Subscriber access provided by Bethel University
Remediation and Control Technologies
Complete defluorination and mineralization of perfluorooctanoic acid by mechanochemical method using alumina and persulfate Nan Wang, Hanqing Lv, Yuqi Zhou, Lihua Zhu, Yue Hu, Tetsuro Majima, and Heqing Tang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.9b00486 • Publication Date (Web): 31 May 2019 Downloaded from http://pubs.acs.org on June 4, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 37
Environmental Science & Technology
1
Complete defluorinationand mineralization of perfluorooctanoic acid by a
2
mechanochemical method using alumina and persulfate
3
Nan Wanga,+, Hanqing Lva,+, Yuqi Zhou a, Lihua Zhua,*,Yue Hub, Tetsuro Majima a, Heqing
4
Tangb,*
5
a
6
Technology, Wuhan 430074, P. R. China
7
bCollege
8
430074, P. R. China
9
+These
College of Chemistry and Chemical Engineering, Huazhong University of Science & of Resourcesand Environmental, South-Central University for Nationalities, Wuhan
authors contributed equally to this work.
10
ACS Paragon Plus Environment
1
Environmental Science & Technology
12
ABSTRACT: Perfluorooctanoic acid (PFOA) is a persistent organic pollutant that has
13
received concerns worldwide due to its extreme resistance to conventional degradation. A
14
mechanochemical (MC) method was developed for complete degradation of PFOA by using
15
alumina (Al2O3) and potassium persulfate (PS) as comilling agents. After ball milling for 2 h,
16
the MC treatment using Al2O3 or PS caused conversion of PFOA to either
17
1-H-1-perfluoroheptene or dimers with a defluorination efficiency lower than 20%, but that
18
using both Al2O3 and PS caused degradation of PFOA with a defluorination of 100% and a
19
mineralization of 98%. This method also caused complete defluorination of other C3~C6
20
homologues of PFOA. The complete defluorination of PFOA attributes to Al2O3 and PS led to
21
the weakening of the C-F bond in PFOA and the generation of hydroxyl radical (•OH),
22
respectively. During the MC degradation, Al2O3 strongly anchors PFOA through COO−-Al
23
coordination and in-situ formed from Lewis-base interaction and PS through hydrogen bond.
24
Meanwhile, mechanical effects induce the homolytic cleavage of PS to produce SO4•−, which
25
reacts with OH group of Al2O3 to generate •OH. The degradation of PFOA is initiated by
26
decarboxylation as a result of weakened C-COO− due to Al3+ coordination. The subsequent
27
addition of •OH, elimination of HF, and reaction with water induce the stepwise removal of
28
all carboxyl groups and F atoms as CO2 and F−, respectively. Thus, complete defluorination
29
and mineralization are achieved.
30
ACS Paragon Plus Environment
2
Page 2 of 37
Page 3 of 37
32
Environmental Science & Technology
TOC Art
33
ACS Paragon Plus Environment
3
Environmental Science & Technology
35
Page 4 of 37
INTRODUCTION
36
Perfluorooctanoic acid (C7F15COOH, PFOA), perfluorooctane sulfonate (C7F15SO3H,
37
PFOS), and other fluorosurfactants have been widely applied in textiles, paper and packaging
38
materials, cookware, and fire-fighting foam productions. Due to their disposal, these
39
fluorocarbons are frequently detected in aquatic environments,1 soils, sludge and sediments.2
40
Because PFOA and PFOS are persistent organic pollutants (POPs),3 their elimination is
41
necessary.
42
PFOA is thermally and chemically stable because of high dissociation energy of C−F
43
bonds (533 kJmol−1).4 To induce the degradation of PFOA in aqueous media, several
44
chemical approaches have been explored, such as UV-sulfite reduction,5,
45
electro-Fenton oxidation,7,8 photocatalytic oxidation,9 and sulfate radical anion (SO4•−)-based
46
oxidation.10-12 However, few methods have been developed for the degradation of PFOA in
47
solids. Incineration at temperatures above 500 °C is a practical method to treat organic
48
pollutants in solid, although the thermal treatment of PFOA produces undesirable greenhouse
49
gases and corrosive gases such as CF4, C2F4, and HF.13 Thus, development of complete and
50
environment-friendly disposal technologies is urgently needed for treating PFOA in solids.
