Subscriber access provided by CORNELL UNIVERSITY LIBRARY
Article
Catalytic Oxidation of Chlorobenzene over MnxCe1-xO2/ HZSM-5 Catalysts: A Study with Practical Implications Xiaole Weng, Pengfei Sun, Yu Long, Qingjie Meng, and Zhongbiao Wu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b06585 • Publication Date (Web): 14 Jun 2017 Downloaded from http://pubs.acs.org on June 17, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
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 28
Environmental Science & Technology
1
Catalytic Oxidation of Chlorobenzene over MnxCe1-xO2/HZSM-5
2
Catalysts: A Study with Practical Implications
3
Xiaole Weng a,b, Pengfei Suna, Yu Long a, Qingjie Menga and Zhongbiao Wu*a,b
4
a Key Laboratory of Environment Remediation and Ecological Health, Ministry of
5
Education, College of Natural Resources and Environmental Science, Zhejiang
6
University, 310058 Hangzhou, P. R. China.
7
b Zhejiang Provincial Engineering Research Centre of Industrial Boiler & Furnace
8
Flue Gas Pollution Control, 388 Yuhangtang Road, 310058 Hangzhou, P. R. China.
9 10
Corresponding author: Dr. Zhongbiao Wu, Fax/Tel: 0086-571-88982863; E-mail:
[email protected].
11
Abstract
12
Industrial-use catalysts usually encounter severe deactivation after long-term
13
operation for catalytic oxidation of chlorinate volatile organic compounds (CVOCs),
14
which becomes a “bottleneck” for large-scale application of catalytic combustion
15
technology.
16
MnxCe1-xO2/HZSM-5 were investigated for the catalytic oxidation of chlorobenzene
17
(CB). The activation energy (Ea), Brønsted and Lewis acidities, CB adsorption and
18
activation behaviors, long-term stabilities, and surficial accumulation compounds
19
(after ageing) were studied using a range of analytical techniques, including XPS,
20
H2-TPR, pyridine-IR, DRIFT, and O2-TP-Ms. Experimental results revealed that the
21
Brønsted/Lewis (B/L) ratio of MnxCe1-xO2/HZSM-5 catalysts could be adjusted by ion
22
exchange of H· in HZSM-5 with Mnn+ (where the exchange with Ce4+ did not
23
distinctly affect the acidity); the long-term aged catalysts could accumulate ca. 14
24
organic compounds at surface, which included highly toxic tetrachloromethane,
25
trichloroethylene,
26
operational environment could ensure a stable performance for MnxCe1-xO2/HZSM-5
27
catalysts; this was due to the effective removal of Cl· and coke accumulations by H2O
In
this
work,
typical
tetrachloroethylene,
acidic
solid-supported
o-Dichlorobenzene,
1 / 28
ACS Paragon Plus Environment
etc.;
catalysts
high
of
humid
Environmental Science & Technology
28
washing, and the distinct increase of Lewis acidity by the interaction of H2O with
29
HZSM-5. This work gives an in-depth view into the CB oxidation over acidic
30
solid-supported catalysts and might provide practical guidelines for the rational design
31
of reliable catalysts for industrial applications.
32 33
Key words: Chlorobenzene; Catalytic Oxidation; CeO2; MnOx; Stability; HZSM-5
34 35
TOC/Abstract Art
36 37
1. Introduction
38
Dioxins are persistent air pollutants that can cause cancer in high doses. They are
39
usually released from municipal solid waste (MSW) incineration and manufacturing
40
processes [1-3]. Catalytic combustion of dioxins and analogue compounds into
41
harmless CO2 and H2O is an efficient measure to reduce dioxin emissions [4-6].
42
In the laboratory, chlorobenzene (CB) is usually selected as an indicator to study
43
the efficiency of catalysts developed for the catalytic oxidation of dioxins [7-9]. The
44
oxidation of CB generally involves three sequential reaction processes: Cl· adsorption,
45
aromatic ring cleavage, and deep oxidation (into CO2 and H2O) [10, 11]. The
46
Cl· adsorption process is best with catalysts that have Brønsted acid sites, the
47
presence of which can induce a nucleophilic substitution between the Cl· and H· that 2 / 28
ACS Paragon Plus Environment
Page 2 of 28
Page 3 of 28
Environmental Science & Technology
48
converts the CB into phenolates and facilitates the ring cleavage [12-14]. Therefore,
49
acidic solids with abundant Brønsted acid sites (e.g. HZSM-5, H-MOR, and H-BETA
50
[12, 15, and 16]) are usually selected to enhance the adsorption process. Aromatic ring
51
cleavage and deep oxidation processes are both best with the use of catalysts that have
52
high redox activities, and transition metal oxides (e.g., CeO2, CrOx, MnOx, FeOx, and
53
their bimetallic oxides [17-19]) are usually selected for this purpose.
54
In literature, catalysts consisting of acidic solids and transition metal oxides are
55
extensively studied for the catalytic oxidation of volatile organic compounds (VOCs)
56
[20], which lead to promising outcomes in terms of high activity and CO2 selectivity.
57
However, most of the reported work did not provide all insights into chlorine
58
by-products, CVOCs adsorption and activation, and Cl· and coke accumulations.
59
Particularly, identification of the long-term stabilities, deactivation mechanisms for
60
recovery and surficial accumulation compounds (after ageing) for waste disposal of
61
developed catalysts has been lacking in previous investigations. Many studies are
62
therefore insufficient to provide practical guidelines for industry, making the
63
engineering technicians unable to practically estimate the efficiencies of catalysts
64
used in industry and, as a result, many current industrial catalysts undergo severe
65
deactivation after a long-term operation.
66
In our previous study [21], we reported typical acid-solid supported catalysts,
67
MnxCe1-xO2/HZSM-5, for the catalytic oxidation of CB and assessed their activities,
68
HCl and CO2 selectivity, and CB oxidation mechanisms. In this study, we further
69
investigated the activation energy (Ea), Brønsted and Lewis acidities, CB adsorption
70
and activation behaviors, long-term stabilities (in both dry and humid conditions), Cl⋅
71
and coke accumulations and surficial accumulation compounds (after ageing) of the
72
catalysts. Our goal in this work is to give an in-depth view into the CB oxidation over
73
these acidic solid-supported catalysts and to estimate the practicality of using these
74
catalysts in industry.
3 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
75
2. Experimental
76
2.1 Catalyst syntheses
77
MnxCe1-xO2/HZSM-5 catalysts were synthesized via a wet impregnation route
78
where precisely measured Mn(NO3)2, Ce(NO3)3, and HZSM-5 were mixed in ethanol,
79
following by continuous stirring for 5 h. The mixture was then dried at 110 °C for 10
80
h and calcinated at 550 °C for 5 h in static air. The HZSM-5, with a Si:Al ratio of 30,
81
was supplied by Zhiyuan Molecular Co., Ltd (Shanghai, P. R. China). The resultant
82
products were denoted as MnxCe1-xO2/HZSM-5, with x values of 1, 0.8, and 0.
83
The Mn0.8Ce0.2O2 catalyst was synthesized as follows. An aqueous solution
84
containing proper amounts of Mn(NO3)2, Ce(NO3)3, and citric acid (citric
85
acid/(Mn+Ce) = 0.3 in molar ratio) was continuously stirred for 2 h at 50 °C. The
86
mixture was then dried at 110 °C for 10 h and calcinated at 550 °C for 5 h in static air.
87
The aforementioned metal salts (> 99.9%) were all supplied by SCRC (Sinopham
88
Chemical Reagent Co., Ltd, P. R. China) and used as obtained.
