Subscriber access provided by READING UNIV
Environmental Processes
Hydroxyl radical based photocatalytic degradation of halogenated organic contaminants and paraffin on silica gel Ruijuan Qu, Chenguang Li, Jiaoqin Liu, Ruiyang Xiao, Xiaoxue Pan, Xiaolan Zeng, Zunyao Wang, and Jichun Wu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b00499 • Publication Date (Web): 11 Jun 2018 Downloaded from http://pubs.acs.org on June 11, 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 23
Environmental Science & Technology
1
Hydroxyl radical based photocatalytic degradation of halogenated
2
organic contaminants and paraffin on silica gel
3
Ruijuan Qu†, Chenguang Li†, Jiaoqin Liu†, Ruiyang Xiao‡, Xiaoxue Pan†, Xiaolan Zeng†, Zunyao
4
Wang†,*, Jichun Wu§,*
5
†
6
Nanjing University, Nanjing 210023, P. R. China
7
‡
8
University, Changsha 410083, P. R. China
9
§
10
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment,
Institute of Environmental Engineering, School of Metallurgy and Environment, Central South
Key Laboratory of Surficial Geochemistry, School of Earth Sciences and Engineering, Nanjing
University, Nanjing 210023, P. R. China
11
*
Corresponding
author.
Tel:
+86-25-89680358;
Fax:
[email protected];
[email protected]. 1
ACS Paragon Plus Environment
+86-25-89680358.
E-mail:
Environmental Science & Technology
12
Abstract
13
Photochemical materials are of scientific and practical importance in the field of
14
photocatalysis. In this study, the photochemistry of several organic contaminants, including
15
decabromodiphenyl ether (BDE-209), halogenated phenols (C6X5OH, X = F, Cl, Br) and paraffin,
16
on silica gel (SG) surface was investigated under simulated solar irradiation conditions. Photolysis
17
of these compounds at the solid/air interface proceeds with different rates yielding various
18
hydroxylation products, and hydroxyl radical was determined as the major reactive species.
19
According to density functional theory (DFT) calculations, the reaction of physically adsorbed
20
water with reactive silanone sites (>Si=O) on silica was indispensable for the generation of •OH
21
radical, where the required energy matches well with the irradiation energy of visible light. Then,
22
the BDE-209 was selected as a representative compound to evaluate the photocatalytic
23
performance of SG under different conditions. The SG material showed good stability in the
24
photodegradation process, and was able to effectively eliminate BDE-209 under natural sunlight.
25
These findings provide new insights into the potential application of silica gel as a solid surface
26
photocatalyst for contaminants removal.
27
2
ACS Paragon Plus Environment
Page 2 of 23
Page 3 of 23
Environmental Science & Technology
28
1. Introduction
29
Photocatalysis that makes use of a catalyst to accelerate a photoreaction is a promising
30
technique for the photocatalytic oxidation of organic pollutants. In this field, photochemical
31
materials have become scientifically and practically important, because they absorb light to
32
generate highly reactive species, particularly hydroxyl radicals, that are able to decompose
33
refractory organic substances via secondary reactions. The •OH radical (E = 2.8 V) is the second
34
most reactive species next to fluorine atom, and they attack a wide range of organic substrates
35
with rate constants usually in the order of 106−109 M−1s−1 1. Since the pioneer work of Fujishima
36
and Honda, many semiconductors such as TiO2 and ZnO have been tested and evaluated as
37
photocatalysts in ultraviolet (UV) light-activated processes2,3. By contrast, photocatalytic activity of
38
non-metal oxides is less well studied.
39
Silica is considered to be an interesting candidate material for future development of catalysts4.
40
Previous studies have shown that pure silica can promote some photocatalytic reactions under UV
41
irradiation, such as direct methane coupling5, photo-metathesis of propene6,7, and photo-epoxidation of
42
propene8. In recent years, the silica nanoparticles (SiO2 NPs) have received growing research
43
interests, due to its high efficiency towards catalytic degradation of water contaminants. The SiO2
44
NPs prepared from tetraethylorthosilicate and rice husk ash were found to be photocatalytically
45
active in the degradation of methyl red dye9,10. Anastasescu et al.11 revealed that tubular SiO2
46
could behave as an efficient photocatalyst for the oxidation of oxalic acid to carbon dioxide. Since
47
most of organic contaminants are slightly soluble or insoluble in water, it is of great significance
48
to develop potential methods for removal of hydrophobic compounds in contaminated
49
environments.
50
Silica gel (SG) is a typical manmade amorphous silica material, which lacks long-range order
51
and possesses a principally different siloxane framework architecture from crystalline silica12. It
52
has been widely used as catalyst supports, because of its large surface area, high optical
53
transparency in the UV/Vis region and excellent characteristics in adsorption of pollutants13-15.
54
Chromatography SG was shown to increase its photocatalytic activity for the CO oxidation
55
reaction with O2 under UV irradiation16. Some researchers have examined the photochemical
56
behavior of organic contaminants, especially polycyclic aromatic hydrocarbons (PAHs), at the
57
SG/air interface. Silica gel catalyzed the photochemical transformation of benzo[e]pyrene and 3
ACS Paragon Plus Environment
Environmental Science & Technology
58
perylene, and the reaction was thought to occur through a radical cation-mediated process17,18. For
59
the photochemical oxidation of phenanthrene on SG, the addition of singlet molecular oxygen (1O2)
60
to the ground state molecules was determined as the initial reaction step19. Sotero and Arce20
61
reported that the irradiation of perylene on SG with 320−510 nm light induced its degradation, and
62
the participation of perylene radical cation and 1O2 was confirmed in the photodegradation process.