6
photo- and
51
Mechanochemical (MC) degradation is a promising method for the disposal of organic
52
pollutants in solid. In an MC treatment, the degradation reaction usually begins with the MC
53
activation of milling agents by particle crushing, build-up of structural defects, rupture of
54
bonds, heating, etc., under intensive mechanical stresses. The efficiency of the MC
55
degradation of organic pollutants is improved by using appropriate additives. For example,
56
CaO14, CaO-SiO215, Fe16, zero-valent metal (e.g., Fe, Al)-SiO217, persulfate (PS)18, and
57
PS-CaO19 have been used as efficient additives in MC degradation of chlorinated and
58
brominated organic pollutants, although these additives are not efficient enough individually
59
for the MC degradation of PFOA (Table S1). Yan et al. investigated the MC degradation of a
ACS Paragon Plus Environment
4
Page 5 of 37
Environmental Science & Technology
60
chlorinated polyfluorinated ether sulfonate (F-53B, a PFOS alternative) by using sodium
61
persulfate and NaOH as comilling agents, and found that the defluorination efficiency of
62
F53-B was 54% after ball milling for 8 h.20 Zhang et al. reported that ball milling for 8 h
63
produced defluorination efficiency of PFOA and F-53B greater than 80% by adding KOH as a
64
milling agent with a KOH/PFOA molar ratio of 170:1, but less than 20% by adding NaOH.21,
65
22
66
The relatively poor defluorination in the MC degradation of PFOA is closely related to
67
the chemical inertness of C−F bonds. Activation of a C−F bond and the following cleavage
68
involve the formation of thermodynamically favorable bonds such as H−F, B−F, Al−F, Si−F,
69
and Ge−F bonds.23, 24 Recently, we have found that the mechanocaloric effect promotes the
70
dehydration of the Al2O3 surface to produce coordinatively unsaturated Lewis acid sites,
71
inducing the activation of C−F bonds to convert PFOA to a valuable product,
72
1-H-1-perfluoroheptene.25 Cagnetta et al. reported that the use of La2O3 as a milling agent
73
converted perfluorinated compounds into LaOF, a luminescent material.26 This MC method is
74
preferable for the degradation of solid PFOA, because not only is PFOA degraded, but
75
organo-fluorine blocks in PFOA are also recovered for synthesizing polyfluoroalkenes or
76
LaOF. However, MC treatment is not appropriate for degradation of PFOA at ratios lower
77
than 20% in solids because the economy of this method becomes poor for recovery of the
78
fluorine resource. In most cases, the PFOA content is too low to be recovered as a resource.
79
Therefore, a degradation of PFOA with complete defluorination in solids to harmless
80
substances is desirable.
81
We recently have reported using the MC method with PS to cause the homolytic
82
cleavage of the peroxide bond in PS to produce SO4•−, inducing the decomposition of
83
brominated organic pollutants.18 Considering the activation effect of Al2O3 on C−F bonds of
84
PFOA and generation of SO4•− from the dissociation of PS by the MC treatment, we
ACS Paragon Plus Environment
5
Environmental Science & Technology
85
anticipated that Al2O3 and PS might lead to complete MC degradation and defluorination of
86
PFOA. In the present work, we developed an MC method for degradation of PFOA by using
87
both Al2O3 and PS as comilling agents, and clarified the reaction mechanism.
