Subscriber access provided by UNIV OF LOUISIANA
Remediation and Control Technologies
Phosphate-Functionalized CeO2 Nanosheets for Efficient Catalytic Oxidation of Dichloromethane Qiguang Dai, Zhiyong Zhang, Jiaorong Yan, Jinyan Wu, Grayson Johnson, Wei Sun, Xingyi Wang, Sen Zhang, and Wangcheng Zhan Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b05002 • Publication Date (Web): 29 Oct 2018 Downloaded from http://pubs.acs.org on November 1, 2018
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 26
Environmental Science & Technology
1
Phosphate-Functionalized CeO2 Nanosheets for
2
Efficient Catalytic Oxidation of Dichloromethane
3
Qiguang Dai,†, # Zhiyong Zhang, ‡, # Jiaorong Yan,† Jinyan Wu,† Grayson Johnson,‡ Wei Sun,†
4
Xingyi Wang,† Sen Zhang,*, ‡ and Wangcheng Zhan*, †
5
†
Key Laboratory for Advanced Materials, Research Institute of Industrial Catalysis, School of
6
Chemistry & Molecular Engineering, East China University of Science & Technology, Shanghai
7
200237, PR China.
8
‡
Department of Chemistry, University of Virginia, Charlottesville, Virginia 22904, United
9
10
States.
ABSTRACT
11
Tuning acid and basic sites nature and profile in the surface of redox-active metal oxide
12
nanostructures is a promising approach to constructing efficient catalyst for the oxidative
13
removal of chlorinated volatile organic compounds (CVOCs). Herein, using the dichloromethane
14
(DCM) oxidation as a model reaction, we report that phosphate (POx) Brønsted acid sites can be
15
incorporated onto CeO2 nanosheets (NSs) surface via an organophosphate-mediated route, which
16
can effectively enhance the CeO2’s catalytic performance by promoting the removal of chlorine
17
poisoning species. Based on the systematic study of the correlation between POx composition,
18
surface structure (acid and basic sites) and catalytic properties, we find that the incorporated
19
Brønsted acid sites can also function to decrease the amount of medium-strong basic sites (O2-),
ACS Paragon Plus Environment
1
Environmental Science & Technology
Page 2 of 26
20
reducing the formation of chlorinated organic by-product monochloromethane (MCM) and
21
leading to the desirable product HCl. At the optimized P/Ce ratio condition (0.2), the POx-CeO2
22
NSs can perform a stable DCM conversion of 65-70% for over 10 hours at 250°C, and over 95%
23
conversion at 300°C, superior to both pristine and other phosphate-modified CeO2 NSs. Our
24
work clearly identifies the critical role of acid and basic sites over functionalized CeO2 for
25
efficient catalytic CVOCs oxidation, guiding the future advanced catalyst design for
26
environmental remediation.
27
INTRODUCTION
28
Most of chlorinated volatile organic compounds (CVOCs), such as dichloromethane (DCM),
29
1, 2-dichloroethane (DCE), vinyl chloride (VC), trichloroethylene (TCE) and chlorobenzene
30
(CB), are deemed as hazardous pollutants that should be stringently monitored and addressed to
31
minimize their negative impacts on environment and public health.1-4 Catalytic oxidation (also
32
called combustion) is one of the most promising technologies for the removal of CVOCs by
33
converting CVOCs to desirable product of HCl, H2O and CO2. In search of an efficient catalyst
34
for this oxidative conversion process, various catalytic materials have been studied, such as solid
35
acids,5-7 supported noble metals,8-10 and transition metal oxides.11-16 CeO2 is a rare-earth metal
36
oxide with an easy conversion between CeO2 and CeO2-x.
37
property allows CeO2 to facilely activate O2 and serve as an active O-carrier for the catalytic
38
oxidation of many molecules including CVOCs. Its high activity, plus high earth abundance and
39
low cost, makes CeO2 an excellent catalyst choice for CVOCs oxidation in mild conditions.11
40
However, a discouraging issue associate with CeO2-catalyzed CVOCs oxidation is that inorganic
41
chlorine species (Cl-) tend to over-strongly adsorbed on the active sites, leading to a rapid decay
42
of CeO2 catalytic performance. Doping CeO2 with first-row transition metal cations (e.g. Mn, Fe,
17-23
Such a unique Ce3+/Ce4+ redox
ACS Paragon Plus Environment
2
Page 3 of 26
Environmental Science & Technology
43
or Co)14-16 or loading of Ru24 can alleviate this poisoning issue, as these dopants can function to
44
promote the Deacon reaction (oxidation of Cl- into Cl2). But disappointingly, the organic-
45
chlorinated by-products will be generated in the presence of highly reactive Cl2, especially at
46
higher temperatures. It is of particular importance to develop a new strategy to enhance CeO2
47
based catalysts’ ability to remove the poisoning species while maintaining their high activity and
48
selectivity for optimized CVOCs oxidation.