63
Photoirradiation of pyrene/SG samples was suggested to involve three primary intermediate
64
species, pyrene radical cation, superoxide radical anion (O2•–) and hydrogen radical (H•), among
65
which O2•– was generated from O2 via capture of trapped electrons on SG surface21 . The
66
decomposition of dibenzo-p-dioxin and 4-chlorophenol on the surface of SG occurs under UV-C
67
light, and the formation of biradical and carbine via excited state molecules was the primary
68
degradation step, respectively22,23. Under external force (e.g. grinding, heat), surface-associated
69
reactive free radicals (silicon-based radicals, O2•– radicals, oxyradicals, CO2• radicals and •OH
70
radicals) are generated on silica particles24,25. In a recent work, Romanias et al.26 found that
71
surface bound •OH radical was the dominant reactive species in pyrene oxidation on solid Al2O3
72
surface, leading to the formation of 28 gas/solid phase products. However, whether
73
photoirradiation can produce •OH radical for the catalytic degradation of organic contaminants on
74
the silica surface still remains unknown.
75
In this work, we examined the surface photochemical reactions of organic chemicals to
76
explore the photocatalytic behavior of SG. The commercial SG material was first characterized by
77
modern analytical techniques for its morphological structures and photoelectrochemical properties.
78
Then, decabromodiphenyl ether (BDE-209), halogenated phenols, and paraffin, as typical
79
environmental contaminants belonging to different categories, were selected as model compounds
80
to conduct a kinetics and product study. These compounds showed different photoreactivity at the
81
SG/air interface under simulated solar irradiation, and many hydroxylation products were detected
82
by mass spectrometry analysis. Mechanistic studies indicate the involvement of surface •OH
83
radical in the photooxidation process, and the generation of •OH radical was further rationalized
84
by density functional theory (DFT) calculations. Finally, the photocatalytic performance of SG
85
was evaluated under various conditions, with BDE-209 as a representative. To the best of our
86
knowledge, this is the first attempt to systematically elucidate the •OH radical initiated
87
photochemical reactions on SG surface. 4
ACS Paragon Plus Environment
Page 4 of 23
Page 5 of 23
Environmental Science & Technology
88
2. Materials and methods
89
2.1 Chemicals and reagents
90
Decabromodiphenyl ether (BDE-209, purity ≥ 98%) was provided by Tokyo Chemical
91
Industry Co., Ltd. Silica gel (SG, particle size: 200-300 mesh) was purchased from Qingdao
92
Haiyang Chemical Co., Ltd. The chromatographic grade methanol, tetrahydrofuran and n-hexane
93
were supplied by Merck Co., Ltd. Derivatization reagent bis(trimethylsilyl)trifluoroacetamide
94
(BSTFA)/trimethylchlorosilane (TMCS) (99:1, v/v), provided by AccuStandard, was diluted 100
95
times with n-hexane for use. Other chemicals, including halogenated phenols (C6X5OH, X = F, Cl,
96
Br) and solid paraffin were obtained from commercial sources and used as received.
97
Radical-trapping
98
2,2,6,6-tetramethyl-4-piperidinol (TEMP, 99%) was purchased from J&K Co., Ltd. and Tokyo
99
Chemical Industry Co., Ltd., respectively.
100
agent
5,5-dimethylpyrroline-N-oxide
(DMPO,
purity
≥
99%)
and
2.2 Characterization of silica gel
101
Field emission scanning electron microscopy (FESEM) images were obtained on a S-3400N
102
II scanning electron microscope (Hitachi, Japan) at an acceleration voltage of 20 kV. Transmission
103
electron microscopy (TEM) images were observed by a JEM-200CX electron microscope (JEOL,
104
Japan). X-ray diffraction (XRD) pattern was recorded on a X’TRA X-ray diffraction-meter (ARL,
105
Switzerland). Fourier transform infrared spectra (FT-IR) were analyzed in the range of 400−4000
106
cm-1 by a Nexus 470 FTIR spectrophotometer (ARL, Switzerland) using KBr pellet technique.
107
The Raman spectra was examined in the 300−1500 cm−1 region using an XploRA ONE Raman
108
spectrometer (Horiba, France) with 785 nm wavelength laser excitation. Specific surface area was
109
measured by the Brunauer−Emmett−Teller (BET) method, and nitrogen adsorption–desorption
110
isotherms were obtained at 77 K using an ASAP 2020 apparatus (Micromeritics, USA). Pore size
111
distribution
112
Barrett−Joyner−Halenda (BJH) model. The photoelectrochemical properties were measured in a
113
standard three-electrode system by a CHI760D electrochemical workstation (Shanghai, China).
114
Ag/AgCl electrode and platinum wire served as the reference and counter electrode, respectively.
115
The silica gel loaded onto ITO electrode (an active area of ca. 1 cm2) was used as the working
116
electrode, which was irradiated with light from a 500W Xenon lamp. 0.1 M Na2SO4 solution was
117
used as the electrolyte.
was
derived
from
the
desorption
branch
5
ACS Paragon Plus Environment
of
isotherms
using
the
Environmental Science & Technology
118
2.3 Sample preparation and photolysis experiment
119
To prepare samples for irradiation, 0.2 mL of BDE-209 stock solution (25 mg/L) in
120
tetrahydrofuran was added to 0.1 g of SG in each cylindrical quartz tube. The dispersion was
121
magnetically stirred, and then evenly applied to the inner wall of the quartz tube for about half its
122
height from the base. After the solvent was removed by drying at 40 oC in an oven, the tubes were
123
cooled down to room temperature and sealed with glass stoppers for irradiation. According to the
124
weight of BDE-209 and SG, the surface coverage was calculated as 50 µg/g. The loading
125
concentration was selected according to previous studies on the photodegradation of BDE-209 in
126
solid phase27-29. Preloaded samples of halogenated phenols and paraffin on SG were achieved in a
127
similar manner, except for the stock solution that was prepared in methanol and n-hexane,
128
respectively.
129
The irradiation experiment was performed in an XPA-1 photochemical reactor (Nanjing
130
Xujiang Electromechanical Plant, China) equipped with a 500 W Xe lamp as the light source. The
131
lamp was vertically placed inside a quartz cooling water jacket to prevent it from overheating.