88
89
Chemicals and Reagents. Perfluorooctanoic acid (PFOA, C7F15COOH, 98% in purity) was
90
purchased from J&K China Chemical, Beijing, China. Shorter chain perfluorocarboxylic
91
acids (PFCAs) including perfluoroheptanoic acid (PFHpA, >98%), perfluorohexanoic acid
92
(PFHxA, >98%) and perfluoropentanoic acid (PFPA,>98%) were purchased from Tokyo
93
Chemical Industry, Shanghai, China. Alumina (Al2O3, 99.9%), 5,5-dimethyl-1-pyrroline
94
(DMPO, analytical grade), and diphenyl-1-picrylhydrazyl (DPPH•, 97%) was obtained from
95
Aladdin, Shanghai, China. H218O with an isotope abundance of 98% was purchased from
96
Jiangsu Huayi Isotopes Chemical Limited. Potassium persulfate (PS, analytical grade) was
97
obtained from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. All solvents used for
98
extraction and high-performance liquid chromatography (HPLC) analysis were analytical
99
grade and HPLC grade, respectively. Milli-Q water with conductivity of 18.2 MΩ cm (Merck
100
EXPERIMENTAL SECTION
Millipore GmbH, Germany) was used in the present work.
101
Degradation Experiments. The MC degradation of PFOA was conducted in a planetary
102
ball mill at room temperature (BM4, Beijing Grinder Instrument, China) by using Al2O3
103
and/or PS as milling agents. Typically, 0.25 g PFOA and 3.95 g milling agents (PS, Al2O3,
104
and/or their mixtures) were mixed in the stainless steel pot (250 mL) of the ball milling
105
machine, which was then filled with 18 stainless steel balls with a diameter (d) of 6 mm and
106
10 balls with d = 10 mm (the total mass of balls was 210 g). When PS and Al2O3 were used as
107
comilling agents, the molar ratio of Al2O3/PS ( nAl O /nPS ) was set at 5 unless otherwise stated.
108
The mass of the total milled sample was 4.2 g, and the ball-to-sample mass ratio was fixed at
109
50:1. After the milling materials were well predistributed in a pot, the pot was sealed tightly
2
3
ACS Paragon Plus Environment
6
Page 6 of 37
Page 7 of 37
Environmental Science & Technology
110
and the planetary ball mill was operated at 350 rpm under atmospheric conditions, with
111
automatic change in the direction of rotation every 15 min. At given time intervals, the ball
112
milling was stopped and solid samples were taken out for chemical analysis, unless specified
113
elsewhere. Replicate runs (n ≥ 3) were carried out for each test, and the relative
114
standard deviations were less than 5%.
115
For isotopic labeling experiments, both the PS and Al2O3 were preheated at 100 °C for
116
24 h and kept in a desiccator prior to MC degradation experiments. H218O (1 mL) was mixed
117
with Al2O3 (10 g) in a desiccator for 1 h to ensure full oxygen exchange between H218O and
118
surface OH on the Al2O3, followed by vacuum drying at 60 °C for 1 h. Then,
119
Al2O3 powders were used for the MC treatment of PFOA.
120
Analysis. Each milled sample (0.020g) was extracted with ethanol (3 times, for a total of 10
121
mL) under ultrasound irradiation (KQ-200KDE, Kunshan Ultrasonic Instruments, China) for
122
10min. After removing the solids by centrifugation, the collected solution was filtered through
123
a 0.22 μm membrane and then subjected to HPLC analysis. Similar to our previous work,27
124
the PFOA concentration was measured with HPLC on an Ultimate 3000 series system
125
(Dionex, Idstein, Germany) equipped with a Corona Ultra RS charged aerosol detector (CAD,
126
Thermo Scientific, Bellefonte, PA, USA). The analysis was performed on a Symmetry C8
127
column (150 mm × 3.9 mm; 5 μm particle size; Waters, Ireland). Column oven temperature
128
was set at 30 °C. The mobile phase was 30% ammonium acetate (5 mmol L−1 with pH
129
adjusted to 5.0 ± 0.2 by acetic acid) and 70% methanol by volume, the flow rate was 1 mL
130
min−1, and the injection volume was 40 μL. The degradation efficiency of PFOA (DGEPFOA)
131
is calculated from eq. 1,
132
133
n DGE PFOA 1 PFOA,t 100% nPFOA,0
18O-labeled
(1)
where nPFOA,0 and nPFOA,t are moles of PFOA at reaction times 0 and t, respectively. In
ACS Paragon Plus Environment
7
Environmental Science & Technology
Page 8 of 37
134
addition, the ethanol-extracted solutions were analyzed using an 1100 LC/MSD trap
135
system (Agilent, USA) with an electrospray ionization (ESI) source to identify possible polar
136
intermediates.