89
2.2 Activity measurements
90
All catalysts were sieved with 40-60 mesh and placed in a quartz fixed-bed
91
reactor. The reaction temperature was monitored using a thermocouple loaded in the
92
core of a catalyst bed with a measured temperature range of 150-400 °C. The reaction
93
feed consisted of 1000 ppm CB and 10 vol% of O2/N2 with a weight hourly space
94
velocity (WHSV) of 15,000 mL g-1 h-1. The inlet and outlet concentrations of CB were
95
analyzed online by a gas chromatograph (GC, Agilent 6890, America) equipped with
96
a flame ionization detector (FID).
97
Humidity was achieved by bubbling the N2 through water to yield a humidity of
98
approximately 10 vol% (as measured by a Hygrograph instrument, Testo 480,
99
Germany).
4 / 28
ACS Paragon Plus Environment
Page 4 of 28
Page 5 of 28
Environmental Science & Technology
100
2.3 Characterizations
101
H2 temperature-programmed reduction (H2-TPR) was conducted using an
102
automatic multi-purpose adsorption instrument (TP-5079, Xianqua, Tianjin, P. R.
103
China) equipped with a custom-made thermal conductivity detector (TCD). In each
104
measurement, precisely weighted 50 mg catalyst was first pretreated in a purge of He
105
at 400 °C for 1 h and then naturally cooled to room temperature. The catalyst was
106
then subjected to a purge of 6 vol% H2/N2 at a flow rate of 30 mL min-1 and heated
107
from room temperature to 800 °C at a linear heating rate of 10 °C min-1. The
108
consumption of H2 was quantitatively measured by integrating the corresponding TPR
109
profiles and using the CuO as a standard agent for calibration.
110
X-ray photoelectron spectroscopy (XPS) was conducted using a Thermo
111
ESCALAB 250 spectrometer (U.S.A.) with an Al Kα X-ray (hν = 1486.6 eV)
112
radiation excitation source. After calibrating the binding energy scale using the signal
113
of adventurous carbon (a binding energy of 284.8 eV), the curves were fitted using a
114
Shirley background and a Gaussian peak shape with 20 % Lorentzian character.
115
Pyridine adsorbed IR spectroscopy (Py-IR) was conducted using an FT-IR
116
(Tensor 27, Bruker, Germany) equipped with a custom made IR cell that was
117
connected by a vacuum adsorption apparatus. The measurement was carried out in
118
situ and the spectra were recorded at a resolution of 4 cm-1. The catalyst was first
119
heated at a rate of 10 °C min-1 to 400 °C and then cooled to room temperature in a
120
vacuum (10-3 Pa). After that, pyridine vapor was introduced until the adsorption
121
approached saturation. The desorption process was conducted by heat-treatment of the
122
adsorbed catalyst at a linear heating rate of 10 °C min-1 to 450 °C.
123
In situ DRIFT was conducted using a Nicolet 6700 FT-IR spectrometer equipped
124
with an MCT detector. A DRIFT cell (Harrick) with CaF2 windows was fitted into a
125
heating cartridge that allowed the catalyst being heated to 400 °C in atmospheric
126
conditions. In each measurement, the catalyst was first pretreated in a flow of He
127
(99.99%, 100 mL min-1) at a temperature of 400 °C for 1 h and then naturally cooled
128
to 200 °C. Thereafter, 100 ppm CB was introduced for 15 min and then together with 5 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
129
a carrier gas of 10 vol% O2/N2 for another 15 min. The spectra (average of 32 scans
130
at 4 cm-1 resolution) were recorded simultaneously in each run.
131
2.4 Stability measurements
132
Surficial accumulation compounds of aged catalysts were analyzed using a
133
GC-MS instrument (Agilent 7890A GC equipped with Agilent 5975C MS, U.S.A.).
134
The catalyst was desorbed in a thermal desorption instrument (TDI, PERSEE-TP7, P.
135
R. China) where the released organic and inorganic species were simultaneously
136
analyzed using a GC-MS.
137
O2-TP-MS experiment was carried out on a custom made setup (TP-5079, Tianjin
138
Xianquan Co., Ltd., China), connected with a mass spectrometer (HIDEN QGA, UK).
139
The amount of 50 mg catalyst was first pretreated in a pure He gas flow at 200 °C for
140
1 h and then cooled to room temperature. Thereafter, catalyst was heated in a flow of
141
5% O2/He from 100 to 900 °C with a heating rate at 10 °C min-1. The signals of
142
desorbed CO2, CO, and HCl were recorded using an MS.
143
SEM-EDS mapping was carried out using an SU8010 FE-SEM operated at an
144
accelerating voltage of 3 kV with an EDS detector.
145
3. Results and Discussion
146
3.1 Apparent activation energy measurements
147
The apparent activation energy (Ea) for CB oxidation over MnxCe1-xO2/HZSM-5
148
and Mn0.8Ce0.2O2 catalysts were calculated using a first-order Arrhenius equation: r =
149
-kc = (-Aexp(-Ea/RT))c, where c, r, k, and A were the CB concentration (µmol g-1),
150
reaction rate (µmol g-1 s-1), rate constant (s-1), and pre-exponential factor, respectively
151
[22] [23].
152
Fig. 1 shows the Arrhenius plots of Mn0.8Ce0.2O2, Mn0.8Ce0.2O2/HZSM-5,
153
MnOx/HZSM-5, and CeO2/HZSM-5 catalysts with CB conversion at CeO2/HZSM-5 >
157
MnOx/HZSM-5 > Mn0.8Ce0.2O2 > Mn0.8Ce0.2O2/HZSM-5. The lowest Ea value, for
158
Mn0.8Ce0.2O2/HZSM-5, indicated that the presence of HZSM-5 and the coexistence of
159
Ce and Mn oxides both benefited the activation and oxidation of CB. To elucidate the
160
cause for such a low Ea of the Mn0.8Ce0.2O2/HZSM-5 catalyst, a series of analyses
161
were then conducted to assess the influence factors, including pore structure, redox
162
potential, oxygen mobility, and reactant absorbability, etc. [24]
163 164
Fig. 1 Arrhenius plots of ln R versus 1000/T for the Mn0.8Ce0.2O2, Mn0.8Ce0.2O2/HZSM-5,
165
MnOx/HZSM-5, CeO2/HZSM-5 catalysts and HZSM-5 in the catalytic oxidation of CB with
166
WHSV at 15000 mL g-1 h-1.
167
3.2 Catalyst characterizations
168
3.2.1 H2-TPR measurements
169
The redox potentials of MnxCe1-xO2/HZSM-5 and Mn0.8Ce0.2O2 catalysts was
170
evaluated by using H2-TPR measurements, which were conducted in the temperature
171
range of 150-600 °C. As shown in Fig. 2, the CeO2/HZSM-5 catalyst did not show
172
obvious H2 consumption peaks within the investigated temperature range. However,
173
with a magnified H2-TPR profile (see supplementary Fig. S2), a weak reduction peak 7 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
174
located within 400-650 °C could be observed, which originated from the reduction of
175
surficial CeO2 in the catalyst [25]. The MnOx/HZSM-5 catalyst revealed two distinct
176
H2 reduction peaks centered at approximately 365 and 453 °C. The former was
177
assigned to the reduction of Mn4+ to Mn3+ and the latter was ascribed to the reduction
178
of Mn3+ to Mn2+ [26, 27]. The extra peak at approximately 237 °C was a result of the
179
reduction of Mn3+ in tetrahedral sites, according to Stobbe et al. [28] and Weimin et al.
180
[29]. For Mn0.8Ce0.2O2 and Mn0.8Ce0.2O2/HZSM-5 catalysts, two distinct H2
181
consumption peaks were also observed. Again, the former originated from the
182
reduction of Mn4+ to Mn3+, but the latter could be an overlap peak resulting from the
183
reduction of both surficial Ce4+ to Ce3+ and Mn3+ to Mn2+ [26, 27].