132
During irradiation, the tubes were rotated around the lamp by a merry-go-round apparatus for
133
uniform light exposure. The light intensity in the sample region was measured as 74.2 mW·cm-2
134
by a FZ-A Radiometer with a 400-1000 nm sensor (Photoelectric Instrument Factory of Beijing
135
Normal University, China). At designated time intervals, tubes were taken from the photoreactor.
136
The unreacted BDE-209 was extracted from the solid surface by adding 2.0 mL of tetrahydrofuran
137
to the tubes, followed by ultrasonic treatment for 10 min. The suspension was then centrifuged for
138
5 min to remove any suspended particles for GC analysis. The extraction efficiency of BDE-209
139
was about 98.1 %, and all kinetics experiments were performed in triplicate. To detect possible
140
products, 0.1 g 400 µg/g BDE-209/SG sample was irradiated for different time, and then
141
ultrasonically extracted with 2.0 mL of methanol for liquid chromatograph-time of flight-mass
142
spectrometry (LC-TOF-MS) analysis. In another batch of experiments, reaction samples of
143
BDE-209/SG were extracted with 2.0 mL of tetrahydrofuran for gas chromatography-mass
144
spectrometry (GC-MS) analysis. Sample pretreatment procedures of other compounds were
145
summarized in Table S1.
146
The degradation of BDE-209/SG was also evaluated under natural sunlight irradiation from
147
9:00 am to 3:00 pm for two successive days (December 1−2, 2016). The quartz tubes were placed 6
ACS Paragon Plus Environment
Page 6 of 23
Page 7 of 23
Environmental Science & Technology
148
on the roof top of our school building. The solar intensity was measured frequently by the FZ-A
149
radiometer (Figure S1).
150
2.4 Mass balance analysis
151
The actual surface coverage was measured by extracting the preloaded samples with organic
152
solvent. Then, the percentage of spiking was calculated from the ratio of the measured
153
concentration to the theoretical value. The un-extractable parts were estimated using adsorption
154
experiments. Adsorption was performed by adding 0.1g of SG into 10 mL of each compound
155
solution and shaking the samples for 24 h at 25 °C. Following centrifugation, the concentration of
156
target compound in the supernatant was analyzed.
157
2.5 Analysis methods
158
An Agilent 6890A gas chromatography equipped with an autosampler and an
159
electron-capture detector (ECD) was used for quantification of BDE-209. 2 µL of sample was
160
injected in splitless mode, and the separation was performed on a DB-XLB column (10 m × 0.25
161
mm, 0.25 µm, J&W Scientific, USA). The carrier gas was nitrogen (N2) at a constant flow rate of
162
3.0 mL/min. The injector and detector temperature was set at 280 °C and 320 °C, respectively. The
163
column temperature was programmed from 60 °C (2 min hold) to 290 °C at a rate of 20 ºC/min,
164
then from 290 °C to 310 ºC at 8 ºC/min and held isothermally at 310 ºC for 14 min. Analysis
165
methods of other compounds were listed in Table S1.
166
The LC-TOF-MS analysis was performed on an Agilent 1260 Infinity HPLC system
167
connected to a quadrupole time-of-flight Triple TOF 5600 mass spectrometer (AB SCIEX, USA).
168
For detailed information, please refer to Text S1.
169
A Thermo TRACE Ultra-TSQ Quantum GC-MS system (Thermo Scientific, USA) was used
170
to identify possible degradation products in BDE-209/SG reaction samples. Specific information
171
is detailed in Text S2.
172
Paraffin oxidation products were derivatized using BSTFA/TMCS and analyzed by a Thermo
173
Trace Ultra-DSQ II GC-MS system (Thermo Scientific, USA) to illustrate the generation of
174
alcohols. The detailed analytical procedure could be seen in Text S3.
175
2.6 Detection of free radicals
176
The electron paramagnetic resonance (EPR) spectra were recorded on a Bruker EMX-10/12
177
EPR spectrometer. Production of •OH and O2•− in the reaction mixtures was monitored using 7
ACS Paragon Plus Environment
Environmental Science & Technology
178
DMPO as spin-trapping agent, while 1O2 was detected with TEMP as spin-trapping agent. 0.025 g
179
SG spiked with 0.05 mL of 100 mM DMPO/TEMP was irradiated for 25 min, and the reaction
180
samples were then diluted by 0.45 mL of water for •OH/1O2 detection. Due to the facile
181
disproportionation in water, formation of the O2•− radical was examined in ethanol solution. The
182
detailed instrumental parameters were set as follows: scanning frequency: 9.7 GHz; central field:
183
3480 G; sweep width: 200 G; microwave power: 20 mW; scanning temperature: 298 K.
184
2.7 Theoretical calculations
185
The density functional theory (DFT) calculations were performed at the B3LYP/LanL2DZ
186
level using the Gaussian 09 software package30. The structural geometry was optimized with the
187
keyword “Opt”, and vibrational frequencies of all stationary points were calculated to identify the
188
structures obtained as true minima or first-order saddle point. The Gibbs free energies (G) were
189
directly obtained from Gaussian output files.
190
3. Results and Discussion
191
3.1 Sample characterization
192
The nitrogen adsorption-desorption isotherms of SG are shown in Figure 1A. Type IV
193
isotherms with a hysteresis loop were observed, indicating that the material is mesoporous in
194
nature. From the pore size distributions (see the inset of Figure 1A), the sample presents a very
195
narrow distribution of pore diameter ranging from 3 to 15 nm. The BET specific surface area was
196
calculated to be 337.9 m2/g, and the average pore diameter was 8.36 nm.
