O2−TPD profiles (Figure S5). Consequently, the
262
capacity of N2O adsorption played an important role in the N2O decomposition reaction.
263
4. Discussion 11
ACS Paragon Plus Environment
Environmental Science & Technology
264 265 266
Page 12 of 32
4.1 Reaction mechanism and kinetic study Many researchers proposed that the N2O decomposition process can be generally described as:17, 19, 21, 42, 43
267
N 2 O(g) +M n+ M n+ ONN
(1)
268
M n+ ONN M (n+1)+ O +N 2(g)
(2)
269
2M (n+1)+ O 2M n+ O 2(g)
(3)
270
First, N2O is adsorbed at an active site on the surface (Reaction 1). Then, the adsorbed N2O can
271
decompose to N2 and a residual O atom (Reaction 2). Finally, two residual O atoms on the surface
272
can combine to generate O2, and the active sites are regenerated (Reaction 3).
273
Thus far, the rate limiting step of N2O decomposition was comprehensively discussed using the
274
DFT calculations and experimental study. However, the conclusions were completely
275
inconsistent.15 Some researchers found that the splitting of N2O (Reaction 2) is the rate
276
determining step, but others considered that the recombination of O2 (Reaction 3) determines the
277
N2O decomposition rate.44-47
278
If the rate-determining step is the splitting of N2O (Reaction 2), the rate of N2O decomposition
279
can be described as:
280
N O
281 282
2
d[M n+ ONN] =k2 [M n+ ONN] dt
(4)
where vN2O, k2 and [Mnn+−ONN] represent the N2O decomposition rate, the kinetic constant of Reaction 2 and the concentration of adsorbed N2O, respectively.
283
While the GHSV is extremely high and the gaseous N2O concentration is relatively low,
284
Reaction 1 can be simply considered as a reversible reaction. Therefore, the concentration of the
285
adsorbed N2O can be approximately described as:
286
[M n+ ONN]=K1[M n+ ][N 2 O(g) ]
287 288 289
(5)
where K1, [Mn+] and [N2O(g)] represent the equilibrium constant of Reaction 1, the quantity of active sites and the concentration of gaseous N2O, respectively. Combined with Equations 4 and 5, the N2O decomposition rate can de depicted as:
12
ACS Paragon Plus Environment
Page 13 of 32
290
Environmental Science & Technology
N O K1k2 [M n+ ][N 2 O(g) ]
(6)
2
291
In Equation 6, K1 and k2 are only related to the reaction temperature over the same catalyst.
292
Therefore, K1 and k2 can be regarded as constants if the reaction reaches steady state. Meanwhile,
293
the quantity of active sites (Mn+) can be rapidly regenerated through Reaction 3. This action
294
suggests that the quantity of active sites (Mn+) can also be regarded as a constant. Therefore,
295
Equation 6 can be simplified to:
296
N O k N O [N 2 O(g) ]
(7)
297
k N2O K1k2 [M n+ ]
(8)
2
2
298
where kN2O denotes the reaction rate constant of the N2O decomposition.
299
Overall, the N2O decomposition is a first-order reaction when the rate-determining step is the
300 301
splitting of N2O. If the recombination of O2 (Reaction 3) determines the N2O decomposition rate, then the rate of
302
N2O decomposition can be described as:
303
N O
304 305
2
d[M (n+1)+ O ] =2k3 [M (n+1)+ O ]2 dt
(9)
where k3 and [M(n+1)+−O−] denote the reaction rate constant of Reaction 3 and the concentration of M(n+1)+−O−, respectively.
306
Reactions 1 and 2 can be regarded as opposing reactions while the GHSV is extremely high;
307
consequently, the gaseous N2O concentration is relatively low, and the recombination of O2 is the
308
rate-determining step. Therefore, Equation 5 remains workable in this case, and the concentration
309
of M(n+1)+−O− can be described as:
310
[M (n+1)+ O ] K 2 [M n+ ONN]
(10)
311
where K2 represents the equilibrium constant of Reaction 2 in this case.