49
Our previous experimental and theoretical studies reveal that Brønsted acid sites can shuttle
50
the protons to facilitate the formation of HCl which are subject to desorption to eliminate Cl-
51
poisoning species.25-26 Such a stability enhancement induced by Brønsted acid surface
52
functionalization have been observed in our previously reported vanadium oxide-modified CeO2
53
catalyst during DCE oxidation.25 But vanadium oxide is highly toxic, prohibiting its practical
54
application for environmental catalysis. As a more promising alternative, we envision that
55
grafting non-toxic Brønsted acid phosphate species in the surface of CeO2 can also deliver the
56
enhanced tolerance to chloride poisoning. Although phosphate-modification has been well-
57
established in other oxides studies such as MCM-41 and SBA-15 by reaction with inorganic
58
phosphoric acid (H3PO4),27-29 the direct functionalization of CeO2 with H3PO4 is prone to
59
generating CePO4 phase in catalyst surface that inhibits the catalyst’s redox property.30-33
60
Recently, Slowing et. al. reported that Brønsted acidic sites could be generated by
61
immobilization of organophosphates on CeO2 without lowering redox activity.34 In the present
62
work, we find that, using the reaction of CeO2 nanosheets (NSs) with trimethylphosphate (TMP)
63
and a subsequent calcination, phosphate (POx)-modified CeO2 NSs can be successfully prepared
64
with controllable surface POx content and without impurity phases. Since such surface Brønsted
65
acid sites are constructed without sacrificing the redox property of CeO2, our POx-modified CeO2
ACS Paragon Plus Environment
3
Environmental Science & Technology
Page 4 of 26
66
NSs exhibit high activity and durability for the catalytic oxidation of CVOCs, using DCM as a
67
model pollutant. More important, the surface acidity/basicity and redox property of POx-modified
68
CeO2 are systematically investigated by employing a suite of probe techniques, including NH3-
69
/CO2-temperature programmed desorption (TPD), H2-temperature programmed reduction (TPR)
70
and pyridine (Py)-mediated diffuse reflectance infrared Fourier transform spectroscopy
71
(DRIFTs). By carefully studying POx content-dependant catalytic properties, we find POx-CeO2
72
NSs with a P/Ce ratio of 0.2 is an excellent catalyst for DCM oxidation with balanced activity
73
and durability, superior to the pristine CeO2, POx-CeO2 NSs with other compositions and CeO2
74
modified with inorganic phosphates (phosphoric acid, ammonium phosphate and sodium
75
phosphate). More importantly, we refine the essential roles of Brønsted acid sites (P-OH, in
76
removing poisoning species) and basic site (surface lattice O2-, in tuning the formation of
77
chlorinated organic by-product) for optimized DCM oxidation performance, which can be
78
generalized to other CVOCs catalytic oxidation and beyond, providing a guidance to designing
79
and engineering catalyst with functional surface sites for advanced environmental catalysis.
80
EXPERIMENTAL SECTION
81
Chemicals and materials. Cerium (III) nitrate hexahydrate (Ce(NO3)3•6H2O), ammonium
82
bicarbonate (NH4HCO3), TMP, and sodium phosphate (SP) were purchased from Aladdin
83
Industrial Corporation, China. Phosphoric acid (PA) and ammonium phosphate (AP) were
84
obtained from Sinopharm Chemical Reagent Co., Ltd, China. Silica (SiO2, amorphous fumed)
85
was purchased from Alfa Aesar. All the chemicals were used as-received without any further
86
purification.
87
Synthesis of ceria nanosheets. In a typical synthesis,35 2.78 g of cerium (III) nitrate
88
hexahydrate (Ce(NO3)3•6H2O) and 1.50 g of ammonium bicarbonate (NH4HCO3) were dissolved
ACS Paragon Plus Environment
4
Page 5 of 26
Environmental Science & Technology
89
separately, each in 100 ml of deionized water, at 30°C under magnetic stirring. These two
90
solutions were then mixed together rapidly, stirred for 0.5 h, and statically aged for 24 h at 30°C.
91
The final product was collected by filtration, washed with deionized water to remove any
92
possible ionic remnants, dried at 80°C for overnight, and calcined at 500°C for 4 h in flowing air
93
to obtain the belt-like CeO2 nanosheets.
94
Synthesis of phosphate-modified ceria. Phosphate modified CeO2 nanosheets was
95
synthesized through an incipient wetness impregnation route. Typically, a desired amount (the
96
molar ratio of P/Ce was controlled to be 0.05, 0.1, 0.2 or 0.3) of TMP, PA, AP or SP was
97
dissolved in 1.1 ml of deionized water. 0.86 g of the CeO2 nanosheets was then added to the
98
phosphoric solution, statically aged for 24 h at room temperature, dried at 80°C, and calcinated at
99
450°C for 4 h with a 5°C/min ramp rate in flowing air. Catalysts prepared with TMP, PA, AP,
100
SP were named as POx-CeO2-n, PA-CeO2-n, AP-CeO2-n, and SP-CeO2-n respectively, where n
101
is the molar ratio of P/Ce. The same approach was applied to prepare POx-functionalized SiO2.
102
Catalyst characterization. Scanning electron microscopy (SEM) images were collected on a
103
Hitachi S-3400N at 8.0 kV. Transmission electron microscopy (TEM) images and mapping
104
analyses were acquired from a JEM-2100 operated at 200 kV. X-ray powder diffraction (XRD)
105
patterns were recorded on a Rigaku D/MAX 2550 VB/PC X-ray diffractometer with Cu Kα
106
radiation (λ= 0.154056 nm). The surface areas of catalysts were measured on a Micrometrics
107
ASAP 2460 adsorption apparatus at -196°C and calculated by the Brunauer-Emmett-Teller
108
(BET) method. FT-IR spectra were obtained in transmission mode on a Nicolet 6700
109
spectrometer with pressed KBr pellets.
110
CO2-, and NH3-TPD were performed in a quartz tube reactor system equipped with a
111
quadrupole mass spectrometer (MS, Hiden, HPR20, for CO2-TPD) or a thermal conductivity
ACS Paragon Plus Environment
5
Environmental Science & Technology
Page 6 of 26
112
detector (TCD, for NH3-TPD). Prior to the TPD analysis, 100 mg of the sample was first pre-
113
heated at 400°C in 20 vol. % O2/Ar flow (30 ml min-1) for 1 h, cooled to 100°C, and swept by Ar
114
for another 1 h. Pure CO2 or NH3 was introduced into the catalyst bed at 10 ml min-1 for 30 min.
115
An Ar flow (30 ml min-1) was then purged through the catalysts for 1 h to remove physiosorbed
116
CO2 or NH3. After that, the sample was heated from 100 to 600°C, at a ramp of 10°C min-1. For
117
CO2 detection, mass signals at m/z = 44 was recorded. H2-TPR was carried out using the same
118
apparatus. The sample (100 mg) was pre-treated following a similar procedure as the TPD
119
experiments, and was heated from 50 to 750°C at 10°C min-1 in a 5 vol.% H2/Ar flow (30
120
ml/min). The consumption of H2 was determined by a TCD.