197
Photoelectrochemical measurements are usually employed to qualitatively study the
198
separation and transfer of photogenerated charge carriers. In this work, the transient photocurrent
199
response of SG was recorded for several on-off cycles of light irradiation. As shown in Figure 1B,
200
the photocurrent increased sharply as soon as the lamp was switched on, while immediately
201
returned to its initial value as soon as the irradiation was terminated, suggesting that the generation
202
of photocurrent was completely attributed to the photoreactivity of the electrode. Moreover, the
203
photoelectrode displayed good reproducibility and stability as the photocurrent intensity did not
204
exhibit obvious changes under several on-off cycles. The steady transient photocurrent density of
205
SG was approximately 0.1 µA/cm2. Characterization results of FESEM, TEM, XRD, FT-IR and
206
Raman are provided in Text S4 and Figure S2.
8
ACS Paragon Plus Environment
Page 8 of 23
Page 9 of 23
Environmental Science & Technology
400 300 200
2
0.20 0.15 0.10 0.05 0.00
0
10
20
30
40
Pore diameter (nm)
100 0 0.0
Control Silica gel
0.25
Photocurrent (µΑ /cm )
500
B 0.16 Pore volume (cm3/gnm)
Quantity adsorbed (cm3/g)
A 600
Adsorption Desorption
0.2
0.4
0.6
0.8
1.0
light on
0.12
0.08
0.04 light off
0.00 0
207
50
100
150
200
Irridiation time (s)
Relative pressure (P/P0)
208
Figure 1 (A) Nitrogen adsorption-desorption isotherms and (B) transient photocurrent response of
209
silica gel.
210
3.2 Mass balance analysis
211
With respect to total mass of the chemicals, (1) some of them were removed during solvent
212
evaporation, (2) some of them were strongly adsorbed onto SG, which are not possible to extract,
213
and (3) some of them were adsorbed onto SG, which are possible to extract. Thus, a mass balance
214
analysis was established to evaluate the reliability of experimental results. As shown in Table S2,
215
the percentage of spiking, indicative of the sample loss from solvent volatilization and the
216
un-extractable loss, ranges from 92.9±0.7% for BDE-209 to 98.5±3.4% for pentachlorophenol.
217
When adsorption equilibrium was achieved between solid particles and liquid phase, the
218
chemicals in solutions accounted for an overwhelming proportion of the total mass
219
(97.1%−99.5%). This suggests only very small parts is not possible to extract from preloaded
220
samples. Considering the negligible loss in sample preparation and extraction process, nominal
221
concentrations were used throughout the work.
222
3.3 Degradation products of BDE-209, halogenated phenols and paraffin
223
Foil covered dark control samples all exhibited no loss of target compound during the
224
reaction. In the Xe lamp irradiation system, BDE-209 was found to degrade with time, giving a
225
removal rate of 91.5% after 60 min (Figure 2A). Söderström and co-workers31 investigated the
226
photolytic debromination of BDE-209 on SG using mercury UV-lamp (1.6 mW/cm2, 300−400
227
nm), and they also observed a fast degradation of BDE-209 with a half-life time of
280
90% matching), this compound was assigned as Heptadecan-1-ol trimethylsilyl ether. Similarly,
281
the chromatographic peak at 24.26 min having a parent ion of m/z 356 was identified as
282
Nonadecan-1-ol trimethylsilyl ether. These results show that SG can also catalyze the degradation
283
of halogenated phenols and paraffin, resulting in the formation of hydroxylation products. Since
284
attack by •OH radical leads to hydroxylation products29,35,36, it is speculated that photodegradation
285
of organic contaminants on the surface of SG mainly involves •OH radical initiated oxidation
286
reaction. A
Cl
F OH C6H2F4O2
Cl
F
287
OH
OH F
F
OH Cl
Cl
Cl OH C6H2Cl4O2
HO
OH Cl
Cl
OH Cl
Br
OH Cl O Cl O C6HCl3O3 C6H3Cl3O3
Br
OH
OH Br
Br
Br Br
Br
Br HO O OH Br OH O Br C6HBr3O3 C6H2Br4O2 C6H3Br3O3
12
ACS Paragon Plus Environment
Page 13 of 23
Environmental Science & Technology
288 289
Figure 4 (A) The representative degradation products of halogenated phenols identified by
290
LC-TOF-MS; (B) The GC-MS mass spectra of two derivatization products during photochemical
291
degradation of paraffin on silica gel.
292
3.4 Determination of the major reactive species
293
The EPR technology was used to characterize the radical species involved in the reaction
294
system. As shown in Figure 5, no peaks of free radicals (i.e., •OH, 1O2 and O2•−) were observed in
295
the reaction solutions in the dark. Four characteristic peaks of a 1:2:2:1 quartet pattern (aN = aH =
296
14.9 G) were detected under light irradiation (Figure 5A), indicating the generation of •OH radical.
297
TEMP could trap singlet oxygen, resulting in the spin adduct 2,2,6,6-tetramethylpiperidine-1-oxyl
298
(TEMPO). From Figure 5B, a 1:1:1 triplet signal characteristic of TEMPO radical (aN = 7.2 G, aH
299
= 4.1 G) were observed, demonstrating the existence of 1O2. Furthermore, the EPR spectra showed
300
six characteristic peaks of the DMPO-O2•− spin adducts (aN = 14 G, aH = 8 G) in the ethanol
301
solution (Figure 5C). After addition of SG, the intensity of the TEMPO and DMPO-O2•− peaks
13
ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 23
302
changed slightly, while the formation of DMPO-•OH spin adducts was significantly induced.
303
These spin-trap experiments revealed that •OH radical contributed mainly to the photochemical
304
transformation of halogenated organic contaminants and paraffin on SG. DMPO-OH
DMPO-25min
B
SG-0min
3450 3460 3470 3480 3490 3500 3510
305
SG-25min
C
TEMP-1O2
Intensity(a.u)
SG-25min
Intensity(a.u)
Intensity(a.u)
A
TEMP-25min
3440
3460
DMPO-O2¯
DMPO-25min
SG-0min
SG-0min
3480
3500
3520
Magnetic field (G)
Magnetic field (G)
SG-25min
3440
3460
3480
3500
3520
Magnetic field (G)
306
Figure 5 EPR spectra of (A) DMPO-•OH adduct, (B) TEMP-1O2 adduct, and (C) DMPO-O2•−
307
adduct formed in the silica gel (SG) system. Note that superoxide anion (O2•−) was measured in
308
ethanol solution.