312
Combining Equations 5, 9 and 10, the rate of N2O decomposition can be formulated as:
d[M (n+1)+ O ] =2k3 [M (n+1)+ O ]2 2 dt 2k3 K 2 2 [M n+ ONN]2 2k3 K 2 2 K12 [M n+ ]2 [N 2 O(g) ]2
N O 313
k N2O [N 2 O(g) ]2 13
ACS Paragon Plus Environment
(10)
Environmental Science & Technology
314 315
k N2O 2k3 K 2 2 K12 [M n+ ]2
Page 14 of 32
(11)
Overall, the N2O decomposition is a second-order reaction when the rate-determining step is the
316
recombination of O2.
317
4.2 Model verification
318
A steady-state kinetic study was performed to judge the rate-limiting step of N2O
319
decomposition. An extremely high GHSV and a relatively low gaseous N2O concentration were
320
employed to satisfy the assumptions of the kinetic equations and to overcome the diffusion
321
limitations. The results of the steady-state kinetic study over the Co3O4, CuxCoy and CuO catalysts
322
are shown in Figure S6. All the materials showed significant linear relationships between the
323
gaseous N2O concentration and the N2O decomposition rate from 350−500 °C, with all lines going
324
through the origin of the coordinates. This result indicates that the N2O decomposition reaction
325
was a first-order reaction in this case, which is in good accordance with Equation 7. Therefore, the
326
splitting of N2O (Reaction 2), rather than the recombination of O2 (Reaction 3), was the
327
rate-determining step of N2O decomposition over the Co3O4, CuxCoy and CuO catalysts. Because
328
if the recombination of O2 was the rate-determining step, then the reaction order of N2O
329
decomposition would be 2. Thus, the linear regression presented in Figure S6 was performed to
330
obtain the reaction rate constant of N2O decomposition, and the results are shown in Figure 5.
331
Hinted by Equations 7 and 8, the reaction rate constant of N2O decomposition (kN2O) positively
332
correlated with the equilibrium constant of N2O adsorption (K1), the kinetic constant of N2O
333
splitting (k2) and the quantity of active sites (Mn+). Generally, the kinetic constant of N2O splitting
334
(k2) increases with the temperature. Therefore, the N2O decomposition rate of all the materials
335
increased with the temperature (shown in Figure 5). The kinetic constant of N2O splitting (k2)
336
relates to the redox ability,21, 38 whereas the quantity of active sites (Mn+) mainly relates to the
337
metal cations in a low valence state (e.g., Cu+ and Co2+) on the surface. The results of the H2−TPR
338
study (Figure 2) suggest that the sequence of redox ability followed Cu2Co1>Cu1.5Co1.5≈Cu1Co2≈
339
CuO≫Co3O4. The AES and XPS results (Figure 3) demonstrate that the percentages of metal
340
cations in a low valence state on the surface of the CuxCoy catalysts were almost the same, but
341
these percentages were significantly higher than those for CuO and Co3O4. The equilibrium
342
constant of N2O adsorption (K1) is mainly related to the ability to adsorb N2O. The adsorptivity of 14
ACS Paragon Plus Environment
Page 15 of 32
Environmental Science & Technology
343
N2O was determined by N2O−TPD (Figure 4g), and the results show that the adsorptivity of N2O
344
decreased in the following sequence: Cu2Co1 > Co3O4 > Cu1.5Co1.5 > Cu1Co2 > CuO. CuO
345
presented the lowest N2O adsorptivity and redox cyclability, as well as a medium quantity of
346
active sites and redox property. Therefore, the lowest N2O decomposition performance of CuO
347
(shown in Figure 5) mainly resulted from its poor N2O adsorptivity and redox cycling between the
348
Cu2+/Cu+. Co3O4 presented a relatively high N2O adsorptivity as well as a medium quantity of
349
active sites. However, its redox ability was the weakest, which was mainly responsible for the
350
inferior N2O decomposition performance of Co3O4 (shown in Figure 5). For the CuxCoy catalysts,
351
the redox ability derived from CuO, the N2O adsorptivity derived from Co3O4, and the higher
352
quantities of active sites derived from the charge interaction between Co and Cu were all
353
responsible for the excellent N2O decomposition performance. Meanwhile, these properties in
354
Cu1.5Co1.5 and Cu1Co2 were almost the same, except the N2O adsorption capacity of Cu1.5Co1.5 was
355
slightly higher than that of Cu1Co2. This result proves that the N2O decomposition performance of
356
Cu1.5Co1.5 was slightly higher than that of Cu1Co2 (shown in Figure 5). Cu2Co1 sustained the
357
highest N2O adsorptivity, redox ability and quantity of active sites. Consequently, Cu2Co1 showed
358
the optimum performance for N2O decomposition (shown in Figure 5).