137
To identify volatilized organic intermediates, both products in the reactor and gaseous
138
products in the head space of the reactor were collected by using acetone as a solvent. The
139
extract was filtered through a 0.22 μm membrane and then directly analyzed by GC/MS
140
(Thermo Fisher, USA) under the same conditions reported in our previous work.25 For the
141
analysis of the CO2 produced, the gases were collected in 20 mL NaOH (0.5 mol L−1) after the
142
ball milling. The CO2 absorbed in NaOH was converted into CO32− and then detected by ion
143
chromatography (IC) on a Dionex ICS-1500 equipped with a CD 25 conductivity detector,
144
IonPac AS 23 column and KOH as the eluent (28 mmol L−1, 1.0 mL min−1). In the isotopic
145
labeling experiments, the produced CO2 was analyzed using Fourier transform infrared
146
(FT-IR) spectrometry (Bruker, Germany) by introducing the gaseous products into the gas
147
cell. The distribution of 18O in CO2 was analyzed not only by GC/MS (Agilent 7890B-5977A
148
GC/MSD) with a DB-5MS column (30 m × 0.25 mm; 0.25 μm particle size) and quadruple
149
mass analyzer, but also by isotope ratio mass spectrometry after converting the gaseous CO2
150
to a precipitate of MnCO3. After the MC reaction, the generated CO2 was absorbed by a
151
NaOH solution, followed by isolating the milled powders using centrifugation and filtration.
152
The filtrate, with a pH approximately 7.5, was adjusted to pH 11 with NaOH (0.5 mol L−1),
153
followed by adding MnSO4•H2O (0.08 g) to yield a MnCO3 precipitate. The
154
abundance in MnCO3 was analyzed by using an elemental analyzer (EA IsoLink CN/OH,
155
Thermo Fisher USA) coupled to an isotope ratio mass spectrometer (Delta V Advantage,
156
Thermo Fisher). In the reaction furnace of the elemental analyzer at 1380 °C, the evolved
157
oxygen in MnCO3 reacted with carbon to yield CO, which was separated and then transported
158
into an isotope ratio mass spectrometer. Because the conversion of MnCO3 to CO strongly
ACS Paragon Plus Environment
8
18O
isotope
Page 9 of 37
Environmental Science & Technology
159
depends on temperature, two reference compounds including CO and MnCO3 that obtained
160
from the reaction of Na2CO3 with MnSO4 (named ref-MnCO3) were subjected to the same
161
measurement conditions.
162
To determine the amount of F− released from PFOA, the milled samples (20 mg) were
163
dispersed in 10 mL water and 0.5 mL aqueous ammonia. The suspensions were centrifuged
164
and filtered through a 0.22 μm membrane. The filtrate then was diluted to 50 mL with 10 mL
165
total ionic strength adjustment buffer (TISAB, consisting of HAc-NaAc and 10 g L−1 sodium
166
citrate) and water. Finally, F− was measured with a F− selective electrode. The recovery yields
167
of F− were greater than 95% as determined using the standard addition method by spiking NaF
168
(0.3 g) into 3.95 g Al2O3 or a mixture of PFOA (0.25 g), K2SO4 (1.57 g), and Al2O3 (2.58 g)
169
(Table S2). The defluorination efficiency of PFOA (DFEPFOA) is evaluated from eq. 2,
170
DFE PFOA
ndetected F ,t nF,0
100%
ndetected F ,t nPFOA,0 15
100%
(2)
171
where ndetected F ,t and nF,0 are moles of F− detected experimentally and F calculated
172
from the initial PFOA (mol), respectively.