184 185
Fig. 2 H2-TPR profiles of the MnxCe1-xO2/HZSM-5 and Mn0.8Ce0.2O2 catalysts.
186
From Fig. 2, it can be seen that the Mn0.8Ce0.2O2/HZSM-5 catalyst had a much
187
higher redox potential than the MnOx/HZSM-5 as a distinct low-temperature peak
188
shift was observed in the Mn0.8Ce0.2O2/HZSM-5 catalyst. The reason could be
189
ascribed to the presence of a Ce4+/Ce3+ labile cycle that enhanced the reducibility of
190
neighboring Mnn+ in MnOx [30, 31]. Moreover, the Mn0.8Ce0.2O2/HZSM-5 catalyst
191
did not show any obvious difference in redox potential as compared with that of
192
Mn0.8Ce0.2O2. Only a slight high-temperature shift was observed in the peak assigned
193
to the surficial Ce4+ to Ce3+ and Mn3+ to Mn2+ reductions (where the peak for Mn4+ to
194
Mn3+ reduction remained unchanged). This implied that the addition of HZSM-5 8 / 28
ACS Paragon Plus Environment
Page 8 of 28
Page 9 of 28
Environmental Science & Technology
195
might only affect the reduction of surficial Ce4+ to Ce3+ that shifted the corresponding
196
peak to a lower temperature range.
197
The amounts of cumulative hydrogen consumption per gram by all of the
198
catalysts were calculated via integration of the corresponding TPR peaks using CuO
199
as standard agent for calibration. As shown in Table 1, the Mn0.8Ce0.2O2 catalyst
200
revealed the highest cumulative value at approximately 1.39 mol g-1. However, if
201
taking the 80 wt% HZSM-5 into account (note: the HZSM-5 could not uptake H2), the
202
Mn0.8Ce0.2O2/HZSM-5 should have a much higher cumulative H2 consumption per
203
gram of Mn0.8Ce0.2O2 than the Mn0.8Ce0.2O2 alone. This was assumed because the ion
204
exchange between the metal ions (M+) and H· (in the HZSM-5) induced a high
205
dispersion of Mn0.8Ce0.2O2 (after calcination) over the HZSM-5 support [32, 33].
206
Indeed, our SEM-EDS mapping (see supplementary Fig. S3) had revealed that the Mn
207
and Ce elemental dots in the Mn0.8Ce0.2O2/HZSM-5 catalyst were well dispersed over
208
the HZSM-5 support.
209
3.2.2 XPS measurements
210
To evaluate the surficial elemental species of MnxCe1-xO2/HZSM-5 and
211
Mn0.8Ce0.2O2 catalysts, XPS analysis was then conducted. As shown in Table 1, the
212
Mn0.8Ce0.2O2/HZSM-5 catalyst had a relatively higher Ce3+ mol% than the
213
Mn0.8Ce0.2O2. This implies that a degree of electron transfer from HZSM-5 to Ce4+
214
occurred. This electron transfer was also observed between the Mn3+ and Ce4+, which
215
led to a higher Mn4+ mol% in the Mn0.8Ce0.2O2/HZSM-5 catalyst (as compared to the
216
MnOx/HZSM-5). In O1s XPS spectra (see supplementary Fig. S4), the
217
Mn0.8Ce0.2O2/HZSM-5 catalyst showed the highest Osur mol% among all HZSM-5
218
supported catalysts. This is unsurprising given the large amounts of Ce3+ in the
219
catalyst that required abundant oxygen vacancies to compensate for charge neutrality
220
[34, 35]. As Osur is reported to be the most active oxygen species in oxidation
221
reactions [36, 37], the Mn0.8Ce0.2O2/HZSM-5 catalyst, with its high Osur mol%, was
222
unsurprising with a remarkable oxidative ability and yielded a low Ea for CB 9 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
Page 10 of 28
223
oxidation (see Fig. 1). The high molar ratio of Oads species (as resulting from surface
224
carbonate or hydroxyl species) observed in all HZSM-5 supported catalysts was
225
mainly due to the large amount of hydroxyl species at the HZSM-5 surface.
226
Table 1 Cumulative H2 consumption and XPS analysis of the MnxCe1-xO2/HZSM-5 and
227
Mn0.8Ce0.2O2 catalysts (The deconvolution spectra of O1s, Mn2p and Ce3d were present in
228
supplementary data).
229
Catalyst
H2 consumption ( mol/g)
MnOx/HZSM-5
Oxygen distribution (%)
Mn4+/Mn Ce3+/Ce (mol%) (mol%)
Olat
Osur
Oads
0.53
13.2
13.5
73.3
32.7
-
Mn0.8Ce0.2O2/HZSM-5
0.64
36.9
20.2
42.9
37.3
21.1
CeO2/HZSM-5
0.0065
19.5
15.4
65.2
-
13.4
Mn0.8Ce0.2O2
1.39
43.1
31.6
25.3
30.2
12.8
3.2.3 Pyridine-IR measurements
230
Ce3+ O
O O
B
e-
Mn4+
AlO
O SiO
Si-
O
H+ O
O
AlAlSiO O O O O O O SiSiSi-
Mn3+ Ce4+
231 232
Fig. 3 (A) Pyridine-IR spectra of the MnxCe1-xO2/HZSM-335, HZSM-5, and Mn0.8Ce0.2O2 10 / 28
ACS Paragon Plus Environment
Page 11 of 28
Environmental Science & Technology
233
catalysts; (B) Schematic diagram describing the Mn and Ce adsorption sites on HZSM-5.
234
To evaluate the ion exchange behaviors between MnxCe1-xO2 and HZSM-5
235
catalysts, pyridine-IR measurements were conducted. In general, bands at
236
approximately 1635 and 1544 cm-1 were related to pyridine adsorbed onto Brønsted
237
sites and bands at approximately 1612 and 1455 cm-1 were related to pyridine
238
adsorbed onto Lewis sites [38-41]. The band at approximately 1490 cm-1 originated
239
from the pyridine adsorbed onto both Lewis sites and Brønsted sites [41, 42] and the
240
band at approximately 1474 cm-1 was assigned to the C-H bending of aliphatic
241
hydrocarbons bound to the oxygen atoms in a zeolite framework [43, 44]. As shown
242
in Fig. 3 and Table 2, the MnOx/HZSM-5 had a high degree of ion exchange between
243
the Mnn+ and H· in the HZSM-5, where the Brønsted/Lewis (B/L) ratio was measured
244
at approximately 0.82, much lower than that of HZSM-5 alone (at approximately
245
5.07). The reason could be ascribed to the Mn3+ ions interacting with the protons of
246
acidic bridging hydroxyl (strong Brønsted acid sites, SiAlOH), forming Mn(OH)3+
247
species that were subsequently transferred into strong Lewis acid cores after
248
calcination [33, 45]. In comparison, the CeO2/HZSM-5 had a much higher B/L ratio
249
(at approximately 3.82) than the MnOx/HZSM-5. This implies that most of the Ce4+
250
ions would interact with the non-acidic terminal silanol (SiOH) groups of the
251
HZSM-5 [46, 47], whereas only a few of the Ce4+ ions interacted with the SiAlOH
252
groups, leading to a slight decrease in the Brønsted acidity. For Mn0.8Ce0.2O2/HZSM-5
253
catalyst, the B/L ratio was approximately 0.55, much lower than those for the
254
HZSM-5 (at approximately 5.07) and Mn0.8Ce0.2O2 (at approximately 1.92), indicating
255
that a high degree of ion exchange occurred in the combined catalyst.