359
Based on the above results and conclusions, the N2O decomposition mechanism over CuxCoy
360
catalysts, and the key roles of CuO and Co3O4 in CuxCoy catalysts for N2O decomposition were
361
proposed (Scheme 1). The DFT calculation results suggest that Co3O4 provided abundant surface
362
oxygen vacancies, and thus, served as the major adsorption site of N2O. CuO was dispersed
363
around Co3O4 and provided high reducibility on the interface of Co3O4−CuOx, which promoted
364
the rate-determining step (N−O break) of N2O decomposition and left O in the defect sites.
365
Meanwhile, the charge interaction became stronger with increasing Cu content and promoted the
366
formation of Cu+ and Co2+, which performed as the active sites and adsorption sites, respectively.
367
Finally, the residual O in the defect sites recombined to release O2.
368
15
ACS Paragon Plus Environment
Environmental Science & Technology
369
Acknowledgements:
370
This work was financially supported by the National Key Research and Development Program
371
(2017YFC0210700 and 2017YFC0212804), and the National Natural Science Foundation of
372
China (21876093 and 21777081).
373
Supporting Information Available
374
This information is available free of charge via the Internet at http://pubs.acs.org/.
375
N2O decomposition performances of Co3O4, Cu1Co2, and CuO under different conditions. XRD
376
patterns of the Co3O4, CuxCoy, and CuO catalysts. XPS spectra of the CuxCoy and CuO catalysts
377
for the spectral region of Cu 2p3/2. TCD signals of N2O−TPD profiles over Co3O4, CuxCoy, and
378
CuO catalysts. MS spectra of N2O, O2, NO, and N2 over Co3O4, CuxCoy, and CuO during
379
N2O−TPD. N2 and NO desorption amounts during N2O−TPD over Co3O4, CuxCoy, and CuO
380
catalysts. Dependence of the N2O decomposition rate on the N2O concentration over the Co3O4,
381
CuxCoy, and CuO catalysts at 350−500 °C.
382
16
ACS Paragon Plus Environment
Page 16 of 32
Page 17 of 32
Environmental Science & Technology
383
References:
384
(1) Hungate, B. A.; Dukes, J. S.; Shaw, M. R.; Luo, Y.; Field, C. B. Nitrogen and climate change.
385
Science 2003, 302, 1512-1513.
386
(2) Li, L.; Xu, J. H.; Hu, J. X.; Han, J. R. Reducing nitrous oxide emissions to mitigate climate
387
change and protect the ozone layer. Environ. Sci. Technol. 2014, 48, 5290-5297.
388
(3) Zabilskiy, M.; Djinović, P.; Erjavec, B.; Dražić, G.; Pintar, A. Small CuO clusters on CeO2
389
nanospheres as active species for catalytic N2O decomposition. Appl. Catal. B 2015, 163,
390
113-122.
391
(4) Ravishankara, A. R.; Daniel, J. S.; Portmann, R. W. Nitrous oxide (N2O): the dominant
392
ozone-depleting substance emitted in the 21st century. Science 2009, 326, 123-125.
393
(5) Dandekar, A.; Vannice, M. A. Decomposition and reduction of N2O over copper catalysts.
394
Appl. Catal. B 1999, 22, 179-200.
395
(6) Liu, Z. M.; He, F.; Ma, L. L.; Peng, S. Recent advances in catalytic decomposition of N2O on
396
noble metal and metal oxide catalysts. Catal. Surv. Asia. 2016, 20, 121-132.