173
The
residual
PS
in
the
milled
samples
was
determined
with
a
KI
174
spectrophotometric method.28 Prior to the measurement, the milled samples (20 mg) were
175
dispersed in 10 mL water under ultrasound (KQ-200KDE, Kunshan Ultrasonic Instruments,
176
China) for 10 min. After centrifugation and filtration through a 0.22 μm membrane, 1.5 mL
177
filtrate was mixed with 1.0 mL phosphate buffer (pH 6.86) and 0.5 mL KI (2 M). After 15
178
min, 0.20 mL of this mixture was diluted to 10 mL with water and then measured at 355 nm
179
on a UV-vis spectrophotometer (Cary 50, Varian, USA). The recovery yields of PS were
180
greater than 98% with the standard addition method by spiking PS (1.37 g) into a mixture of
181
PFOA (0.25 g) and Al2O3 (2.58 g). The utilization efficiency of PS (PS) is defined as the ratio
182
of the stoichiometric requirement to the actual consumption, which is calculated from the
ACS Paragon Plus Environment
9
Environmental Science & Technology
Page 10 of 37
183
moles of generated F− and consumed PS ( ndetected F ,t and nPS, 0 ndetected PS, t , respectively) (eq.
184
3),
PS
185
ndetected F
,t
nPS, 0 ndetected PS,t
100%
(3)
186
where ndetected PS, t
and nPS,0 are moles of PS detected experimentally and calculated
187
from the initial PS, respectively, and is the stoichiometric ratio of the moles of
188
generated F− to the moles of consumed PS.
189
When necessary, solid samples were taken at specified time intervals during the ball
190
milling and then were characterized using spectroscopic techniques. FT-IR spectra were
191
measured on an Equinox 55 (Bruker, Germany) with the KBr disk method from 400 cm−1 to
192
4000 cm−1. X-ray photoelectron spectrometry (XPS) was conducted on Kratos AXIS-ULTRA
193
DLD-600W (Shimadzu, Japan), in which the binding energy is calibrated using the C 1s peak
194
as 284.6 eV. 19F and 27Al MAS NMR experiments were performed at room temperature on a
195
Bruker Avance III 500 MHz spectrometer (Bruker, Germany) equipped with a 2.5 mm Bruker
196
1H/19F/X
197
kHz.
probe. The samples were loaded into a zirconia rotor and spun at an MAS rate of 20
198
A Bruker EMX-nano electron spin resonance (ESR) instrument was employed to detect
199
radicals with DMPO as the spin trapping reagent. Prior to measurements, a small amount of
200
DMPO instead of PFOA was added to the milling mixture. After being milled for 15 min at
201
300 rpm, a portion of the milled sample (ca. 0.1 g) was directly collected and another portion
202
(20 mg) of the milled mixture was dispersed in 10 mL water. Both the solid and
203
water-extracted samples were measured under the following conditions: center field, 3480 G;
204
sweep width, 200 G; microwave frequency, 9.77 GHz; modulation frequency, 40 kHz; power,
205
10 mW; scan times, 3.
206
RESULTS ANDDISCUSSION
ACS Paragon Plus Environment
10
Page 11 of 37
Environmental Science & Technology
207
Complete Degradation and Defluorination of PFOA using Mechanochemistry. Figures 1a
208
and 1b illustrate the MC degradation and defluorination of PFOA (0.1 g) in kaolin soil (1.5 g)
209
with PS, Al2O3 or a mixture of PS and Al2O3 as milling agents (5.7 g). After milling for 2 h,
210
all the added PFOA in the three cases was nearly degraded, but defluorination efficiencies
211
calculated from the generated F− were 10%, 27% and 85% in MC-Al2O3, MC-PS and
212
MC-Al2O3-PS, respectively. When Al2O3 (2.4 g) and PS (3.3 g) were divided into three equal
213
portions and added for the three reaction times of 0, 30, and 60 min, the degradation and
214
defluorination extents of PFOA achieved 99% and 96% at 2 h, respectively. This indicates
215
that the use of both Al2O3 and PS favors the MC defluorination of PFOA. Because the soil
216
matrix will bring more difficulties in clarifying the roles of Al2O3 and PS, the subsequent
217
degradation experiments were carried out using pure PFOA as a model pollutant.