256
In
particular,
as
compared
with
the
MnOx/HZSM-5
catalyst,
the
257
Mn0.8Ce0.2O2/HZSM-5 catalyst showed an enhanced Lewis acidity, as the
258
characteristic bands at 1612 and 1455 cm-1 both increased. This could be a result of
259
the labile Ce3+/Ce4+ cycle that induced electron transfers from Ce4+ to Mn3+ (as
260
verified by XPS analysis, see Table 1), hence improving the electron accepting ability
261
of neighboring MnOx molecules (see Fig. 2 for the H2-TPR profile). Since the Lewis
262
acid sites were reported to be the core active sites for C-C band cleavage [14, 48], the 11 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 28
263
Mn0.8Ce0.2O2/HZSM-5 with its high Lewis acidity was expected to yield high CO2
264
selectivity in the catalytic oxidation of CB. This assumption was consistent with
265
results presented in our previous report [21]. In sum, the pyridine-IR results suggest
266
that by tuning the molar ratio of Mn:Ce, we should be able to adjust the Brønsted and
267
Lewis acid sites of the MnxCe1-xO2/HZSM-5 catalysts for various catalytic
268
applications.
269
Table 2 Brønsted/Lewis (B/L) ratios of MnxCe1-xO2/HZSM-5, HZSM-5, and Mn0.8Ce0.2O2
270
catalysts Catalysts
Brønsted/Lewis (B/L) ratio
5.07 0.82 0.55 3.82 1.11
HZSM-5 MnOx/HZSM-5 Mn0.8Ce0.2O2/HZSM-5 CeO2/HZSM-5 Mn0.8Ce0.2O2
Mn0.8Ce0.2O2/HZSM-5
Fresh Water treated Used in dry condition Used in humid condition
0.55 0.47 1.07 0.50
271
Note: Quantitative method of B/L ratios [49]: C(pyridine on B sites= 1.88 IA(B) R2/W and
272
C(pyridine on L sites) = 1.42 IA(L) R2/W, Where C = concentration (mmol/g catalyst), IA(B, L) =
273
integrated absorbance of B or L band (cm-1), R = radius of catalyst disk (cm) , W = weight of disk
274
(mg).
275
3.2.4 DRIFT measurements
276 12 / 28
ACS Paragon Plus Environment
Page 13 of 28
Environmental Science & Technology
277 278
Fig. 4 DRIFT spectra taken at 200 °C upon passing 10 vol% O2/He over the CB pre-sorbed onto
279
Mn0.8Ce0.2O2 (A) and Mn0.8Ce0.2O2/HZSM-5 (B) catalysts at different times
280
To get insights into the CB adsorption and reaction behaviors over
281
Mn0.8Ce0.2O2/HZSM-5 and Mn0.8Ce0.2O2 catalysts, in situ DRIFT measurements were
282
conducted while flowing the CB at 200 °C for 15 min followed by CB with 10% vol
283
O2 for other 15 min.
284
Fig. 4A shows the CB adsorption onto the Mn0.8Ce0.2O2 catalyst as a function of
285
time. A series of peaks were identified at approximately 1248, 1320, 1378, 1440,
286
1460, 1540, 1560, 1620, 1700, 2350, 3378, 3659, 3740 cm-1. The peaks at 1440 and
287
1540 cm-1 corresponded to the stretching vibrations of C=C in an aromatic ring [50,
288
51]. The peak at 1700 was assigned to aldehyde-type species [21, 52], and that at
289
1248 cm-1 to the C-H vibration in a plane [13, 50]. The peaks at 1320, 1378, and 1590
290
cm-1 corresponded to COOH- from bidentate formate [50] and those at 1460 and 1560
291
cm-1 corresponded to the COOH- from acetate-type species [53, 54]. The peak at 2350
292
cm-1 originated from CO2 adsorption [54, 55]. With an increase in the purging time,
293
the peaks at 1440 and 1540 cm-1 (assigned to the C=C on an aromatic ring) quickly
294
decreased. This implies that the aromatic ring in CB was cleaved once in contact with
295
the Mn0.8Ce0.2O2. The increases in the peaks at 1248 cm-1 (assigned to the vibration of
296
C-H in the plane) 1460, and 1560 cm-1 (assigned to the COOH- of acetate-type species)
297
also support this conclusion. 13 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
298
After O2 was introduced, the intensities of the peaks at 1700 cm-1 (assigned to
299
aldehyde-type species) 1320, 1378, and 1590 cm-1 (assigned to COOH- of bidentate
300
formates) all increased. This implies that O2 was involved in the CB oxidation
301
reaction, which dissociated the COOH- from the bidentate formates. In the hydroxyl
302
region of 3000-4000 cm-1, the peak at 3740 cm-1 originated from the produced H2O
303
[56], while peaks at 3659 and 3378 cm-1 corresponded to the consumption of hydroxyl
304
at the metal oxide surfaces [14, 57].
305
Fig. 4B shows adsorption of the CB onto the Mn0.8Ce0.2O2/HZSM-5 catalyst as a
306
function of time. As with Mn0.8Ce0.2O2, a series of peaks were identified at
307
approximately 1248, 1320, 1378, 1440, and 1460 cm-1. However, an extra peak at
308
1640 cm-1 was observed for Mn0.8Ce0.2O2/HZSM-5, which corresponded to the
309
vibration of CB adsorption onto the Brønsted sites [38, 41]. This peak quickly
310
decreased, indicating that the Brønsted sites were consistently consumed by CB
311
adsorption. As illustrated in our previous work [21], the adsorption of CB onto
312
Brønsted sites could convert the CB into phenolates via a nucleophilic substitution
313
process. These phenolates were then quickly transformed into benzoquinone or
314
cyclohexanone and facilitated the aromatic ring cleavage and deep oxidation
315
processes. Moreover, because the peaks at 1440 and 1540 cm-1 (assigned to C=C on
316
an aromatic ring) and those in the range of 1250-1500 cm-1 (corresponding to
317
oxidation products from aromatic ring cleavage) did not change as obviously as they
318
did with Mn0.8Ce0.2O2, CB oxidation over the Mn0.8Ce0.2O2/HZSM-5 barely occurred
319
at this stage. After O2 was introduced, peaks characteristic of CB oxidative products
320
appeared, indicating that the O2 had induced the cleavage of the CB aromatic rings,
321
which facilitated the deep oxidation of CB over the Mn0.8Ce0.2O2/HZSM-5 catalyst. In
322
the hydroxyl region of 3000-4000 cm-1, as with Mn0.8Ce0.2O2, peaks at approximately
323
3659 and 3378 cm-1 were also identified. However, extra peaks were observed at
324
approximately 3440 and 3290 cm-1, which could be originated from the surface
325
hydroxyl (in HZSM-5) interacting with the CB molecules that led to the formation of
326
weakened hydrogen-bonded OH- [12, 58].
327
Combined with our previous publication [21], it can be concluded that the 14 / 28
ACS Paragon Plus Environment
Page 14 of 28
Page 15 of 28
Environmental Science & Technology
328
HZSM-5 was not only able to convert the CB into phenolates via the adsorption of
329
CB onto the Brønsted sites; this had facilitated the deep oxidation of CB; but also able
330
to adjust the Brønsted/Lewis (B/L) ratio via the ion exchange between Mnn+ and H·;
331
this had promoted the redox potential of Mn1-xCexO2, both of which led to a low Ea in
332
the catalytic oxidation of CB over the Mn0.8Ce0.2O2/HZSM-5 catalyst (see Fig. 1).
333
3.3 Stability measurements
334
As mentioned above, the stability of industrial catalysts in the catalytic oxidation
335
of CVOCs is still a technological problem in industry [59, 60, 61]. To assess the
336
practicality of the Mn0.8Ce0.2O2/HZSM-5 catalyst, a long-term ageing test was
337
conducted at T90 (i.e., 90% CB conversion) with a WHSV at 15,000 mL g-1h-1, in both
338
dry and humid conditions.