397
(7) Komvokis, V. G.; Marti, M.; Delimitis, A.; Vasalos, I. A.; Triantafyllidis, K. S. Catalytic
398
decomposition of N2O over highly active supported Ru nanoparticles (≤3nm) prepared by
399
chemical reduction with ethylene glycol. Appl. Catal. B 2011, 103, 62-71.
400
(8) Kawi, S.; Liu, S. Y.; Shen, S. C. Catalytic decomposition and reduction of N2O on
401
Ru/MCM-41 catalyst. Catal. Today. 2001, 68, 237-244.
402
(9) Parres-Esclapez, S.; Illán-Gómez, M. J.; de Lecea, C. S.-M.; Bueno-López, A. On the
403
importance of the catalyst redox properties in the N2O decomposition over alumina and ceria
404
supported Rh, Pd and Pt. Appl. Catal. B 2010, 96, 370-378. 17
ACS Paragon Plus Environment
Environmental Science & Technology
405
(10) Rutkowska, M.; Piwowarska, Z.; Micek, E.; Chmielarz, L. Hierarchical Fe-, Cu- and Co-Beta
406
zeolites obtained by mesotemplate-free method. Part I: Synthesis and catalytic activity in N2O
407
decomposition. Micropor. Mesopor. Mat. 2015, 209, 54-65.
408
(11) Jisa, K.; Novakova, J.; Schwarze, M.; Vondrova, A.; Sklenak, S.; Sobalik, Z. Role of the
409
Fe-zeolite structure and iron state in the N2O decomposition: Comparison of Fe-FER, Fe-BEA,
410
and Fe-MFI catalysts. J. Catal. 2009, 262, 27-34.
411
(12) Pirngruber, G. D.; Pieterse, J. A. Z. The positive effect of NO on the N2O decomposition
412
activity of Fe-ZSM-5: A combined kinetic and in situ IR spectroscopic study. J. Catal. 2006, 237,
413
237-247.
414
(13) Yao, X. J.; Tang, C. J.; Gao, F.; Dong, L. Research progress on the catalytic elimination of
415
atmospheric molecular contaminants over supported metal-oxide catalysts. Catal. Sci. Technol.
416
2014, 4, 2814-2829.
417
(14) Konsolakis, M.; Sgourakis, M.; Carabineiro, S. A. C. Surface and redox properties of cobalt–
418
ceria binary oxides: On the effect of Co content and pretreatment conditions. Appl. Surf. Sci. 2015,
419
341, 48-54.
420
(15) Konsolakis, M. Recent advances on nitrous oxide (N2O) decomposition over non-noble-metal
421
oxide catalysts: Catalytic performance, mechanistic considerations, and surface chemistry aspects.
422
ACS Catalysis 2015, 5, 6397-6421.
423
(16) Pasha, N.; Lingaiah, N.; Babu, N.; Reddy, P.; Prasad, P. Studies on cesium doped cobalt
424
oxide catalysts for direct N2O decomposition in the presence of oxygen and steam. Catal.
425
Commun. 2008, 10, 132-136.
426
(17) Xue, L.; Zhang, C. B.; He, H.; Teraoka, Y. Catalytic decomposition of N2O over CeO2 18
ACS Paragon Plus Environment
Page 18 of 32
Page 19 of 32
Environmental Science & Technology
427
promoted Co3O4 spinel catalyst. Appl. Catal. B 2007, 75, 167-174.
428
(18) Russo, N.; Fino, D.; Saracco, G.; Specchia, V. N2O catalytic decomposition over various
429
spinel-type oxides. Catal. Today. 2007, 119, 228-232.
430
(19) Piskorz, W.; Zasada, F.; Stelmachowski, P.; Kotarba, A.; Sojka, Z. DFT modeling of reaction
431
mechanism and ab initio microkinetics of catalytic N2O decomposition over alkaline earth oxides:
432
From molecular orbital picture account to simulation of transient and stationary rate profiles. J.
433
Phys. Chem. C 2013, 117, 18488-18501.