Degradation efficiency /%
75
50
PS Al2O3+PS
25
0
218
Al2O3
Al2O3+PS(1/33)
0
40 80 Milling time /min
Defluorination efficiency /%
100 (b)
100 (a)
75 Al2O3 PS Al2O3+PS
50
Al2O3+PS(1/33) 25
0
120
0
30
60 90 Milling time /min
120
219
Figure 1. Time profiles of degradation (a) and defluorination (b) of PFOA (0.1 g) in kaolin
220
soil (1.5 g) under 350 rpm and a ball-to-sample mass ratio (mb/ms) of 50 in the presence of
221
Al2O3 (5.7 g), PS (5.7 g), or the mixture of Al2O3 (2.4 g) and PS (3.3 g). All milling agents
222
were added in one time before milling, except for the system of Al2O3+PS (1/3 3), in which
223
PS and Al2O3 were divided into three equal portions and added at reaction time of 0, 30, and
224
60 min.
225
Figures 2a and 2b compare the MC degradation and defluorination of powdered PFOA
ACS Paragon Plus Environment
11
Environmental Science & Technology
Page 12 of 37
226
(0.25 g) by using 3.95 g milling agents. After a 2 h milling with either PS ( nPS /nPFOA = 24) or
227
Al2O3 ( nAl O /nPFOA = 64), the degradation efficiency of PFOA was greater than 60%, but the 2
3
228
defluorination extent was lower than 20%. When using comilling agents at nAl O /nPS = 5,
229
both degradation and defluorination efficiencies reached 100% after 2 h milling. Moreover, at
230
any time, the defluorination extent of PFOA was close to the degradation in MC-Al2O3-PS,
231
thus confirming that successive degradation and defluorination continued until there is a
232
complete removal of F atoms. In addition, the fate of fluorine was identified by solid-state
233
nuclear magnetic resonance (SS NMR) and XPS measurements. High-resolution XPS spectra
234
of F 1s and NMR spectra of
235
MC-Al2O3-PS is bonded to the Al2O3 surface to form Al−F bond (Text S1, Figures S1 and
236
S2). This provides additional merit for reducing the secondary pollution because the bonding
237
of F− on Al2O3 reduces the dissociation of F− into water and thus avoids the possible
238
promotion effect of bone disease from excess F−.
2
19F
27Al
and
3
demonstrated that F− eliminated from PFOA in
239
Figure 2c shows the effect of nAl O /nPS on the degradation and defluorination ratios of
240
PFOA after a 2 h MC treatment. As nAl O /nPS increases from 0 to 2, the degradation extent
241
of PFOA increases from 61% to 96%. A further increase in nAl O /nPS to 5 results in a
242
degradation ratio for PFOA of 100%. However, a different volcano plot is observed for the
243
effect of nAl O /nPS on the defluorination efficiency of PFOA. The defluorination ratio of
244
PFOA increases from 6.2% to 98% when nAl O /nPS increases from 0 to 3 and reaches 100%
245
at nAl O /nPS = 5. After that, the defluorination ratio of PFOA significantly decreases to 42% as
246
nAl O /nPS increases to 10. The use of Al2O3 alone leads to a defluorination ratio of 18%, even
247
though all the added PFOA is removed. The above observations are related to different roles
248
of comilling agents on the degradation and defluorination of PFOA. Al2O3 exhibits better
2
3
2
3
2
2
3
2
2
2
3
3
3
3
ACS Paragon Plus Environment
12
Page 13 of 37
Environmental Science & Technology
249
performance than PS on the degradation of PFOA via a decarboxylation process as a result of
250
weakened C-COO− due to Al3+ coordination (Figure 2a), leading to that a larger nAl O /nPS
251
caused a higher degradation ratio of PFOA. However, neither the in situ generated electrons
252
in MC-Al2O3 nor the generated SO4•− in MC-PS was efficient in the defluorination of PFOA
253
(Figure 2b). The complete defluorination of PFOA is attributed to Al2O3 and PS, which
254
caused the weakening of the C-F bond in PFOA and the generation of hydroxyl radicals
255
(•OH), respectively. During ball milling, mechanical effects induce the decomposition of PS
256
into SO4•−, which reacts with surface OH− (OH−surf) on Al2O3 to generate •OHsurf. As
257
nAl O /nPS decreases, the total amount of OH−surf on Al2O3 decreases. In contrast, as nAl O /nPS
258
increases, the amount of PS decreases. Either of them would result in the insufficient
259
generation of •OHsurf from the reaction of SO4•− with OH−surf. Thus, a complete defluorination
260
of PFOA requires a moderate value of nAl O /nPS .