121
FT-IR spectrum of the pyridine adsorption was measured on a Nicolet 6700 spectrometer
122
equipped with an MCT detector in diffuse reflection mode (DRIFTS). The sample was first
123
purged with O2/Ar (20 vol. %) at 400 ºC for 1 h, and background measurements were collected at
124
300, 200, and 100°C, respectively. The pyridine vapor was then admitted to the IR cell for 15
125
min at 100°C. After sweeping with Ar for 30 min, the spectra of adsorbed pyridine were
126
collected at the desired desorption temperature (100, 200, or 300°C).
127
Catalytic oxidation of dichloromethane. Catalytic oxidation of DCM was carried out in a
128
temperature-controlled flow reactor (U-shaped quartz tube with an inner diameter of 3 mm)
129
containing 200 mg of the as-prepared catalyst. The flow rate of feeding gas (air containing 500
130
ppm of DCM) was controlled by mass flow controllers (Alicat), and the space velocity was
131
always fixed to 15,000 ml/g•h. Here, air (dried by silica gel and 5A zeolite) and dichloromethane
132
(Macklin, 99.9%, with molecular sieves) were used to avoid the presence of water in the feeding
133
gas. The effluent gases were analysed on-line using a GC equipped with an FID detector, and the
134
conversion of DCM was calculated based on the peak area after running for 15 min at the setting
ACS Paragon Plus Environment
6
Page 7 of 26
Environmental Science & Technology
135
temperature. The amount of MCM by-product was calculated from the DCM peak using a
136
relative correction factor of 0.92.
137
RESULTS AND DISCUSSION
138
CeO2 NSs were synthesized through an aqueous phase precipitation reaction of Ce(NO3)3 and
139
NH4HCO3 at room temperature followed by a calcination in air at 500°C, according to our
140
previously reported method.35 Subsequent incipient wetness impregnation of TMP allows the
141
generation of POx-functionalized CeO2 nanosheets, as modified from Slowing’s route
142
(Experimental Section).34 SEM and TEM images in Figure 1a & S1b indicate that as-
143
synthesized CeO2 samples have a belt-like nanosheet morphology with a thickness of 20-50 nm,
144
a length of 5-10 µm and a width of 0.5-1.2 µm. The NS structure is well retained after the
145
incorporation of functional POx species in POx-CeO2-0.05 (Figure S1a and S1c) and POx-CeO2-
146
0.2 (Figure 1b and 1c) samples. Meanwhile, these pristine and POx-modified CeO2 NSs are
147
polycrystalline, which are confirmed by the concentric circles of the selected area electron
148
diffraction (SAED) patterns (insets of Figure 1c, S1b and S1c).36. Such morphological and
149
crystallographic similarities between pristine and POx-modified CeO2 NSs allow us to
150
unambiguously refine POx species effect on sample’s catalytic properties.
151
SEM-EDS elemental mappings of POx-CeO2-0.05 and POx-CeO2-0.2 NSs demonstrate that P
152
is uniformly distributed over our samples regardless of the P concentration (Figure 1d and S1d).
153
Based on the SEM-EDS analysis, the element atomic ratios of P/Ce on POx-CeO2-0.05, POx-
154
CeO2-0.1, POx-CeO2-0.2 and POx-CeO2-0.3 are quantified to be 0.04, 0.09, 0.15 and 0.28,
155
respectively, only slightly lower than the precursor values. N2-physisorption measurement
156
indicates that the POx-modification leads to the decline of the specific surface areas. The BET
157
specific surface areas (SBET) are 75, 71, 65 and 63 m2/g for POx-CeO2-0.05, -0.1, -0.2 and -0.3,
ACS Paragon Plus Environment
7
Environmental Science & Technology
Page 8 of 26
158
respectively, lower than that of pristine CeO2 (85 cm2/g). The decrease is related with the
159
formation of phosphate34, which is bonded on the surface of CeO2, resulting in the monotonously
160
decrease in specific surface area with the P contents.
161 162
Figure 1. (a) SEM images of CeO2 NSs; (b-d) SEM (b) and TEM images with the
163
corresponding SAED pattern (c), and SEM-EDS elemental mapping (d) of POx-CeO2-0.2 NSs.
164
All POx-CeO2 NSs with different POx contents show the typical cubic fluorite structure
165
identical to CeO2 (JCPDS no. 34-0394),37 with no impurity diffraction peaks observed in their
166
XRD patterns (Figure 2a). This indicates that our TMP modification process does not lead to the
167
formation of crystalline phosphate impurity (such as CePO4 or CeP2O7) which are usually
168
observed in the traditional approach via inorganic phosphoric acid treatment.30 FT-IR was
ACS Paragon Plus Environment
8
Page 9 of 26
Environmental Science & Technology
169
conducted to confirm the successful POx-grafting on CeO2 surface. As seen in Figure 2b, all
170
POx-modified CeO2 present three new IR bands at 1158 cm-1, 1000 cm-1 and 950 cm-1 compared
171
to the pristine CeO2, and the intensities of these peaks increase with POx content. These peaks
172
are attributed to the symmetric stretching of non-bridging (O-P-O)sym oxygen atoms in Q2
173
phosphate units, the stretching vibrations of P-O- in Q0 species and the antisymmetric stretching
174
mode of non-bridging (P-O-P)antisym oxygen in isolated orthophosphate groups, respectively.38
175
The band at 1625 cm-1 is assigned to the bending motions of physiosorbed water molecules
176
and/or (P-OH) groups,38 while the latter contribution is more dominant as the peak is absent on
177
the pristine CeO2. More importantly, we find that the band at 1324 cm-1, which can be assigned
178
to carbonate species adsorbed on CeO2 basic sites,39 diminishes with increasing POx content and
179
eventually disappears on the POx-CeO2-0.2 sample, suggesting that the basic surface sites are
180
fully consumed by POx-modification in high POx content samples.
181 182
Figure 2. (a) XRD patterns and (b) FT-IR spectra of the pristine CeO2 and POx-CeO2 NSs with
183
different POx contents.
184
The nature of surface O and P over POx-CeO2 NSs was probed via XPS, as shown in Figure 3.