434
(20) Piskorz, W.; Zasada, F.; Stelmachowski, P.; Kotarba, A.; Sojka, Z. Decomposition of N2O
435
over the surface of cobalt spinel: A DFT account of reactivity experiments. Catal. Today. 2008,
436
137, 418-422.
437
(21) Zhou, H. B.; Huang, Z.; Sun, C.; Qin, F.; Xiong, D. S.; Shen, W.; Xu, H. L. Catalytic
438
decomposition of N2O over CuxCe1−xOy mixed oxides. Appl. Catal. B 2012, 125, 492-498.
439
(22) Zabilskiy, M.; Djinović, P.; Tchernychova, E.; Tkachenko, O. P.; Kustov, L. M.; Pintar, A.
440
Nanoshaped CuO/CeO2 materials: Effect of the exposed ceria surfaces on catalytic activity in N2O
441
decomposition reaction. ACS Catalysis 2015, 5, 5357-5365.
442
(23) Kapteijn, F.; Marbán, G.; Rodriguez-Mirasol, J.; Moulijn, J. A. Kinetic Analysis of the
443
Decomposition of Nitrous Oxide over ZSM-5 Catalysts. J. Catal. 1997, 167, 256-265.
444
(24) Obalová, L.; Fíla, V. Kinetic analysis of N2O decomposition over calcined hydrotalcites.
445
Appl. Catal. B 2007, 70, 353-359.
446
(25) Liu, X. H.; Qiu, G. Z.; Li, X. G. Shape-controlled synthesis and properties of uniform spinel
447
cobalt oxide nanocubes. Nanotechnology 2005, 16, 3035-3040.
448
(26) Shojaee, K.; Montoya, A.; Haynes, B. S. Insight into oxygen stability and vacancy formation 19
ACS Paragon Plus Environment
Environmental Science & Technology
449
on Co3O4 model slabs. Computational Materials Science 2013, 72, 15-25.
450
(27) Maimaiti, Y.; Nolan, M.; Elliott, S. D. Reduction mechanisms of the CuO(111) surface
451
through surface oxygen vacancy formation and hydrogen adsorption. Phys. Chem. Chem. Phys.
452
2014, 16, 3036-3046.
453
(28) Yang, B. X.; Ye, L. P.; Gu, H. J.; Huang, J. H.; Li, H. Y.; Luo, Y. A density functional theory
454
study of CO oxidation on CuO1-x(111). J. Mol. Model. 2015, 21, 195.
455
(29) Zasada, F.; Piskorz, W.; Sojka, Z. Cobalt spinel at various redox conditions: DFT+U
456
investigations into the structure and surface thermodynamics of the (100) facet. J. Phys. Chem. C
457
2015, 119, 19180-19191.
458
(30) Xiong, S. C.; Xiao, X.; Liao, Y.; Dang, H.; Shan, W. P.; Yang, S. J. Global kinetic study of
459
NO reduction by NH3 over V2O5-WO3/TiO2: Relationship between the SCR performance and the
460
key factors. Ind. Eng. Chem. Res. 2015, 54, 11011-11023.
461
(31) Xiong, S. C.; Weng, J. X.; Liao, Y.; Li, B.; Zou, S. J.; Geng, Y.; Xiao, X.; Huang, N.; Yang,
462
S. J. Alkali metal deactivation on the low temperature selective catalytic reduction of NOx with
463
NH3 over MnOx-CeO2: A mechanism study. J. Phys. Chem. C 2016, 120, 15299-15309.
464
(32) Wang, D.; Peng, Y.; Xiong, S. C.; Li, B.; Gan, L. N.; Lu, C. M.; Chen, J. J.; Ma, Y. L.; Li, J.
465
H. De-reducibility mechanism of titanium on maghemite catalysts for the SCR reaction: An in situ
466
DRIFTS and quantitative kinetics study. Appl. Catal. B 2018, 221, 556-564.
467
(33) Pérez-Ramı́rez, J.; Kapteijn, F.; Schöffel, K.; Moulijn, J. A. Formation and control of N2O in
468
nitric acid production. Appl. Catal. B 2003, 44, 117-151.