309
In addition, the UV/H2O2 experiment that generates •OH radical as reactive species was
310
conducted to confirm the above conclusion. Due to the very strong hydrophobicity of BDE-209
311
and paraffin, the experiment was performed on halogenated phenols (see Text S5 for experimental
312
details). As shown in Figure S7, the typical hydroxylation products generated from photolysis of
313
halogenated phenols on SG were also detected in the UV/H2O2 oxidation process. Moreover,
314
experimental observations and quantum chemical investigations have revealed that atmospheric
315
photooxidation of polybrominated diphenyl ethers (PBDEs) by •OH, taking 4,4’-dibromodiphenyl
316
ether (BDE-15) as an example, leads to the formation of OH-PBDEs and brominated phenols as
317
reaction products37,38. These results provide reliable evidence for the role of •OH radical in
318
photodegradation of organic compounds on SG surface.
319
3.5 Detection of hydroxyl radical
320
As illustrated in Figure 6A, the DMPO-•OH adduct was clearly observed in the SG system
321
after Xe lamp irradiation for 25 min. The •OH signal intensity increased with irradiation time, due
322
to the accumulation of •OH generated from the surface of SG. Under the sunlight irradiation, the
323
•OH radical was also detected in SG reaction solution (Figure 6B). In presence of cut-off filters,
324
the signal intensity of •OH became lower than the control, and no radical was observed when 14
ACS Paragon Plus Environment
Page 15 of 23
Environmental Science & Technology
325
orange filters were used (Figure 6C). Considering the light transmittance of different cut-off filters
326
(Figure 6D), we can conclude that short wavelength irradiation is more favorable to the generation
327
of •OH radical.
A
B
25 min
SG-sun
Intensity (a.u.)
Intensity (a.u.)
5 min
1.5 min
SG-Xe
0 min
3460
3470
3480
3490
3500
3510
3400
3440
Magnetic field (G)
C
3520
3560
D 100 Light transmittance (%)
orange yellow green
Intensity (a.u.)
3480
Magnetic field (G)
cyan purple uv no filter
100
80
80
60
60
40
40
20
20
0 200
400
600
Relative intensity (%)
3450
0 1000
800
Wavelength (nm) 3400
328
3440
3480
3520
red cyan
3560
Magnetic field (G)
orange purple
yellow uv
green lamp
329
Figure 6 (A) EPR spectra of the silica gel (SG) system after irradiation by the Xe lamp for
330
different time; (B) EPR spectra of the SG reaction mixture under Xe lamp and solar irradiation; (C)
331
EPR spectra of DMPO−•OH adduct in the SG system with different cut-off filters; (D) The
332
emission spectrum of the lamp and the light transmittance of different cut-off filters.
333
3.6 Theoretical analysis for the generation of •OH radical
15
ACS Paragon Plus Environment
Environmental Science & Technology
57
Page 16 of 23
∆rG (kJ/mol)
OH + BDE-209 reaction
TS1 56.19
56 55 TS6 54.10 TS3 53.68 TS5 53.14
54 53
TS2 52.05
52
TS4 50.84
51 50 0
OH
0
OH generation
OH+BDE-209
F
Br
path
way
-100
-200
OH + Br
E pa th wa yA pa t hw ay C
th w ay
D ay
(434.29,435.35)(428.31)
-500 Si-O-Si
Br
Br Br +
HO Br
O (PR1)
OH
+ Br
Br
(540.19)
PR5 -152.97
Br4
O Br
-152
Br
Br
Br
PR2 -151.96
Br
Br
Br
PR3 -151.84
Br
Br
Br
3MRs H2O (389.32) Si10-OH Si10=O (418.05)
B
-150
-151
(211.51)
pa
ay pa
w th
pa th
w
-400
PR4 -150.25
Br 3 4 Br 5 Br
Br
Si10=O-H2O
-300
334
Br
Br 2 O1 6 BrBr
(PR2-PR6)
PR6 -159.75
-153 -160 -170
PR1 -178.82
-180
335
Figure 7 Calculated Gibbs free energies for possible generation of •OH radical and its subsequent
336
reaction with BDE-209 at the B3LYP/LANL2DZ level.
337
Previous studies have suggested the presence of some radical species on silica surface,
338
however, the generation mechanism of •OH radical still remains unclear39-41 . In this work,
339
theoretical calculations were employed to gain a deep insights into how •OH radical is produced
340
on the surface of SG. Figure 7 shows energy profiles for possible generation of •OH radical and its
341
subsequent reaction with BDE-209. Details of the six reaction pathways for •OH generation are
342
given in Figure S8. In the first case, homolytic cleavage of water molecules leads to the formation
343
of •OH and •H radicals (Figure S8A). The Gibbs free energy of reaction (∆rG) was estimated as
344
418.05 kJ/mol, corresponding to the irradiation energy of 286 nm UV light. Then, the [Si8O25H18]
345
model from the work of Narayanasamy and Kubicki42 was adopted for calculations. Surface ≡Si•
346
and ≡SiO• radicals are generated from homolytic cleavage of siloxane bonds (Si−O−Si) on the
347
silica surface, and the ≡SiO• radicals can easily react with H2O to produce •OH (Figure S8B).
348
Obviously, cleavage of the Si−O−Si bond is the energy-demanding step, which requires an energy
349
as high as 540.19 kJ/mol. Three-membered rings (3MRs), formed at high temperature and
350
“frozen-in” by rapid quenching in fumed silica, have been established as precursors of
351
oxyradicals40,41. Based on our calculations, the reaction of surface biradicals with water to 16
ACS Paragon Plus Environment
Page 17 of 23
Environmental Science & Technology
352
generate •OH radical can occur spontaneously, but an additional energy of 389.32 kJ/mol is
353
needed for homolytic cleavage of 3MRs (Figure S8C).