339
In dry conditions (see Fig. 5A), the Mn0.8Ce0.2O2/HZSM-5 and Mn0.8Ce0.2O2
340
catalysts both showed a stable performance in the first 10 h. However, their activities
341
began to drop as the experimental time approached 15 h. The Mn0.8Ce0.2O2/HZSM-5
342
catalyst retained approximately 60% CB conversion after 50 h of ageing, whereas the
343
Mn0.8Ce0.2O2 retained only approximately 25% CB conversion.
344
As indicated in our previous study [21], the activity of Mn0.8Ce0.2O2 is mainly
345
dependent on the redox potential of metal oxides where CB is dechlorinated to
346
benzene, which could then be oxidized by chemisorbed oxygen species to CO2 and
347
H2O [61]. Accordingly, deactivation of the Mn0.8Ce0.2O2 catalyst should be caused by
348
the accumulation of chloride species or coke at the metal oxide surfaces, which
349
inhibited their redox activities [61, 62]. For the Mn0.8Ce0.2O2/HZSM-5 catalyst,
350
oxidation of the CB was initiated by the adsorption of CB onto the Brønsted acid sites,
351
where they formed phenolates followed by benzoquinone or cyclohexanone species
352
[13, 14]. As such, the chloride species or coke might be in part trapped by the
353
HZSM-5 [60, 63], which to some extent protected the Mn0.8Ce0.2O2 active sites, and
354
yielded better stability than the Mn0.8Ce0.2O2 catalyst alone.
355
In a humid ageing condition (see Fig. 5B), the Mn0.8Ce0.2O2/HZSM-5 catalyst 15 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
356
had very stable performance within the 50 h experiment, whereas the Mn0.8Ce0.2O2
357
yielded even worse stability than in the dry ageing condition. The activity of the
358
catalyst quickly dropped after 10 h, and only approximately 10% CB conversion was
359
retained after 50 h of ageing. It was widely believed that the addition of
360
low-concentration H2O would induce a competitive adsorption with CB, hence
361
inhibiting catalytic activities [64, 65], whilst large amounts of H2O addition could
362
inhibit coke formation and promote Cl· removal from the catalyst surface [20, 66, 67].
363
To
364
Mn0.8Ce0.2O2/HZSM-5 and Mn0.8Ce0.2O2 catalysts, XPS, EDS-Mapping and
365
O2-TP-MS were then conducted.
assess the actual effects of H2O on the
stability performance of
366
367 368
Fig. 5 Stability measurements of the Mn0.8Ce0.2O2 and Mn0.8Ce0.2O2/HZSM-5 catalysts at their
369
respective T90 in (A) dry and (B) humid (RH = approximately 10 vol%) conditions.
370
371
3.3.1 XPS and pyridine-IR analyses The XPS analyses (see supplementary Fig. S5) indicated that dry-aged 16 / 28
ACS Paragon Plus Environment
Page 16 of 28
Page 17 of 28
Environmental Science & Technology
372
Mn0.8Ce0.2O2 did not show obvious changes in the O1s, Mn2p, and Ce3d spectra.
373
Only distinct Cl2p characteristic peaks were observed, indicating that the long-term
374
ageing had indeed caused Cl· accumulation at the surface. For humid ageing, the XPS
375
revealed a lack of Mn2p and Ce3d characteristic peaks. Only the Oads characteristic
376
peak originated from the introduced H2O was observed, indicating that the humid
377
ageing had led to formation of a H2O layer over the catalyst surface. The aged
378
Mn0.8Ce0.2O2 therefore suffered a more severe deactivation because the presence of a
379
H2O layer inhibited the adsorption of CB and O2 for reaction.
380
For Mn0.8Ce0.2O2/HZSM-5 catalyst, the XPS did not show obvious changes in
381
O1s, Mn2p, and Ce3d spectra (in either dry or humid ageing conditions). However, a
382
slight decrease in Olat and increase in Ce3+, together with the appearance of Cl2p
383
characteristic peaks were observed in the aged sample. After pyridine-IR analysis (see
384
Fig. 6 and Table 2), it was noted that the amounts of Brønsted acidity did not change
385
distinctly after dry and humid ageing. However, the Lewis acidities decreased
386
remarkably in dry ageing, but significantly increased in humid ageing. The former
387
could be ascribed to an accumulation of Cl· over the metal oxides that suppressed
388
their electron accepting abilities, while the latter was assumed to be caused by the
389
effective removal of Cl· by H2O washing [66, 67]. Indeed, EDS-mapping (see Fig. 7)
390
indicated that the dry-aged sample had similar distributions of Cl and Mn elements,
391
implying that the Cl· species mainly accumulated on the Mn0.8Ce0.2O2 sites. In
392
comparison, for the humid-aged sample, much less Cl· and C species were observed.
393
One might argue that the increased Lewis acidity might be also caused by H2O
394
interacting with the HZSM-5 catalyst. As Marie-Rose et al. [68] reported, H2O
395
adsorption onto protonic zeolite (HZSM-5 herein) would induce a proton transfer to
396
form H2O dimers, which bonded to the anionic zeolite framework to form H5O2+ ions,
397
leading to an increase in Lewis acidity. Indeed, after H2O treatment in which the fresh
398
Mn0.8Ce0.2O2/HZSM-5 was aged in humid condition for 50 h without the addition of
399
CB, the treated catalyst also yielded enhanced Lewis acidity (see Fig. 6). This result
400
suggests that an increase in the Lewis acidity of humid-aged Mn0.8Ce0.2O2/HZSM-5
401
catalyst was caused by both H2O washing and H5O2+ formation. Since the Lewis acid 17 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 28
402
sites have been reported to be the core active sites for C-C cleavage [14, 48], their
403
enhancement would reduce the accumulations of Cl· and coke on the catalyst surface,
404
hence enabling stable CB oxidation. This is all consistent with the stability
405
measurement results (see Fig. 5).
406
407 408
Fig. 6 Pyridine-IR spectra of the Mn0.8Ce0.2O2/HZSM-5 catalyst treated in dry or humid conditions
409 410
Fig. 7 EDS-mapping of Mn0.8Ce0.2O2/HZSM-5 catalysts aged in dry or humid conditions.
411
Considering that real-time application generally requires a pre-spraying process
412
to
remove
acidic
gases
(i.e.,
H2S)
and
water
413
dimethylformamide–DMF), an adjustment in humidity to assist the subsequent CVOC
414
oxidation process could be easily achieved. As such, the finding that a high-humidity
415
operating environment could enable stable performance for acidic solid-supported 18 / 28
ACS Paragon Plus Environment
soluble
species
(i.e.,
Page 19 of 28
Environmental Science & Technology
416
catalysts might provide a solution to resolve the deactivation problem in industrial
417
applications. However, the increase in HCl corrosion and a possible collapse of the
418
monolithic catalyst structure are still of concern.
419
3.3.2 Surficial accumulation compounds analyses
420
In industry, analysis of surficial accumulation compounds for used catalysts is
421
crucial for responsible waste disposal. To evaluate the surficial accumulation
422
compounds of aged catalysts, they were all subjected to thermal desorption treatments
423
where the desorbed gaseous compounds were simultaneously analyzed using an
424
GC-MS. Fig. 8A shows the main surficial accumulation compounds for aged
425
Mn0.8Ce0.2O2 catalysts. It was noted that the compounds desorbed from the dry-aged
426
sample were mainly benzene (label 4’) and CB (label 6’) (see Table 3), while for the
427
humid-aged sample, a distinct H2O peak (label 3’) was observed, implying that the
428
surface of the catalyst was covered by significant amounts of H2O. This was
429
consistent with the XPS results (see supplementary Fig. S5).