185
The O1s XPS spectra for both POx-CeO2-0.05 and POx-CeO2-0.2 can be deconvoluted into three
ACS Paragon Plus Environment
9
Environmental Science & Technology
Page 10 of 26
186
peaks: O2- (529 eV, lattice oxygen species, denoted as OⅠ), O22-/O- (530.8 eV, surface active
187
oxygen species, denoted as OⅡ) and -OH (532.5 eV, hydroxyl species and adsorbed water,
188
denoted as OⅢ).
189
functionalized CeO2 NSs can retain their redox properties. Moreover, the relative content of
190
oxygen species (OⅠ, OⅡ, OⅢ/Ototal) calculated from XPS measurement is 67.2, 22.1 and 10.8%,
191
61.8, 21.0 and 17.2% for POx-CeO2-0.05 and POx-CeO2-0.2, respectively, which indicates that
192
O22-/O- species decreases (redox sites) with the increasing of POx content while -OH species
193
increases (Brønsted acid sites). The P2p XPS spectrum for POx-CeO2-0.2 is deconvoluted into
194
two peaks. While the one at 132.5 eV is assigned to the monodentate surface complex, the one at
195
a higher binding energy (134.2 eV) is assigned to the bidentate surface complex, in which POx is
196
bound to two surface hydroxyl groups. Such a co-existence of surface mono- and bidentate POx
197
has also been reported in other POx-modified metal oxide.41 Surprisingly, however, POx-CeO2-
198
0.05 only shows a single peak of monodentate phosphate complex, indicating the bidentate is not
199
favourable under low concentration surface modification.
40
The presence of surface O22-/O- on the two samples suggests that our POx-
200 201
Figure 3. O1s and P2p XPS spectra of the (a) POx-CeO2-0.05 and (b) POx-CeO2-0.2 NSs.
ACS Paragon Plus Environment
10
Page 11 of 26
Environmental Science & Technology
202
The redox properties of POx-CeO2 NSs were further evaluated by H2-TPR. As shown in
203
Figure 4a, the pristine CeO2 and POx-CeO2-0.05 present a broad peak between 325 and 600°C in
204
their H2-TPR profiles, which are attributed to the reduction of surface oxygen species (adsorbed
205
oxygen species over O-vacancy and surface lattice oxygen species of CeO2).42 This peak
206
becomes more symmetric when incorporating POx, along with the disappearance of adsorbed
207
oxygen species due to POx coverage on CeO2 surface. Meanwhile, a small reduction peak at
208
lower temperatures (200 to 300°C) is detected on POx-CeO2-0.05 and POx-CeO2-0.1, which can
209
be ascribed to oxygen species adsorbed on the interface of POx and CeO2, as illustrated in
210
Scheme S1. Interestingly, we note this low-temperature peak becomes negligible on the POx-
211
CeO2-0.2 and POx-CeO2-0.3, since the high coverage of POx over CeO2 surface can effectively
212
mitigate the interface-mediated oxygen adsorption. Such high surface coverage of POx when
213
P/Ce>0.2 is consistent with aforementioned FT-IR analysis wherein surface carbonate species
214
absorption is also prohibited.
ACS Paragon Plus Environment
11
Environmental Science & Technology
Page 12 of 26
215 216
Figure 4. (a) H2-TPR, (b) NH3-TPD, and (c) CO2-TPD profiles of POx-CeO2 NSs with different
217
POx contents; (d) Schematic illustration of surface acidic and basic sites on POx-CeO2 NSs.
218
Surface acidity and basicity of POx-CeO2 NSs were systematically investigated by using NH3-
219
TPD and CO2-TPD. The NH3-TPD profile (Figure 4b) on pristine CeO2 NSs displays a peak in
220
the range of 125 to 250°C, which is resulted from the desorption of NH3 adsorbed on Ce4+/3+
221
(dominant) and surface acidic hydroxyl group (bridged OHad, as illustrated in Figure 4d), in
222
agreement with the mild acidic nature of CeO2.34 The POx modification significantly enhances
223
the acidic strength of CeO2 NSs, and correspondingly, NH3 desorption peak is broadened to
224
450°C, confirming the generation of the new medium-strong acid sites (P-OH).34 However, the
225
overall number of surface acid sites on POx-CeO2 NSs is reduced with the increase of POx
226
content. Based on integrated peak area, the total acid sites decrease drastically from 7.5x for
ACS Paragon Plus Environment
12
Page 13 of 26
Environmental Science & Technology
227
POx-CeO2-0.05 (relative to pristine CeO2) to 6.0x for POx-CeO2-0.1, 5.7x for POx-CeO2-0.2, and
228
4.7x for POx-CeO2-0.3. This is probably due to the transformation from monodentate to bidentate
229
POx when more species are grafted, as implied by XPS and IR results.
230
The CO2-TPD results in Figure 4c clearly exhibit two CO2 desorption peaks on pristine CeO2.
231
The one at 175°C is correlated to surface basic hydroxyl group (weak basic sites, as illustrated in
232
Figure 4d), while another one around 375°C reflects basic oxygen ion (O2-, medium-strong basic
233
site).
234
strong basic sites on pristine CeO2. Once POx is grafted, the density of weak basic sites is
235
substantially reduced and becomes even negligible on POx-CeO2-0.2 and POx-CeO2-0.3 NSs.
236
This can also be explained by the aforementioned mono-to-bidentate POx transition when
237
increasing POx content.34,
238
converted to bidentate -HPO4 through the dehydration reaction between acidic P-OH and the
239
adjacent surface basic hydroxyl groups over CeO2,27, 41 consuming these weak basic sites. More
240
interestingly, the CO2 desorption peak corresponding to the medium-strong basic sites O2- is
241
heightened and broadened after POx modification, especially for the POx-CeO2-0.05 NSs, which
242
is an indication that more of the medium-strong basic sites are formed. The lower
243
electronegativity of P (2.19) than O (3.44) and their interfacial effect are probably responsible for
244
the basicity change over POx-CeO2-0.05 NSs.