469
(34) Zabilskiy, M.; Djinović, P.; Tchernychova, E.; Pintar, A. N2O decomposition over CuO/CeO2
470
catalyst: New insights into reaction mechanism and inhibiting action of H2O and NO by operando 20
ACS Paragon Plus Environment
Page 20 of 32
Page 21 of 32
Environmental Science & Technology
471
techniques. Appl. Catal. B 2016, 197, 146-158.
472
(35) Pasha, N.; Lingaiah, N.; Siva Sankar Reddy, P.; Sai Prasad, P. S. An investigation into the
473
effect of Cs promotion on the catalytic activity of NiO in the direct decomposition of N2O. Catal.
474
Lett. 2007, 118, 64-68.
475
(36) Perezalonso, F.; Meliancabrera, I.; Lopezgranados, M.; Kapteijn, F.; Fierro, J. Synergy of
476
FexCe1−xO2 mixed oxides for N2O decomposition. J. Catal. 2006, 239, 340-346.
477
(37) Bloemen, P. J. H.; van de Vorst, M. T. H.; Johnson, M. T.; Coehoorn, R.; de Jonge, W. J. M.
478
Magnetic layer thickness dependence of the interlayer exchange coupling in (001) Co/Cu/Co. J.
479
Appl. Phys. 1994, 76, 7081-7083.
480
(38) Konsolakis, M.; Carabineiro, S. A. C.; Papista, E.; Marnellos, G. E.; Tavares, P. B.; Moreira,
481
J. A.; Romaguerabarcelay, Y.; Figueiredo, J. L. Effect of preparation method on the solid state
482
properties and the deN2O performance of CuO–CeO2 oxides. Catal. Sci. Technol. 2015, 5,
483
3714-3727.
484
(39) Svintsitskiy, D. A.; Kardash, T. Y.; Stonkus, O. A.; Slavinskaya, E. M.; Stadnichenko, A. I.;
485
Koscheev, S. V.; Chupakhin, A. P.; Boronin, A. I. In situ XRD, XPS, TEM, and TPR study of
486
highly active in CO oxidation CuO nanopowders. J. Phys. Chem. C 2013, 117, 14588-14599.
487
(40) Timmermans, B.; Reniers, F.; Hubin, A.; Buess-Herman, C. Chemical effects in the Auger
488
spectrum of copper-oxygen compounds. Appl. Surf. Sci. 1999, 144-145, 54-58.
489
(41) Fierro, G.; Lo Jacono, M.; Inversi, M.; Dragone, R.; Porta, P. TPR and XPS study of
490
cobalt-copper mixed oxide catalysts: evidence of a strong Co-Cu interaction. Top. Catal. 2000, 10,
491
39-48.
492
(42) Grzybek, G.; Stelmachowski, P.; Gudyka, S.; Indyka, P.; Sojka, Z.; Guillén-Hurtado, N.; 21
ACS Paragon Plus Environment
Environmental Science & Technology
493
Rico-Pérez, V.; Bueno-López, A.; Kotarba, A. Strong dispersion effect of cobalt spinel active
494
phase spread over ceria for catalytic N2O decomposition: The role of the interface periphery. Appl.
495
Catal. B 2016, 180, 622-629.
496
(43) Piskorz, W.; Zasada, F.; Stelmachowski, P.; Diwald, O.; Kotarba, A.; Sojka, Z.
497
Computational and experimental investigations into N2O decomposition over MgO nanocrystals
498
from thorough molecular mechanism to ab initio microkinetics. J. Phys. Chem. C 2011, 115,
499
22451-22460.
500
(44) Liu, X.; Yang, Z. Y.; Li, Y. P.; Zhang, F. Z. Theoretical study of N2O decomposition
501
mechanism over binuclear Cu-ZSM-5 zeolites. J. Mol. Catal. A 2015, 396, 181-187.
502
(45) Yamashita, T.; Vannice, A. N2O decomposition over manganese oxides. J. Catal. 1996, 161,
503
254-262.