354
When the crystal structure model of silica [Si10O28H16] was chosen, direct bond-breaking of
355
surface silanol groups would yield •OH radical, and the energy required for the two types of
356
Si−OH groups is very close, 435.35 kJ/mol and 434.29 kJ/mol, respectively (Figure S8D). Owing
357
to the presence of >Si=O sites on silica surface42, it is reasonable to slightly modify the previous
358
model [Si10O28H16] by replacing geminal SiOH group with the silanone >Si=O sites. As for the
359
newly constructed structure model [Si10O22H4], a similar energy of 428.31 kJ/mol should be
360
provided to create •OH radical by cleavage of silanol group (Figure S8E). Another possibility for
361
the generation of •OH could be the reaction of physically adsorbed water with reactive silanone
362
sites (>Si=O) on silica, which only consumes energy of 211.51 kJ/mol (Figure S8F). This energy
363
is equal to the irradiation of 566 nm visible light, largely consistent with above EPR analysis
364
results that the •OH radical was still detected until the use of orange cut-off filters transmitting
365
light of wavelength above 550 nm. Since the Xe lamp used in this work emits lights at
366
wavelength > 290 nm, cleavage of 3MRs via absorbing at least the energy of 307 nm photon may
367
also contribute slightly to the formation of •OH.
368
Next, the •OH radical would attack BDE-209, producing different OH-Nona-BDEs species
369
and a bromine atom (PR2-PR6), and pentabromophenol and phenoxyl radical (PR1) via the
370
transition states (see details in Text S6 and Figure S9). The fast combination of bromine
371
atom/phenoxyl radical and the silicon-based radical formed in the first step makes it available to
372
regenerate the [Si10O22H4] structure with associated hydrogen bromide/pentabromophenol. In this
373
sense, SG could catalyze the decomposition of H2O to generate •OH continuously. Furthermore,
374
the M06-2x functional is also used to test the robustness of the DFT results. As shown in Table S4,
375
the ∆rG values calculated at the M062X/LANL2DZ level are largely consistent with the data at the
376
B3LYP/LANL2DZ level, suggesting that it is reliable to interpret possible •OH generation
377
pathways by the calculation results.
378
3.7 Degradation kinetics of BDE-209 under different conditions
379
The dependence of BDE-209 decomposition on the wavelength was investigated using
380
different cut-off filters (Figure 8A). Obviously, BDE-209 losses on SG followed the order of no
381
filter > purple > UV > cyan > green > yellow, while the red and orange light caused no 17
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 23
382
degradation of BDE-209. This is related to the light transmittance of different cut-off filters and
383
the emission spectrum of the lamp (Figure 6D). Except for the UV, the removal efficiency
384
decreased with increasing wavelength of light. Although the photon energy of the UV light is
385
greater than that of the purple light, the emission intensity of the lamp is very low in the reaction
386
wavelength region, thus leading to a smaller reaction rate. When a cut-off filter of 550 nm (orange)
387
was used, the SG photocatalyst finally lost its activity. This agrees well with the EPR results that
388
the •OH radical was not observed in presence of orange filters, further confirming the dominant
389
•OH radical oxidation reaction mechanism on SG surface.
390
The reaction process was repeated five times to assess the stability and reusability of the SG
391
for BDE-209 degradation. The SG after reaction was collected, washed by tetrahydrofuran to
392
remove the residual reactant, and then dried in a vacuum oven for use. As seen in Figure 8B, the
393
removal rate of BDE-209 in the fifth cycling run was not remarkably reduced, still reaching 90.1%
394
after 6 h. Thus, the SG possessed adequate properties in catalyzing the degradation of BDE-209. When exposed to the natural sunlight, the SG can also induce the effective degradation of
395
BDE-209, and the removal efficiency was 94.2% after 11 h (Figure 8C). B
1.0
0.8
0.8
0.8
0.6
0.6
0.6 0.4 0.2 0.0 0
40
80
120
Time (min)
397
C 1.0
1.0
Ct/C0
Ct/C0
A
red cyan
orange uv
yellow purple
1st
2nd
3rd
4th
5th
Ct/C0
396
0.4 0.2 0.0
green 0 no filter
0.4 0.2
60
120
180
240
300
Time (min)
0.0
0
2
4
6
8
10
12
Time (h)
398 399
Figure 8 (A) Wavelength dependence of BDE-209 decomposition on silica gel (SG) under Xe
400
lamp irradiation; (B) The BDE-209 degradation kinetics in five recycling runs of SG; (C) The
401
degradation of BDE-209/SG under natural sunlight irradiation. Reaction conditions: sample mass
402
= 0.1 g, [BDE-209] = 50 µg/g.
403
Environmental Implications
404
Our results indicated that the reaction of physically adsorbed water with reactive silanone
405
sites (>Si=O) on photoactivated silica gel produced surface •OH radical to decompose various
406
organic contaminants, resulting in the formation of hydroxylation products. This sheds a new light
18
ACS Paragon Plus Environment
Page 19 of 23
Environmental Science & Technology
407
on the application of SG as a simple and cost-effective photocatalyst for pollutants degradation.
408
The visible light responsive property of this material (active wavelength < 550 nm) is promising in
409
photocatalysis, as visible light covers the largest proportion of the solar irradiation beyond Earth’s
410
atmosphere. Because of its porous structure, large surface area and high surface energy, SG has
411
been widely used as adsorbents for the cleaning of polluted air. Further decomposition of air
412
pollutants by light irradiation, instead of just accumulating them, is expected to make SG
413
attractive in the field of air purification. Moreover, silicon dioxide is the most abundant
414
component of mineral dust that contributes to the largest mass emission rate of atmospheric
415
aerosol particles at a global scale43, and therefore photodegradation on mineral aerosols during
416
long range atmospheric transport may be an important fate process for hydrophobic organic
417
pollutants in the environment. The photochemical transformation into hydroxylation products
418
seems to be of interest in terms of health risk assessment due to the changed molecular polarity
419
and biological activity.