43
Moreover, the total number of weak basic sites is clearly much higher than medium-
41, 44-45
As illustrated in Scheme S2, monodentate -H2PO4 can be
ACS Paragon Plus Environment
13
Environmental Science & Technology
Page 14 of 26
245 246
Figure 5. DRIFT spectra of pyridine adsorption on (a) Al2O3, pristine CeO2, and POx-CeO2 NSs
247
with different POx contents at 200°C and (b) POx-CeO2-0.2 NSs at different temperatures.
248
Furthermore, the nature of acid sites (Brønsted or Lewis acid sites) and their corresponding
249
strengths were revealed in detail via pyridine (Py) adsorption coupled with IR spectroscopy.
250
Figure 5a shows the DRIFT spectra for Py adsorption on POx-CeO2 NSs with different POx
251
contents at 200°C, with Al2O3 as a reference. For pristine CeO2 and POx-CeO2, the presence of
252
Lewis acid sites is proven by two bands at approximately 1595 cm-1 (ν8a vibration mode of
253
adsorbed Py) and 1440 cm-1 (ν19b mode), both shifting to lower frequencies compared with
254
Al2O3 standard (1620 cm-1 and 1450 cm-1), suggesting the relatively weak strength of Lewis acid
255
sites on both pristine and POx-modified CeO2.46 The Brønsted acid site is undetectable on
256
pristine CeO2, but becomes evident on POx-CeO2 NSs due to the existence of P-OH as
257
demonstrated by the peak at 1525 cm-1.34 By comparing the peak area, we notice that the number
258
of Lewis acid site is significantly increased after the POx-grafting, with POx-CeO2-0.05 being the
259
highest one. Meanwhile, POx-CeO2-0.05 and POx-CeO2-0.3 NSs present fewer Brønsted acid
260
sites than the POx-CeO2-0.1 and POx-CeO2-0.2, probably due to the low POx content for POx-
261
CeO2-0.05 and the formation of bidentate POx (less P-OH) for POx-CeO2-0.3. In addition, these
ACS Paragon Plus Environment
14
Page 15 of 26
Environmental Science & Technology
262
acid sites are thermally stable. Temperature-dependant DRIFT spectra of POx-CeO2 (Figure 5b
263
and S2) show that the strong Py adsorption on the acid sites can still be observed even at 300°C,
264
which is consistent with our NH3-TPD results (the higher and broader NH3 desorption
265
temperature on POx-CeO2).
266 267
Figure 6. Catalytic oxidation of DCM on different catalysts: (a) effect of POx content, (b)
268
stability of catalysts, (c) effect of phosphate precursors, and (d) MCM selectivity on pristine
269
CeO2 and POx-CeO2 NSs with different POx contents.
270
With acid and basic sites over CeO2 being fine-tuned through POx-incorporation, we have
271
systematically studied its effect on catalytic oxidation of DCM. As shown in the light-off curves
272
in Figure 6a, the pristine CeO2 NSs exhibit an irregular increase of DCM conversion with two
ACS Paragon Plus Environment
15
Environmental Science & Technology
Page 16 of 26
11
273
plateaus, which is caused by the deactivation/poisoning of catalyst during the reaction.
The
274
complete oxidation of DCM is not achieved even at 400°C. The POx-modification can effectively
275
eliminate these plateaus and clearly improve the conversion of DCM when temperature is above
276
175°C. Among catalysts with different POx content, POx-CeO2-0.05 shows the maximum
277
increase rate in DCM conversion, while POx-CeO2-0.2 and POx-CeO2-0.3 present almost
278
overlapping curves. We believe that such a performance enhancement is attributed to the
279
inhibition of catalyst deactivation, given the facts that the pristine CeO2 shows a better initial
280
activity than POx-CeO2 while POx itself is not catalytically active towards DCM oxidation (only
281
20% DCM conversion was obtained over POx-SiO2-0.2 catalyst even at 400 °C, Figure S3).28-29
282
Such Brønsted acid site-induced anti-poisoning effect can be further demonstrated in the
283
stability testing (Figure 6b). At 250°C, a rapid deactivation of the pristine CeO2 NSs is observed
284
with DCM conversion decreased from 87% to 20% within 4 hours. For POx-CeO2 NSs with low
285
POx content (POx-CeO2-0.05 and POx-CeO2-0.1), although the deactivation is still recorded at
286
the initial stage (the conversion of DCM drops from 85-90% to 65-70% within 90 min), the
287
DCM conversion can be eventually stabilized at a much higher level compared with the pristine
288
CeO2 (approximately 65% vs. 20%). Very encouragingly, the decay of initial activity is
289
negligible on POx-CeO2-0.2 and POx-CeO2-0.3 NSs, and a stable DCM conversion of 65-70% is
290
delivered. The initial activity loss on POx-CeO2-0.05 and POx-CeO2-0.1 is correlated to their
291
relatively low coverage of POx species. With more POx attached on CeO2, the adsorbed Cl-
292
poisoning species can be rapidly removed in the form of HCl under the assistance of Brønsted
293
acid sites. Meanwhile, the re-dissociation of formed HCl to Cl- can be inhibited by Brønsted acid
294
site, which further strengthens the anti-poisoning property of the catalyst.25 It is noteworthy that
295
the higher reaction temperature is favourable to the subsequent removal of HCl as a final
ACS Paragon Plus Environment
16
Page 17 of 26
Environmental Science & Technology
296
product. Slightly increasing the stability testing temperature from 250 to 300°C leads to a
297
significantly enhanced DCM conversion stabilized at approximately 95%.
298
Our POx-CeO2 NSs obtained via a TMP process show significantly higher activity than CeO2
299
treated with inorganic phosphates, such as PA, AP and SP, as summarized in Figure 6c. The SP-
300
CeO2 catalyst presents the lowest activity (the conversion of DCM is only 40% even at 400°C),
301
which is probably due to the presence of Na+ and unsuccessful grafting of Brønsted acid sites.
302
The limited performance of PA-CeO2 and AP-CeO2 catalysts might be related to the formation of
303
cerium phosphate over their surfaces, which is detrimental to the surface redox property for
304
catalysis.