2
2
2
3
2
3
3
3
261
Figure 2d illustrates the effect of the ball-to-sample mass ratio (mb/ms) on the MC
262
degradation of PFOA at a fixed nAl O /nPS /nPFOA of 42: 8.4: 1. As mb/ms increases from 15 to
263
50, both the degradation and defluorination ratios of PFOA at 2 h increases from
264
approximaterly 35% to 100%. However, slight increases in the degradation and defluorination
265
ratios of PFOA are observed with an increase in mb/ms to 75 or more. Larger mb/ms values can
266
provide higher mechanical energy or a stronger mechanocaloric effect, which may induce an
267
increase in temperature. However, as mb/ms varied from 25 to 100, the change in temperature
268
on the surface of the milled samples was almost the same and the maximum temperature was
269
below 50 °C (Figure 2e), due to the instantaneous temperature produced by ball milling
270
having a short lifetime of 2–10 ns.29 After the ball milling, both PS and Al2O3 were not
271
molten (Figure S3), because the surrounding temperature was much lower than their melting
272
points. In a control test, when the powdery mixtures of PFOA, Al2O3 and PS were heated at
2
3
ACS Paragon Plus Environment
13
Environmental Science & Technology
Page 14 of 37
273
100 °C for 2 h, the defluorination efficiency of PFOA was negligible (Figure 2f). Thus, the
274
activation of solid-to-solid reactions by ball milling is not a simple heat activation mechanism.
275
The high energy impact between surfaces of the balls not only produces a transient high
276
temperature but also initiates the rotation, alignment and reorientation of molecules, all of
277
which facilitate changes in the solid surface and chemical bonds. A larger mb/ms can increase
278
the total number of impacts between balls to achieve reaction completion for a given milling
279
time, and any further increase of mb/ms can only act to accelerate to accelerate reaction rate.30,
280
31
(a)
75
50
25
Al2O3-PS Al2O3
(c)
only Al2O3
75
25
25
0
2
4
6
nAl O /nPS 2
(e)
8
10
0
30
100
15
30
60 90 Milling time /min
100
50
50
25
25
0
15
3
25 50 75 100
(d)
120
75
0
mb/ms
60 90 Milling time /min
75
100
30
0
0
100
45
T / C
25
120
50
0
PS
50
45
(f)
mb/ms
Defluorination efficiency /%
60 90 Milling time /min
50
60
Al2O3
0
75
105 100
PFOA-PS
75
PFOA-Al2O3-PS
50
50
25
0
30
ACS Paragon Plus Environment
14
25
PFOA-Al2O3-PS PFOA-PS
0
120
75
60 Time /min
90
0 120
Residual PS /%
30
75
0
Al2O3-PS
75
Defluorination efficiency /%
Degradation efficiency /%
100
(b)
PS
0
Defluorination efficiency /%
0
281
Defluorination efficiency /%
100
Degradation efficiency /%
Degradation efficiency /%
100
Page 15 of 37
Environmental Science & Technology
282
Figure 2. Time profiles of degradation (a) and defluorination (b) of PFOA (0.25 g) under 350
283
rpm and mb/ms of 50 with milling agents such as PS (3.95 g), Al2O3 (3.95 g), or the mixture of
284
Al2O3 (2.58 g) and PS (1.37 g) at nAl O /nPS of 5. Effects of nAl O /nPS (c) and mb/ms (d) on
285
degradation (triangles) and defluorination ratios (circles) of PFOA in MC-Al2O3-PS at a 2 h
286
milling. (e) Time profiles of the temperature on the surface of milled samples during the
287
milling. (f) Time profiles of defluorination efficiency of PFOA with PS or the mixture of PS
288
and Al2O3 in solid state by heating to 100 °C.