420 421
ASSOCIATED CONTENT
422
Supporting Information
423
Texts S1-S6, Figures S1-S9, and Tables S1-S4 can be found in the Supporting Information. This
424
material is available free of charge via the Internet at http://pubs.acs.org.
425
AUTHOR INFORMATION
426
Corresponding Author
427
* Phone: +86-25-89680358. Fax: +86-25-89680358.
428
E-mail:
[email protected] (Zunyao Wang);
[email protected] (Jichun Wu).
429
Notes
430
The authors declare no competing financial interest.
431
ACKNOWLEDGMENTS
432
This research was financially supported by the National Natural Science Foundation of China (No.
433
21607073; 21577063) and the Natural Science Foundation of Jiangsu Province (No.
434
BK20160651).
435 436
References 19
ACS Paragon Plus Environment
Environmental Science & Technology
(1) Hoigné, J.; Bader, H. Rate constants of reactions of ozone with organic and inorganic compounds in water—II: dissociating organic compounds. Water Res. 1983, 17, 185−194. (2) Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37−38. (3) Cheng, M.; Zeng, G.; Huang, D.; Lai, C.; Xu, P.; Zhang, C.; Liu, Y. Hydroxyl radicals based advanced oxidation processes (AOPs) for remediation of soils contaminated with organic compounds: A review. Chem. Eng. J. 2016, 284, 582−598. (4) Arshad, A.; Iqbal, J.; Mansoor, Q.; Ahmed, I. Graphene/SiO2 nanocomposites: The enhancement of photocatalytic and biomedical activity of SiO2 nanoparticles by graphene. J. Appl. Phys. 2017, 121, 244901. (5) Yuliati, L.; Tsubota, M.; Satsuma, A.; Itoh, H.; Yoshida, H. Photoactive sites on pure silica materials for nonoxidative direct methane coupling. J. Catal. 2006, 238, 214−220. (6) Inaki, Y.; Yoshida, H.; Hattori, T. Two photoexcitation steps for photometathesis of propene over FSM-16. J. Phys. Chem. B 2000, 104, 10304−10309. ( 7 ) Yoshida, H.; Kimura, K.; Inaki, Y.; Hattori, T. Catalytic activity of FSM-16 for photometathesis of propene. Chem. Commun. 1997, 1, 129−130. (8) Yoshida, H.; Murata, C.; Hattori, T. Screening study of silica-supported catalysts for photoepoxidation of propene by molecular oxygen. J. Catal. 2000, 194, 364−372. (9) Badr, Y.; El-Wahed, M. A.; Mahmoud, M. A. Photocatalytic degradation of methyl red dye by silica nanoparticles. J. Hazard. Mater. 2008, 154, 245−253. (10) Vinoda, B. M.; Vinuth, M.; Bodke, Y. D.; Manjanna, J. Photocatalytic degradation of toxic methyl red dye using silica nanoparticles synthesized from rice husk ash. J. Environ. Anal. Toxicol. 2015, 5, 336. (11) Anastasescu, C.; Zaharescu, M.; Balint, I. Unexpected photocatalytic activity of simple and platinum modified tubular SiO2 for the oxidation of oxalic acid to CO2. Catal. Lett. 2009, 132, 81−86. (12) Hobbs, L. W.; Yuan, X.; Qin, L. C.; Pulim, V.; Coventry, A. The nanostructures of amorphous silicas. Microsc. Microanal. 2002, 8, 29−34. (13) Xu, Y.; Lei, B.; Guo, L.; Zhou, W.; Liu, Y. Preparation, characterization and photocatalytic activity of manganese doped TiO2 immobilized on silica gel. J. Hazard. Mater. 2008, 160, 78−82. ( 14 ) Echavia, G. R. M.; Matzusawa, F.; Negishi, N. Photocatalytic degradation of organophosphate and phosphonoglycine pesticides using TiO2 immobilized on silica gel. Chemosphere 2009, 76, 595−600. ( 15 ) Zainudin, N. F.; Abdullah, A. Z.; Mohamed, A. R. Characteristics of supported nano-TiO2/ZSM-5/silica gel (SNTZS): Photocatalytic degradation of phenol. J. Hazard. Mater. 2010, 174, 299−306. (16) Ogata, A.; Kazusaka, A.; Enyo, M. Photoactivation of silica gel with UV light during the reaction of carbon monoxide with oxygen. J. Phys. Chem. 1986, 90, 5201−5205. (17) Fioressi, S.; Arce, R. Photochemical transformations of benzo[e]pyrene in solution and adsorbed on silica gel and alumina surfaces. Environ. Sci. Technol. 2005, 39, 3646−3655. (18) Reyes, C. A.; Medina, M.; Hernandez, C. C.; Cedeno, M. Z.; Arce, R.; Rosario, O.; Steffenson, D. M.; Ivanov, I. N.; Sigman, M. E.; Dabestani, R. Photochemistry of pyrene on unactivated and activated silica surfaces. Environ. Sci. Technol. 2000, 34, 415−421. (19) Barbas, J. T.; Sigman, M. E.; Dabeatani, R. Photochemical oxidation of phenanthrene 20
ACS Paragon Plus Environment
Page 20 of 23
Page 21 of 23
Environmental Science & Technology