305
The surface basic sites on CeO2 and POx-CeO2 NSs are found to be important for controlling
306
the production of monochloromethane (MCM) by-product during DCM oxidation reaction
307
(Figure 6d). MCM is barely generated over the pristine CeO2 NSs with less than 0.5%
308
selectivity at all temperatures. The incorporation of surface POx drastically increases the
309
production of MCM at relatively low temperatures, reaching a maximum of 18% selectivity at
310
200°C for POx-CeO2-0.05 NSs. The transition from DCM to MCM normally involves two steps
311
in the absence of gaseous hydrogen: the dissociation of C-Cl bond and hydride transfer. The
312
latter one is rate-determining under oxidative condition and is widely observed on the medium-
313
strong basic sites in previous studies.47-48 Our TPD and Py-IR results have indicated that among
314
all POx-CeO2 NSs samples, POx-CeO2-0.05 presents the largest amount of medium-strong
315
surface basic site O2-, which well explains its highest selectivity for MCM. The higher POx
316
content can prohibit MCM generation, which is decreased to 5% at 200°C and 2% at 250°C on
317
POx-CeO2-0.2. Coupled with the excellent stability as aforementioned, it makes POx-CeO2-0.2
318
the most promising catalyst for DCM oxidation in our study. In addition, MCM by-product
ACS Paragon Plus Environment
17
Environmental Science & Technology
Page 18 of 26
319
disappears above 300°C on all the POx-modified CeO2 catalysts. All these point to a principle
320
that, when working at low temperatures, an ideal transition metal oxide catalyst for CVOCs
321
oxidation should minimize the medium-strong surface basic sites to efficiently prevent the
322
production of chlorinated organic by-products.
323
Based on the above discussion, mechanism of the enhanced oxidative DCM removal over the
324
as-prepared POx-CeO2 NSs catalysts is clearly derived. The employment of TMP ensures a
325
successful surface POx grafting without any change in surface CeO2 phase or sacrifice of surface
326
redox site (Ce4+/3+), leading to a balanced Lewis – Brønsted acid sites combination, which is the
327
key to achieve the highly efficient oxidative DCM removal on the CeO2-based catalysts: While
328
the Lewis acid sites (Ce4+/3+, also as the active redox sites) facilitate the C-Cl bond dissociation
329
and catalyze DCM oxidation, the presence of POx Brønsted acid sites effectively lower the
330
medium-strong basic sites (O2-) intensity, reducing the hydrogen transfer rate as well as the
331
MCM by-product formation. In addition, the Brønsted acid sites can accelerate the removal of
332
adsorbed chlorine species on active redox sites, substantially enhancing the catalyst stability. Our
333
study highlights that the rational surface decoration should be an effective way to improve the
334
catalysts’ performance in DCM removal, and also underlies a promising approach for other air
335
pollution treatment.
336
ASSOCIATED CONTENT
337
Supporting Information. Addition Schemes, SEM and TEM images, SEM-EDS elemental
338
mapping, and in situ DRIFTS spectra are provided. The supporting information is available free
339
of charge on the ACS Publications website.
340
AUTHOR INFORMATION
341
Corresponding Author
ACS Paragon Plus Environment
18
Page 19 of 26
Environmental Science & Technology
342
*E-mail: Sen Zhang (
[email protected]), Wangcheng Zhan (
[email protected])
343
Author Contributions
344
The manuscript was written through contributions of all authors. All authors have given approval
345
to the final version of the manuscript. #These authors contributed equally.
346
Notes
347
The authors declare no competing financial interest.
348
ACKNOWLEDGMENT
349
This work was supported by the National Key Research and Development Program of China
350
(No. 2016YFC0204300) and the National Natural Science Foundation of China (No. 21777043)
351
and Jeffress Trust Awards Program in Interdisciplinary Research from Thomas F. and Kate
352
Miller Jeffress Memorial Trust.
353
REFERENCES
354 355 356 357 358 359 360 361
1.
Atkinson, R.; Arey, J., Atmospheric Degradation of Volatile Organic Compounds. Chem.
Rev. 2003, 103, 4605-4638. 2.
Hitchman, M. L.; Spackman, R. A.; Ross, N. C.; Agra, C., Disposal Methods for
Chlorinated Aromatic Waste. Chem. Soc. Rev. 1995, 24, 423-430. 3.
Martin, E. T.; McGuire, C. M.; Mubarak, M. S.; Peters, D. G., Electroreductive
Remediation of Halogenated Environmental Pollutants. Chem. Rev. 2016, 116, 15198-15234. 4.
Henschler, D., Toxicity of Chlorinated Organic Compounds: Effects of the Introduction
of Chlorine in Organic Molecules. Angew. Chem. Int. Ed. 1994, 33, 1920-1935.
ACS Paragon Plus Environment
19
Environmental Science & Technology
362
5.
Page 20 of 26
González-Velasco, J. R.; López-Fonseca, R.; Aranzabal, A.; Gutiérrez-Ortiz, J. I.;
363
Steltenpohl, P., Evaluation of H-type Zeolites in the Destructive Oxidation of Chlorinated
364
Volatile Organic Compounds. Appl. Catal., B 2000, 24, 233-242.
365
6.
Aranzabal, A.; Pereda-Ayo, B.; González-Marcos, M.; González-Marcos, J.; López-
366
Fonseca, R.; González-Velasco, J., State of the Art in Catalytic Oxidation of Chlorinated Volatile
367
Organic Compounds. Chem. Pap., 2014, 68, 1169-1186.
368
7.
Guillemot, M.; Mijoin, J.; Mignard, S.; Magnoux, P., Volatile Organic Compounds
369
(VOCs) Removal Over Dual Functional Adsorbent/Catalyst System. Appl. Catal., B 2007, 75,
370
249-255.
371
8.
El Assal, Z.; Ojala, S.; Pitkäaho, S.; Pirault-Roy, L.; Darif, B.; Comparot, J.-D.; Bensitel,
372
M.; Keiski, R. L.; Brahmi, R., Comparative Study on the Support Properties in the Total
373
Oxidation of Dichloromethane over Pt Catalysts. Chem. Eng. J. 2017, 313, 1010-1022.