430
Fig. 8B shows the main surficial accumulation compounds for aged
431
Mn0.8Ce0.2O2/HZSM-5. There were approximately 14 organic compounds detected
432
(see Table 3). The dry-aged sample accumulated significant organic species,
433
including non-chlorinated (e.g. labels 1, 2, 3, 5, and 10) and chlorinated ones (e.g.,
434
labels 6, 7, 8, 9, 12, 13, and 14), while the humid-aged sample had significantly
435
decreased amounts of trichloromethane (label 6), tetrachloromethane (label 7),
436
trichloroethylene (label 8), tetrachloroethylene (label 9), o-Dichlorobenzene (label 12)
437
and p-Dichlorobenzene (label 14), some of which were even undetectable. This
438
indicates that the addition of H2O indeed inhibited Cl· accumulation via an increase in
439
the Lewis acidity and removed Cl· from the catalyst via washing. It should be noted
440
that although the presence of HZSM-5 has significantly promoted the catalytic
441
performance for CB oxidation, the increase in high toxic poly-chlorinated by-products
442
(e.g. label 7, 9, 12, etc.) is still a big concern, particularly for waste disposal of used
443
catalysts in industry. 19 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
Page 20 of 28
444
445 446
Fig. 8 GC-MS analysis for surficial accumulation compounds on (A) Mn0.8Ce0.2O2 and (B)
447
Mn0.8Ce0.2O2/HZSM-5 aged in dry or humid conditions
448
Table 3 Ingredients of surficial accumulation compounds on aged Mn0.8Ce0.2O2/HZSM-5 and
449
Mn0.8Ce0.2O2 Mn0.8Ce0.2O2 Label
Compound name
1’ 2’ 3’
Propylene Butylene Water
4’
Benzene
5’
Acetaldehyde
6’
Chlorobenzene
7’
p-Dichlorobenzene
Mn0.8Ce0.2O2/HZSM-5 Structure
Label
Compound name
H2O
1 2 3
Propylene Butane Butylene
4
Water
5
Pentene
6
Trichloromethane
CHCl3
7
tetrachloromethan e
CHCl4
o Cl Cl
Cl 20 / 28
ACS Paragon Plus Environment
Structure
H2O
Page 21 of 28
Environmental Science & Technology
8
Trichloroethylene
9
Tetrachloroethyle ne
10
Benzene
11
Chlorobenzene
12
o-Dichlorobenzen e
13
m-Dichlorobenzen e
14
p-Dichlorobenzen e
H
Cl
Cl
Cl
Cl
Cl
Cl
Cl
Cl Cl Cl Cl Cl
Cl
Cl
450
Evaluation on coke deposition of the aged catalyst was conducted using an
451
O2-TPD-MS. As shown in Fig. 9, the dry-aged sample had yielded much higher CO2
452
and CO desorption peaks than those for the humid-aged sample. This is unsurprising
453
given that humid aging induces a distinct increase in Lewis acidity, which promoted
454
the C-C band cleavage and deep oxidation processes.
455 456
Fig. 9 O2-TP-MS analysis of the Mn0.8Ce0.2O2/HZSM-5 catalysts aged in dry or humid conditions
21 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
457
Supporting Information
458 459 460
Arrhenius plots to eliminate the external and internal diffusion effects, XPS of deconvolution spectra, H2-TPR for CeO2, etc. are present in supplemental section. This material is available free of charge via the Internet at http://pubs.acs.org.
461
Acknowledgements
462
This work was financially supported by the National Natural Science Foundation of
463
China (Grant No. 51478418) and the Program for Zhejiang Leading Team of S&T
464
Innovation (Grant No. 2013TD07).
465
References
466
[1] R. Weber, T. Sakurai, H. Hagenmaier, Low temperature decomposition of PCDD/PCDF,
467
chlorobenzenes and PAHs by TiO2-based V2O5-WO3 catalysts, Appl. Catal. B, 20 (1999)
468
249-256.
469
[2] C. He, Y. Yu, Q. Shen, J. Chen, N. Qiao, Catalytic behavior and synergistic effect of
470
nanostructured mesoporous CuO-MnOx-CeO2 catalysts for chlorobenzene destruction,
471
Appl.Surf.Sci., 297 (2014) 59-69.
472
[3] C. He, Y. Yu, J. Shi, Q. Shen, J. Chen, H. Liu, Mesostructured Cu-Mn-Ce-O composites with
473
homogeneous bulk composition for chlorobenzene removal: Catalytic performance and
474
microactivation course, Macromol. Chem. Phys., 157 (2015) 87-100.
475
[4] H. Li, G. Lu, Q. Dai, Y. Wang, Y. Guo, Y. Guo, Efficient low-temperature catalytic combustion
476
of trichloroethylene over flower-like mesoporous Mn-doped CeO2 microspheres, Appl. Catal.
477
B, 102 (2011) 475-483.
478 479
[5] M. Guillemot, J. Mijoin, S. Mignard, P. Magnoux, Mode of zeolite catalysts deactivation during chlorinated VOCs oxidation, Appl.Catal.A, 327 (2007) 211-217.
480
[6] X. Ma, X. Feng, J. Guo, H. Cao, X. Suo, H. Sun, M. Zheng, Catalytic oxidation of 1,
481
2-dichlorobenzene over Ca-doped FeOx hollow microspheres, Appl. Catal. B, 147 (2014)
482
666-676.
483
[7] V. De Jong, M.K. Cieplik, R. Louw, Formation of dioxins in the catalytic combustion of 22 / 28
ACS Paragon Plus Environment
Page 22 of 28
Page 23 of 28
Environmental Science & Technology
484
chlorobenzene and a micropollutant-like mixture on Pt/γ-AL2O3, Environ.Sci.Tec., 38 (2004)
485
5217-5223.
486 487
[8] Y. Liu, M. Luo, Z. Wei, Q. Xin, P. Ying, C. Li, Catalytic oxidation of chlorobenzene on supported manganese oxide catalysts, Appl. Catal. B, 29 (2001) 61-67.
488
[9] Q. Dai, S. Bai, J. Wang, M. Li, X. Wang, G. Lu, The effect of TiO2 doping on catalytic
489
performances of Ru/CeO2 catalysts during catalytic combustion of chlorobenzene, Appl.
490
Catal. B, 142 (2013) 222-233.
491
[10] W. Cen, Y. Liu, Z. Wu, J. Liu, H. Wang, X. Weng, Cl Species Transformation on CeO2 (111)
492
Surface and Its Effects on CVOCs Catalytic Abatement: A First-Principles Investigation, J.
493
Phys. Chem. C , 118 (2014) 6758-6766.
494
[11] J. Wang, X. Wang, X. Liu, J. Zeng, Y. Guo, T. Zhu, Kinetics and mechanism study on
495
catalytic oxidation of chlorobenzene over V2O5/TiO2 catalysts, J. Mol. Catal. A-Chem., 402
496
(2015) 1-9.
497
[12] B.H. Aristizábal, C.M. de Correa, A.I. Serykh, C.E. Hetrick, M.D. Amiridis, In situ FT-IR
498
study of the adsorption and reaction of ortho-dichlorobenzene over Pd-promoted Co-HMOR,
499
Microporous Mesoporous Mater, 112 (2008) 432-440.
500 501
[13] J. Lichtenberger, M.D. Amiridis, Catalytic oxidation of chlorinated benzenes over V2O5/TiO2 catalysts, J.Catal., 223 (2004) 296-308.