sorbed on silica gel. Environ. Sci. Technol. 1996, 30, 1776−1780. (20) Sotero, P.; Arce, R. Surface and adsorbates effects on the photochemistry and photophysics of adsorbed perylene on unactivated silica gel and alumina. J. Photochem. Photobiol. A 2004, 167, 191−199. (21) Mao, Y.; Thomas, J. K. Chemical reactions of molecular oxygen in surface-mediated photolysis of aromatic compounds on silica-based surfaces. J. Phys. Chem. 1995, 99, 2048−2056. (22) Ferreira, L. F. V.; Da Silva, J. P.; Machado, I. F.; Branco, T. J. F.; Moreira, J. C. Surface photochemistry: Dibenzo-p-dioxin adsorbed onto silicalite, cellulose and silica. J. Photochem. Photobiol. A 2007, 186, 254−262. (23) Da Silva, J. P.; Ferreira, L. F. V.; Da Silva, A. M.; Oliveira, A. S. Photochemistry of 4-chlorophenol on cellulose and silica. Environ. Sci. Technol. 2003, 37, 4798−4803. (24) Vallyathan, V.; Shi, X. L.; Dalal, N. S.; Irr, W.; Castranova, V. Generation of free radicals from freshly fractured silica dust. Am. Rev. Respir. Dis. 1988, 138, 1213−1219. (25) Fubini, B.; Hubbard, A. Reactive oxygen species (ROS) and reactive nitrogen species (RNS) generation by silica in inflammation and fibrosis. Free Radical Bio. Med. 2003, 34, 1507−1516. (26) Romanias, M. N.; Andrade-Eiroa, A.; Shahla, R.; Bedjanian, Y.; Zogka, A. G.; Philippidis, A.; Dagaut, P. Photodegradation of pyrene on Al2O3 surfaces: A detailed kinetic and product study. J. Phys. Chem. A 2014, 118, 7007−7016. (27) Hua, I.; Kang, N.; Jafvert, C. T.; Fábrega-Duque, J. R. Heterogeneous photochemical reactions of decabromodiphenyl ether. Environ. Toxicol. Chem. 2003, 22, 798−804. ( 28 ) Ahn, M. Y.; Filley, T. R.; Jafvert, C. T.; Nies, L.; Hua, I.; Bezares-Cruz, J. Photodegradation of decabromodiphenyl ether adsorbed onto clay minerals, metal oxides, and sediment. Environ. Sci. Technol. 2006, 40, 215−220. (29) Qu, R.; Li, C.; Pan, X.; Zeng, X.; Liu, J.; Huang, Q.; Feng, J.; Wang, Z. Solid surface-mediated photochemical transformation of decabromodiphenyl ether (BDE-209) in aqueous solution. Water Res. 2017, 125, 114−122. (30) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, M. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09 (Revision A.02), Gaussian, Inc.: Wallingford, CT, 2009. (31) Söderström, G.; Sellström, U.; de Wit, C. A.; Tysklind, M. Photolytic debromination of decabromodiphenyl ether (BDE 209). Environ. Sci. Technol. 2004, 38, 127−132. (32) Bezares-Cruz, J.; Jafvert, C. T.; Hua, I. Solar decomposition of decabrominated ether: Products and quantum yield. Environ. Sci. Technol. 2004, 18, 358−371. (33) Eriksson, J.; Green, N.; Marsh, G.; Bergman, A. Photochemical decomposition of 15 21
ACS Paragon Plus Environment
Environmental Science & Technology
polybrominated diphenyl ethers in methanol/water. Environ. Sci. Technol. 2004, 38, 3119−3125. ( 34 ) Schenker, U.; Soltermann, F.; Scheringer, M.; Hungerbühler, K. Modeling the environmental fate of polybrominated diphenyl ethers (PBDEs): The importance of photolysis for the formation of lighter PBDEs. Environ. Sci. Technol. 2008, 42, 9244−9249. (35) An, T. C.; Gao, Y. P.; Li, G. Y.; Kamat, P. V.; Peller, J.; Joyce, M. V. Kinetics and mechanism of •OH mediated degradation of dimethyl phthalate in aqueous solution: Experimental and theoretical studies. Environ. Sci. Technol. 2014, 48, 641−648. (36) Qu, R.; Xu, B.; Meng, L.; Wang, L.; Wang, Z. Ozonation of indigo enhanced by carboxylated carbon nanotubes: Performance optimization, degradation products, catalytic mechanism and toxicity evaluation. Water Res. 2015, 68, 316−327. (37) Raff, J. D.; Hites, R. A. Gas-phase reactions of brominated diphenyl ethers with OH radicals. J. Phys. Chem. A 2006, 110, 10783−10792. (38) Zhou, J.; Chen, J. W.; Liang, C. H.; Xie, Q.; Wang, Y. N.; Zhang, S. Y.; Qiao, X. L.; Li, X. H. Quantum chemical investigation on the mechanism and kinetics of PBDE photooxidation by • OH: A case study for BDE-15. Environ. Sci. Technol. 2011, 45, 4839−4845. (39) Wang, D.; Buriak, J. M. Trapping silicon surface-based radicals. Langmuir 2006, 22, 6214−6221. (40) Zhang, H. Y.; Dunphy, D. R.; Jiang, X. M.; Meng, H.; Sun, B. B.; Tarn, D.; Xue, M.; Wang, X.; Lin, S. J.; Ji, Z. X.; Li, R. B.; Garcia, F. L.; Yang, J.; Kirk, M. L.; Xia, T.; Zink, J. I.; Nel, A.; Brinker, C. J. Processing pathway dependence of amorphous silica nanoparticle toxicity: Colloidal vs pyrolytic. J. Am. Chem. Soc. 2012, 134, 15790−15804. (41) Feng, G. D.; Cheng, P.; Yan, W. F.; Boronat, M.; Li, X.; Su, J. H.; Wang, J. Y.; Li, Y.; Corma, A.; Xu, R. R.; Yu, J. H. Accelerated crystallization of zeolites via hydroxyl free radicals. Science 2016, 351, 1188−1191. (42) Narayanasamy, J.; Kubicki, J. D. Mechanism of hydroxyl radical generation from a silica surface: Molecular orbital calculations. J. Phys. Chem. B 2005, 109, 21796−21807. (43) Satheesh, S. K.; Moorthy, K. K. Radiative Effects of Natural Aerosols: A Review. Atmos. Environ. 2005, 39, 2089−2110.
22
ACS Paragon Plus Environment
Page 22 of 23
Page 23 of 23
Environmental Science & Technology
TOC art
23
ACS Paragon Plus Environment