374 375 376 377 378 379
9.
Chen, Q.-Y.; Li, N.; Luo, M.-F.; Lu, J.-Q., Catalytic Oxidation of Dichloromethane over
Pt/CeO2–Al2O3 catalysts. Appl. Catal., B 2012, 127, 159-166. 10. Guillemot, M.; Mijoin, J.; Mignard, S.; Magnoux, P., Mode of Zeolite Catalysts Deactivation During Chlorinated VOCs Oxidation. Appl. Catal., A 2007, 327, 211-217. 11. Dai, Q.; Wang, X.; Lu, G., Low-Temperature Catalytic Combustion of Trichloroethylene over Cerium Oxide and Catalyst Deactivation. Appl. Catal., B 2008, 81, 192-202.
380
12. Yang, P.; Meng, Z.; Yang, S.; Shi, Z.; Zhou, R., Highly Active Behaviors of CeO2–CrOx
381
Mixed Oxide Catalysts in Deep Oxidation of 1,2-Dichloroethane. J. Mol. Catal. A: Chem. 2014,
382
393, 75-83.
ACS Paragon Plus Environment
20
Page 21 of 26
Environmental Science & Technology
383
13. Weng, X.; Sun, P.; Long, Y.; Meng, Q.; Wu, Z., Catalytic Oxidation of Chlorobenzene
384
over MnxCe1–xO2/HZSM-5 Catalysts: A Study with Practical Implications. Environ. Sci. Technol.
385
2017, 51, 8057-8066.
386
14. Li, H.; Lu, G.; Dai, Q.; Wang, Y.; Guo, Y.; Guo, Y., Efficient Low-Temperature
387
Catalytic Combustion of Trichloroethylene over Flower-like Mesoporous Mn-Doped CeO2
388
Microspheres. Appl. Catal., B 2011, 102, 475-483.
389
15. Wang, W.; Zhu, Q.; Dai, Q.; Wang, X., Fe doped CeO2 Nanosheets for Catalytic
390
Oxidation of 1,2-Dichloroethane: Effect of Preparation Method. Chem. Eng. J. 2017, 307, 1037-
391
1046.
392
16. Cai, T.; Huang, H.; Deng, W.; Dai, Q.; Liu, W.; Wang, X., Catalytic Combustion of 1,2-
393
Dichlorobenzene at Low Temperature over Mn-Modified Co3O4 Catalysts. Appl. Catal., B 2015,
394
166-167, 393-405.
395 396
17. Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P., Fundamentals and Catalytic Applications of CeO2-Based Materials. Chem. Rev. 2016, 116, 5987-6041.
397
18. Cargnello, M.; Wieder, N. L.; Montini, T.; Gorte, R. J.; Fornasiero, P., Synthesis of
398
Dispersible Pd@CeO2 Core−Shell Nanostructures by Self-Assembly. J. Am. Chem. Soc. 2010,
399
132, 1402-1409.
400
19. Cargnello, M.; Jaén, J. J. D.; Garrido, J. C. H.; Bakhmutsky, K.; Montini, T.; Gámez, J. J.
401
C.; Gorte, R. J.; Fornasiero, P., Exceptional Activity for Methane Combustion over Modular
402
Pd@CeO2 Subunits on Functionalized Al2O3. Science 2012, 337, 713-717.
403
20. Cargnello, M.; Doan-Nguyen, V. V. T.; Gordon, T. R.; Diaz, R. E.; Stach, E. A.; Gorte,
404
R. J.; Fornasiero, P.; Murray, C. B., Control of Metal Nanocrystal Size Reveals Metal-Support
405
Interface Role for Ceria Catalysts. Science 2013, 341, 771-773.
ACS Paragon Plus Environment
21
Environmental Science & Technology
Page 22 of 26
406
21. Nie, L.; Mei, D.; Xiong, H.; Peng, B.; Ren, Z.; Hernandez, X. I. P.; DeLaRiva, A.; Wang,
407
M.; Engelhard, M. H.; Kovarik, L.; Datye, A. K.; Wang, Y., Activation of surface lattice oxygen
408
in single-atom Pt/CeO2 for low-temperature CO oxidation. Science 2017, 358, 1419-1423.
409
22. Yang, C.; Yu, X.; Heißler, S.; Nefedov, A.; Colussi, S.; Llorca, J.; Trovarelli, A.; Wang,
410
Y.; Wöll, C., Surface Faceting and Reconstruction of Ceria Nanoparticles. Angew. Chem. Int. Ed.
411
2017, 56, 375-379.
412 413 414 415
23. Trovarelli, A., Catalytic Properties of Ceria and CeO2-Containing Materials. Catal. Rev. 1996, 38, 439-520. 24. Dai, Q.; Bai, S.; Wang, Z.; Wang, X.; Lu, G., Catalytic Combustion of Chlorobenzene over Ru-Doped Ceria Catalysts. Appl. Catal., B 2012, 126, 64-75.
416
25. Dai, Q.; Yin, L.-L.; Bai, S.; Wang, W.; Wang, X.; Gong, X.-Q.; Lu, G., Catalytic Total
417
Oxidation of 1,2-Dichloroethane over VOx/CeO2 Catalysts: Further Insights via Isotopic Tracer
418
Techniques. Appl. Catal., B 2016, 182, 598-610.
419
26. Dai, Q.; Wang, W.; Wang, X.; Lu, G., Sandwich-Structured CeO2@ZSM-5 Hybrid
420
Composites for Catalytic Oxidation of 1, 2-Dichloroethane: An Integrated Solution to Coking
421
and Chlorine Poisoning Deactivation. Appl. Catal., B 2017, 203, 31-42.
422 423 424 425 426 427
27. Kawi, S.; Shen, S. C.; Chew, P. L., Generation of Brønsted Acid Sites on Si-MCM-41 by Grafting of Phosphorus Species. J. Mater. Chem. 2002, 12, 1582-1586. 28. Li, D.; Zheng, Y.; Wang, X., Effect of Phosphoric Acid on Catalytic Combustion of Trichloroethylene over Pt/P-MCM-41. Appl. Catal., A 2008, 340, 33-41. 29. Wang, T.; Dai, Q.; Yan, F., Effect of Acid Sites on Catalytic Destruction of Trichloroethylene over Solid Acid Catalysts. Korean J. Chem. Eng. 2017, 34, 664-671.