502
[14] J. Wang, X. Wang, X. Liu, T. Zhu, Y. Guo, H. Qi, Catalytic oxidation of chlorinated benzenes
503
over V2O5/TiO2 catalysts: The effects of chlorine substituents, Catal. Today, 241 (2015)
504
92-99.
505 506 507 508
[15] S. Scirè, S. Minicò, C. Crisafulli, Pt catalysts supported on H-type zeolites for the catalytic combustion of chlorobenzene, Appl. Catal. B, 45 (2003) 117-125. [16] M. Taralunga, J. Mijoin, P. Magnoux, Catalytic destruction of 1, 2-dichlorobenzene over zeolites, Catal. Commun., 7 (2006) 115-121.
509
[17] Q. Dai, H. Huang, Y. Zhu, W. Deng, S. Bai, X. Wang, G. Lu, Catalysis oxidation of 1,
510
2-dichloroethane and ethyl acetate over ceria nanocrystals with well-defined crystal planes,
511
Appl. Catal. B, 117 (2012) 360-368.
512 513
[18] X. Wang, Q. Kang, D. Li, Low-temperature catalytic combustion of chlorobenzene over MnOx-CeO2 mixed oxide catalysts, Catal. Commun, 9 (2008) 2158-2162. 23 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
514
[19] J. Su, Y. Liu, W. Yao, Z. Wu, Catalytic Combustion of Dichloromethane over
515
HZSM-5-Supported Typical Transition Metal (Cr, Fe, and Cu) Oxide Catalysts: A Stability
516
Study, J. Phys. Chem. C , 120 (2016) 18046-18054.
517
[20] aQ. Dai, X. Wang, G. Lu, Low-temperature catalytic destruction of chlorinated VOCs over
518
cerium oxide, Catal. Commun., 8 (2007) 1645-1649; bS. Chatterjee, H.L. Greene, Y.J. Park,
519
Comparison of modified transition metal-exchanged zeolite catalysts for oxidation of
520
chlorinated hydrocarbons, J. Catal., 138 (1992) 179-194; cH. Greene, D. Prakash, K. Athota,
521
G. Atwood, C. Vogel, Energy efficient dual-function sorbent/catalyst media for chlorinated
522
VOC destruction, Catal. Today, 27 (1996) 289-296;
523
Adsorption and catalytic destruction of trichloroethylene in hydrophobic zeolites, Appl. Catal.
524
B, 14 (1997) 37-47.
d
P.S. Chintawar, H.L. Greene,
525
[21] P. Sun, X. Dai, W. Wang, X. Weng, Z. Wu, Mechanism Study on Catalytic Oxidation of
526
Chlorobenzene over MnxCe1-xO2/H-ZSM5 Catalysts under Dry and Humid Conditions, Appl.
527
Catal. B, (2016) 389-397.
528 529 530 531
[22] M. Alifanti, M. Florea, S. Somacescu, V. Parvulescu, Supported perovskites for total oxidation of toluene, Appl. Catal. B, 60 (2005) 33-39. [23] H. Huang, Q. Dai, X. Wang, Morphology effect of Ru/CeO2 catalysts for the catalytic combustion of chlorobenzene, Appl. Catal. B, 158-159 (2014) 96-105.
532
[24] Y. Liu, H. Dai, J. Deng, Y. Du, X. Li, Z. Zhao, Y. Wang, B. Gao, H. Yang, G. Guo, In situ
533
poly (methyl methacrylate)-templating generation and excellent catalytic performance of
534
MnOx/3DOM LaMnO3 for the combustion of toluene and methanol, Appl. Catal. B, 140
535
(2013) 493-505.
536 537 538 539 540 541 542 543
[25] M.M. Natile, G. Boccaletti, A. Glisenti, Properties and reactivity of nanostructured CeO2 powders: Comparison among two synthesis procedures, Chem. Mat., 17 (2005) 6272-6286. [26] F. Arena, T. Torre, C. Raimondo, A. Parmaliana, Structure and redox properties of bulk and supported manganese oxide catalysts, Phys. Chem. Chem. Phys, 3 (2001) 1911-1917. [27] J. Carnö, M. Ferrandon, E. Björnbom, S. Järås, Mixed manganese oxide/platinum catalysts for total oxidation of model gas from wood boilers, Appl.Catal.A, 155 (1997) 265-281. [28] E. Stobbe, B. De Boer, J. Geus, The reduction and oxidation behaviour of manganese oxides, Catal. Today, 47 (1999) 161-167. 24 / 28
ACS Paragon Plus Environment
Page 24 of 28
Page 25 of 28
Environmental Science & Technology
544 545
[29] W. Weimin, Y. Yongnian, Z. Jiayu, Selective reduction of nitrobenzene to nitrosobenzene over different kinds of trimanganese tetroxide catalysts, Appl.Catal.A, 133 (1995) 81-93.
546
[30] G. Liu, R. Yue, Y. Jia, Y. Ni, J. Yang, H. Liu, Z. Wang, X. Wu, Y. Chen, Catalytic oxidation of
547
benzene over Ce-Mn oxides synthesized by flame spray pyrolysis, Particuology, 11 (2013)
548
454-459.
549 550
[31] S.T. Hussain, A. Sayari, F.ç. Larachi, Enhancing the stability of Mn-Ce-O WETOX catalysts using potassium, Appl. Catal. B, 34 (2001) 1-9.
551
[32] N. Hadi, A. Niaei, S. Nabavi, M.N. Shirazi, R. Alizadeh, Effect of second metal on the
552
selectivity of Mn/H-ZSM-5 catalyst in methanol to propylene process, J. Ind. and Eng. Chem,
553
29 (2015) 52-62.
554
[33] D. Mao, W. Yang, J. Xia, B. Zhang, Q. Song, Q. Chen, Highly effective hybrid catalyst for
555
the direct synthesis of dimethyl ether from syngas with magnesium oxide-modified HZSM-5
556
as a dehydration component, J.Catal., 230 (2005) 140-149.
557 558
[34] R. Peng, X. Sun, S. Li, L. Chen, M. Fu, J. Wu, D. Ye, Shape effect of Pt/CeO2 catalysts on the catalytic oxidation of toluene, Chem. Eng. J., 306 (2016) 1234-1246.
559
[35] S. Xie, Y. Liu, J. Deng, X. Zhao, J. Yang, K. Zhang, Z. Han, H. Dai, Three-dimensionally
560
ordered macroporous CeO2-supported Pd@ Co nanoparticles: Highly active catalysts for
561
methane oxidation, J. Catal., 342 (2016) 17-26.
562 563
[36] H. Huang, Q. Dai, X. Wang, Morphology effect of Ru/CeO2 catalysts for the catalytic combustion of chlorobenzene, Appl. Catal. B, 158 (2014) 96-105.
564
[37] F. Arena, G. Trunfio, J. Negro, B. Fazio, L. Spadaro, Basic evidence of the molecular
565
dispersion of MnCeO x catalysts synthesized via a novel “redox-precipitation” route, Chem.
566
Mat, 19 (2007) 2269-2276.
567 568
[38] C. Emeis, Determination of integrated molar extinction coefficients for infrared absorption bands of pyridine adsorbed on solid acid catalysts, J.Catal., 141 (1993) 347-354.
569
[39] M.A. Makarova, K. Karim, J. Dwyer, Limitation in the application of pyridine for
570
quantitative studies of brönsted acidity in relatively aluminous zeolites, Microporou. Mater, 4
571
(1995) 243-246.