504
(46) Smeets, P.; Sels, B.; Vanteeffelen, R.; Leeman, H.; Hensen, E.; Schoonheydt, R. The catalytic
505
performance of Cu-containing zeolites in N2O decomposition and the influence of O2, NO and
506
H2O on recombination of oxygen. J. Catal. 2008, 256, 183-191.
507
(47) Wu, L. N.; Qin, W.; Hu, X. Y.; Ju, S. D.; Dong, C. Q.; Yang, Y. P. Decomposition and
508
reduction of N2O on CaS (100) surface: A theoretical account. Surf. Sci. 2015, 632, 83-87.
509
22
ACS Paragon Plus Environment
Page 22 of 32
Page 23 of 32
Environmental Science & Technology
510
Table Captions
511
Table 1. Crystal sizes, crystal parameters and BET surface areas of the Co3O4, CuxCoy, and CuO
512
catalysts.
513
Table 2. Surface chemical compositions of the Co3O4, CuxCoy, and CuO catalysts.
514
23
ACS Paragon Plus Environment
Environmental Science & Technology
Page 24 of 32
515
Table 1. Crystal sizes, crystal parameters and BET surface areas of the Co3O4, CuxCoy, and CuO
516
catalysts. crystal size a /nm
Co3O4
517
BET surface area /m2 g−1
CuO
Co3O4
CuO
Co3O4
−
21
−
a=b=c=0.8090, α=β=γ=90o
27
a=b=c=0.8098, α=β=γ=90o
30
a=b=c=0.8097, α=β=γ=90o
28
a=b=c=0.8095, α=β=γ=90o
40
−
8.1
Cu1Co2
22
18
Cu1.5Co1.5
24
17
Cu2Co1
14
13
CuO
27
−
a Calculated
crystal parameter a /nm
a=0.4689, b=0.3433, c=0.5137, α=γ=90o, β= 99.44o a=0.4678, b=0.3446, c=0.5127, α=γ=90o, β= 99.47o a=0.4687, b=0.3399, c=0.5104, α=γ=90o, β= 99.49o a=0.4687, b=0.3420, c=0.5135, α=γ=90o, β= 99.37o
from the XRD patterns.
518
24
ACS Paragon Plus Environment
Page 25 of 32
519
Environmental Science & Technology
Table 2. Surface chemical compositions of the Co3O4, CuxCoy, and CuO catalysts /%. Cua
520 521 522
Cob Cu+
Co3+
Co2+
Co3O4
−
−
13.2
7.7
79.1
Cu1Co2
9.8
3.8
11.6
9.5
65.3
Cu1.5Co1.5
13.4
4.7
10.3
8.5
63.1
Cu2Co1
17.1
6.9
8.5
6.6
60.9
CuO
36.5
4.8
−
−
58.7
a Calculated b Obtained
Ob
Cu2+
from the AES spectra of Cu LMM. from the XPS spectra of Co 2p3/2.
25
ACS Paragon Plus Environment
Environmental Science & Technology
523
Figure Captions
524
Figure 1. (a) N2O decomposition performance of the Co3O4, CuxCoy, and CuO catalysts. (b) N2O
525
decomposition performance of Cu2Co1 under different conditions. Reaction conditions: [N2O] =
526
1000 ppm, [O2] = 2% (when used), [NO] = 200 ppm (when used), [H2O] = 0.5% (when used),
527
catalyst mass = 100 mg, flow rate = 100 ml min−1, and GHSV=60,000 cm3 g−1 h−1.
528
Figure 2. (a) H2−TPR profiles of the Co3O4, CuxCoy, and CuO catalysts. (b) The initial H2
529
consumption rates of the CuxCoy and CuO catalysts in the H2−TPR study.
530
Figure 3. (a) AES spectra of the CuxCoy and CuO catalysts for the spectral region of the Cu
531
LMM. (b) XPS spectra of the Co3O4 and CuxCoy catalysts for the spectral region of the Co 2p3/2.