ACS Paragon Plus Environment
22
Page 23 of 26
Environmental Science & Technology
428
30. Yao, X.; Wang, Z.; Yu, S.; Yang, F.; Dong, L., Acid Pretreatment Effect on the
429
Physicochemical Property and Catalytic Performance of CeO2 for NH3-SCR. Appl. Catal., A
430
2017, 542, 282-288.
431 432 433 434 435 436
31. Yi, T.; Zhang, Y.; Li, J.; Yang, X., Promotional Effect of H3PO4 on Ceria Catalyst for Selective Catalytic Reduction of NO by NH3. Chin. J. Catal. 2016, 37, 300-307. 32. Masui, T.; Hirai, H.; Imanaka, N.; Adachi, G., Characterization and Thermal Behavior of Amorphous Cerium Phosphate. Phys. Status Solidi A 2003, 198, 364-368. 33. Singh, B.; Im, H.-N.; Park, J.-Y.; Song, S.-J., Studies on Ionic Conductivity of Sr2+Doped CeP2O7 Electrolyte in Humid Atmosphere. J. Phys. Chem. C 2013, 117, 2653-2661.
437
34. Nelson, N. C.; Wang, Z.; Naik, P.; Manzano, J. S.; Pruski, M.; Slowing, I. I., Phosphate
438
Modified Ceria as a Brønsted Acidic/Redox Multifunctional Catalyst. J. Mater. Chem. A 2017, 5,
439
4455-4466.
440
35. Dai, Q.; Bai, S.; Li, H.; Liu, W.; Wang, X.; Lu, G., Template-Free and Non-
441
Hydrothermal Synthesis of CeO2 Nanosheets via a Facile Aqueous-Phase Precipitation Route
442
with Catalytic Oxidation Properties. CrystEngComm 2014, 16, 9817-9827.
443 444 445 446 447 448
36. Chunwen, S.; Hong, L.; ZhaoXiang, W.; Liquan, C.; Xuejie, H., Synthesis and Characterization of Polycrystalline CeO2 Nanowires. Chem. Lett. 2004, 33, 662-663. 37. Tan, H.; Wang, J.; Yu, S.; Zhou, K., Support Morphology-Dependent Catalytic Activity of Pd/CeO2 for Formaldehyde Oxidation. Environ. Sci. Technol. 2015, 49, 8675-8682. 38. Lu, M.; Wang, F.; Chen, K.; Dai, Y.; Liao, Q.; Zhu, H., The Crystallization and Structure Features of Barium-Iron Phosphate Glasses. Spectrochim. Acta, Part A 2015, 148, 1-6.
ACS Paragon Plus Environment
23
Environmental Science & Technology
Page 24 of 26
449
39. Manzoli, M.; Menegazzo, F.; Signoretto, M.; Cruciani, G.; Pinna, F., Effects of Synthetic
450
Parameters on the Catalytic Performance of Au/CeO2 for Furfural Oxidative Esterification. J.
451
Catal. 2015, 330, 465-473.
452
40. Hu, Z.; Wang, Z.; Guo, Y.; Wang, L.; Guo, Y.; Zhang, J.; Zhan, W., Total Oxidation of
453
Propane over a Ru/CeO2 Catalyst at Low Temperature. Environ. Sci. Technol. 2018, 52, 9531-
454
9541.
455
41. Comsup, N.; Panpranot, J.; Praserthdam, P., The effect of phosphorous precursor on the
456
CO oxidation activity of P-modified TiO2 supported Ag catalysts. Catal. Commun. 2010, 11,
457
1238-1243.
458
42. Wu, Z.; Cheng, Y.; Tao, F.; Daemen, L.; Foo, G. S.; Nguyen, L.; Zhang, X.; Beste, A.;
459
Ramirez-Cuesta, A. J., Direct Neutron Spectroscopy Observation of Cerium Hydride Species on
460
a Cerium Oxide Catalyst. J. Am. Chem. Soc. 2017, 139, 9721-9727.
461
43. Chang, H.; Wu, Q.; Zhang, T.; Li, M.; Sun, X.; Li, J.; Duan, L.; Hao, J., Design
462
Strategies for CeO2–MoO3 Catalysts for DeNOx and Hg0 Oxidation in the Presence of HCl: The
463
Significance of the Surface Acid–Base Properties. Environ. Sci. Technol. 2015, 49, 12388-
464
12394.
465 466
44. Henderson, M. A., Surface Chemistry of Trimethyl Phosphate on α-Fe2O3. J. Phys. Chem. C 2011, 115, 23527-23534.
467
45. Mäkie, P.; Persson, P.; Österlund, L., Solar Light Degradation of Trimethyl Phosphate
468
and Triethyl Phosphate on Dry and Water-Precovered Hematite and Goethite Nanoparticles. J.
469
Phys. Chem. C 2012, 116, 14917-14929.
470 471
46. Wu, Z.; Mann, A. K. P.; Li, M.; Overbury, S. H., Spectroscopic Investigation of SurfaceDependent Acid–Base Property of Ceria Nanoshapes. J. Phys. Chem. C 2015, 119, 7340-7350.
ACS Paragon Plus Environment
24
Page 25 of 26
472 473 474 475
Environmental Science & Technology
47. Deno, N. C.; Peterson, H. J.; Saines, G. S., The Hydride-Transfer Reaction. Chem. Rev. 1960, 60, 7-14. 48. Zhou, J.; Mullins, D. R., Adsorption and Reaction of Formaldehyde on Thin-Film Cerium Oxide. Surf. Sci. 2006, 600, 1540-1546.
476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494
ACS Paragon Plus Environment
25
Environmental Science & Technology
495
Page 26 of 26
Table of Contents
496
ACS Paragon Plus Environment
26