572 573
[40] E. Parry, An infrared study of pyridine adsorbed on acidic solids. Characterization of surface acidity, J.Catal., 2 (1963) 371-379. 25 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
574 575 576 577 578 579 580 581
[41] T. Yashima, N. Hara, Infrared study of cation-exchanged mordenites and Y faujasites adsorbed with ammonia and pyridine, J.Catal., 27 (1972) 329-333. [42] T. Barzetti, E. Selli, D. Moscotti, L. Forni, Pyridine and ammonia as probes for FT-IR analysis of solid acid catalysts, J. Chem. Soc., Faraday Trans, 92 (1996) 1401-1407. [43] J. Weitkamp, New directions in zeolite catalysis, Studies in Surface Science and Catalysis, 65 (1991) 21-46. [44] W. Farneth, R. Gorte, Methods for characterizing zeolite acidity, Chem. Rev., 95 (1995) 615-635.
582
[45] H.D. Setiabudi, A.A. Jalil, S. Triwahyono, N.H.N. Kamarudin, R.R. Mukti, IR study of
583
iridium bonded to perturbed silanol groups of Pt-HZSM5 for n-pentane isomerization,
584
Appl.Catal.A, 417 (2012) 190-199.
585
[46] J. Bi, M. Liu, C. Song, X. Wang, X. Guo, C2-C4 light olefins from bioethanol catalyzed by
586
Ce-modified nanocrystalline HZSM-5 zeolite catalysts, Appl. Catal. B, 107 (2011) 68-76.
587
[47] B. de Rivas, C. Sampedro, E. Ramos-Fernández, R. López-Fonseca, J. Gascon, M. Makkee, J.
588
Gutiérrez-Ortiz, Influence of the synthesis route on the catalytic oxidation of 1,
589
2-dichloroethane over CeO2/H-ZSM5 catalysts, Appl.Catal.A, 456 (2013) 96-104.
590
[48] B. de Rivas, R. López-Fonseca, J.R. González-Velasco, J.I. Gutiérrez-Ortiz, On the
591
mechanism of the catalytic destruction of 1, 2-dichloroethane over Ce/Zr mixed oxide
592
catalysts, J. Mol. Catal. A-Chem., 278 (2007) 181-188.
593
[49] C.A.Emeis, Determination of Integrated Molar Extinction Coefficients for Infrared
594
Absorption Bands of Pyridine Adsorbed on Solid Acid Catalysts, J.Catal.,141, no. 2 (1993)
595
347-354.
596
[50] X. Ma, Q. Sun, X. Feng, X. He, J. Guo, H. Sun, H. Cao, Catalytic oxidation of 1,
597
2-dichlorobenzene over CaCO3/α-Fe2O3 nanocomposite catalysts, Appl.Catal.A, 450 (2013)
598
143-151.
599
[51] M.A. Larrubia, G. Busca, An FT-IR study of the conversion of 2-chloropropane,
600
o-dichlorobenzene and dibenzofuran on V2O5-MoO3-TiO2 SCR-DeNOx catalysts, Appl. Catal.
601
B, 39 (2002) 343-352.
602 603
[52] M. Nagao, Y. Suda, Adsorption of benzene, toluene, and chlorobenzene on titanium dioxide, Langmuir, 5 (1989) 42-47. 26 / 28
ACS Paragon Plus Environment
Page 26 of 28
Page 27 of 28
Environmental Science & Technology
604
[53] S. Lomnicki, J. Lichtenberger, Z. Xu, M. Waters, J. Kosman, M.D. Amiridis, Catalytic
605
oxidation of 2, 4, 6-trichlorophenol over vanadia/titania-based catalysts, Appl. Catal. B, 46
606
(2003) 105-119.
607 608
[54] X. Ma, X. Suo, H. Cao, J. Guo, L. Lv, H. Sun, M. Zheng, Deep oxidation of 1, 2-dichlorobenzene over Ti-doped iron oxide, PCCP, 16 (2014) 12731-12740.
609
[55] X. Ma, H. Sun, H. He, M. Zheng, Competitive reaction during decomposition of
610
hexachlorobenzene over ultrafine Ca-Fe composite oxide catalyst, Catal.Lett., 119 (2007)
611
142-147.
612 613
[56] M. Kantcheva, A.S. Vakkasoglu, Cobalt supported on zirconia and sulfated zirconia: II. Reactivity of adsorbed NOx compounds toward methane, J.Catal., 223 (2004) 364-371.
614
[57] S.L. Alderman, G.R. Farquar, E.D. Poliakoff, B. Dellinger, An infrared and X-ray
615
spectroscopic study of the reactions of 2-chlorophenol, 1, 2-dichlorobenzene, and
616
chlorobenzene with model CuO/silica fly ash surfaces, Environ.Sci.Tec., 39 (2005)
617
7396-7401.
618
[58] D. Carmello, E. Finocchio, A. Marsella, B. Cremaschi, G. Leofanti, M. Padovan, G. Busca,
619
An FT-IR and Reactor Study of the Dehydrochlorination Activity of CuCl2/γ-Al2O3-Based
620
Oxychlorination Catalysts, J.Catal., 191 (2000) 354-363.
621 622 623
[59] Q. Dai, S. Bai, Z. Wang, X. Wang, G. Lu, Catalytic combustion of chlorobenzene over Ru-doped ceria catalysts, Appl. Catal. B, 126 (2012) 64-75. [60] A. Aranzabal, M. Romero-Sáez, U. Elizundia, J.R. González-Velasco, J.A. González-Marcos,
624
Deactivation of H-zeolites during catalytic oxidation of trichloroethylene, J.Catal., 296 (2012)
625
165-174.
626 627 628 629 630
[61] X. Wang, L. Ran, Y. Dai, Y. Lu, Q. Dai, Removal of Cl adsorbed on Mn-Ce-La solid solution catalysts during CVOC combustion, J. Colloid Interface Sci, 426 (2014) 324-332. [62] W. Xingyi, K. Qian, L. Dao, Catalytic combustion of chlorobenzene over MnOx-CeO2 mixed oxide catalysts, Appl. Catal. B, 86 (2009) 166-175. [63] A. Aranzabal, J. González-Marcos, M. Romero-Sáez, J. González-Velasco, M. Guillemot, P.
631
Magnoux, Stability of protonic zeolites
632
2-dichloroethane), Appl. Catal. B, 88 (2009) 533-541.
633
in the catalytic oxidation of chlorinated VOCs (1,
[64] F. Bertinchamps, A. Attianese, M. Mestdagh, E.M. Gaigneaux, Catalysts for chlorinated 27 / 28
ACS Paragon Plus Environment
Environmental Science & Technology
634
VOCs abatement: Multiple effects of water on the activity of VOx based catalysts for the
635
combustion of chlorobenzene, Catal. Today, 112 (2006) 165-168.
636 637 638 639
[65] C.E. Hetrick, F. Patcas, M.D. Amiridis, Effect of water on the oxidation of dichlorobenzene over V2O5/TiO2 catalysts, Appl. Catal. B, 101 (2011) 622-628. [66] Q. Dai, S. Bai, X. Wang, G. Lu, Catalytic combustion of chlorobenzene over Ru-doped ceria catalysts: Mechanism study, Appl. Catal. B, 129 (2013) 580-588.
640
[67] G. Sinquin, C. Petit, S. Libs, J. Hindermann, A. Kiennemann, Catalytic destruction of
641
chlorinated C2 compounds on a LaMnO3+δ perovskite catalyst, Appl. Catal. B, 32 (2001)
642
37-47.
643
[68] S. Marie-Rose, T. Belin, J. Mijoin, E. Fiani, M. Taralunga, F. Nicol, X. Chaucherie, P.
644
Magnoux, Catalytic combustion of polycyclic aromatic hydrocarbons (PAHs) over zeolite
645
type catalysts: Effect of water and PAHs concentration, Appl. Catal. B, 90 (2009) 489-496.
28 / 28
ACS Paragon Plus Environment
Page 28 of 28