532
Figure 4. Model structures of N2O adsorbed on CuO: (a) CuO with an oxygen vacancy (b), Co3O4
533
(c), and Co3O4 with an oxygen vacancy (d−f). The white balls represent N, red balls represent O,
534
the blue balls represent Cu, and the navy-blue balls represent Co. (g) N2O desorption amounts
535
during N2O−TPD over Co3O4, CuxCoy, and CuO catalysts.
536
Figure 5. N2O decomposition rate constants of the Co3O4, CuxCoy, and CuO catalysts.
537
Scheme 1. The key roles of CuO and Co3O4 in the CuxCoy catalysts for N2O decomposition.
538 539 540 541
26
ACS Paragon Plus Environment
Page 26 of 32
Page 27 of 32
Environmental Science & Technology
N2O decomposition/%
100 80 60
Co3O4 Cu1Co2 Cu1.5Co1.5 Cu2Co1 CuO
40 20 0
N2O decomposition/%
300
100
350
400
450
Temperature/oC (a)
no addition with H2O
500
with O2 with NO with O2+NO+H2O
80 60 40 20 0
350
400
Temperature/oC
450
542
(b) Figure 1. (a) N2O decomposition performance of the Co3O4, CuxCoy, and CuO catalysts. (b) N2O
543
decomposition performance of Cu2Co1 under different conditions. Reaction conditions: [N2O] =
544
1000 ppm, [O2] = 2% (when used), [NO] = 200 ppm (when used), [H2O] = 0.5% (when used),
545
catalyst mass = 100 mg, flow rate = 100 ml min−1, and GHSV=60,000 cm3 g−1 h−1.
546
27
ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 32
215 158
Intensity/a.u.
Cu2Co1
213 184
Cu1.5Co1.5 Cu1Co2
367 320
100
Initial H2 consumption rate/mol g-1 s-1
CuO
202
200
300
Co3O4 400
Temperature/oC (a)
4.0
500
600
Cu1Co2 Cu1.5Co1.5 Cu2Co1 CuO
3.2 2.4 1.6 0.8 0.0
2.28
2.32
2.36
2.40
2.44
2.48
547
1000 T-1/K-1 (b) Figure 2. (a) H2−TPR profiles of the Co3O4, CuxCoy, and CuO catalysts. (b) The initial H2
548
consumption rates of the CuxCoy and CuO catalysts in the H2−TPR study.
549
28
ACS Paragon Plus Environment
Page 29 of 32
Environmental Science & Technology
550 Co 2p3/2
Cu LMM +
Cu
Satellites
+
Cu1.5Co1.5Cu Cu1Co2
Cu2Co1
Intensity/a.u.
Intensity/a.u.
Cu2Co1
Cu+
Co2+
Cu1.5Co1.5
Satellites
Cu1Co2
Cu
Co3O4
CuO
Co2+
Satellites
+
908
Co2+ 779.2 eV
Co2+
779.7 eV
Satellites
912
916
792
920
787
782
777
Binding Energy/eV
Kinetic Energy/eV
551
(a) (b) Figure 3. (a) AES spectra of the CuxCoy and CuO catalysts for the spectral region of the Cu
552
LMM. (b) XPS spectra of the Co3O4 and CuxCoy catalysts for the spectral region of the Co 2p3/2.
553
29
ACS Paragon Plus Environment
Environmental Science & Technology
554
555 556
Figure 4. Model structures of N2O adsorbed on CuO: (a) CuO with an oxygen vacancy (b), Co3O4
557
(c), and Co3O4 with an oxygen vacancy (d−f). The white balls represent N, red balls represent O,
558
the blue balls represent Cu, and the navy-blue balls represent Co. (g) N2O desorption amounts
559
during N2O−TPD over Co3O4, CuxCoy, and CuO catalysts.
30
ACS Paragon Plus Environment
Page 30 of 32
Page 31 of 32
Environmental Science & Technology
561 562
Figure 5. N2O decomposition rate constants of the Co3O4, CuxCoy, and CuO catalysts.
563
31
ACS Paragon Plus Environment
Environmental Science & Technology
564 565
Scheme 1. The key roles of CuO and Co3O4 in the CuxCoy catalysts for N2O decomposition.
566
32
ACS Paragon Plus Environment
Page 32 of 32