289
Mineralization of Degradation Intermediates. The degradation intermediates produced
290
during the 1 h MC degradation of PFOA were extracted with ethanol or acetone as described
291
in the Experimental Section. The organic intermediates in the ethanol-extracted solutions
292
were identified by LC/MS, and those in the acetone-extracted solutions were identified by
293
GC/MS.
2
2
3
3
294
When Al2O3 was used as a milling agent, three products were detected in the acetone
295
extracts from milled samples, where CF3(CF2)4CF=CFH was observed as a major product in a
296
yield of 87% with the almost constant selectivity at 88% after a 2 h milling (Figures 3a, S4a,
297
and S5). This indicates that the transformation to CF3(CF2)4CF=CFH was the main
298
degradation pathway of PFOA in MC-Al2O3. When using PS as a milling agent, large
299
amounts of fluoroketones, fluoroethers, and fluoroesters in the acetone-extracted solutions
300
was observed as well as trace amounts of shorter chain PFCAs (such as perfluorohexanoic
301
acid (C5F11COOH, PFHxA) and perfluoroheptanoic acid (C6F13COOH, PFHpA) in
302
ethanol-extracted solutions (Text S1, Figures S4b, S6, and S7). Due to the lack of authentic
303
compounds, we did not quantify the individual intermediates, but measured the total organic
304
carbon (TOC) in the milled sample. As shown in Figure 3b, the TOC resulting from the
305
generated organic products and residual PFOA was 89%, although 61% of the added PFOA
ACS Paragon Plus Environment
15
Environmental Science & Technology
306
was degraded after a 2 h milling.
307
When both Al2O3 and PS were used as comilling agents, two components were detected
308
in the ethanol-extracted solutions (Figure S8). The first component was detected at 3.9 min
309
with m/z = 314.9(C5F11COO−), and 272.1(C5F11−) was identified as PFHxA (C5F11COOH),
310
and the second component was detected at 5.6 min with m/z = 362.8(C6F13COO−) and
311
319.7(C6F13−) was identified as PFHpA (C6F13COOH). Four main polyfluorocarbons (PFCs)
312
were detected in the acetone-extracted solutions (Figure S4a). The component that was
313
detected at 1.40 min with the mass fragments of 219.12 (CF3(CF2)3), 169.17 (CF3(CF2)2),
314
119.15 (CF3CF2) and 69.2 (CF3) was assigned to CF3(CF2)2CF2H (Figure S9a). Similarly, the
315
other three components were identified as CF3(CF2)3CF=CF2 at 1.46 min, CF3(CF2)4CF=CFH
316
at 1.50 min, and CF3(CF2)5CF2H at 1.60 min, respectively (Figures S9b-d). The maximum
317
ratio of total organic intermediates was less than 10% of the initial PFOA level, and their
318
accumulation gradually decreased to zero as the ball milling time was prolonged to 2 h
319
(Figure 3c). These comparisons confirm that using both Al2O3 and PS not only alters the
320
pathway of PFOA degradation but also significantly reduces the accumulation of
321
intermediates.
322
In addition, CO2 yields were measured in all three tested mixtures when the gaseous
323
products were collected in NaOH solution. After a 2 h milling, the ratios of CO2 to the initial
324
amount of carbon in PFOA were 18, 7.9, and 98% for MC-Al2O3, MC-PS, and MC-PS-Al2O3,
325
respectively (Figures 3a~c). Throughout the MC treatment, the total carbon included in the
326
residual PFOA, organic intermediates and produced CO2 was almost equal to the initial value
327
for the three tested reactions. Thus, it is emphasized that PFOA was converted to either
328
CF3(CF2)4CF=CFH in MC-Al2O3, or dimeric molecules in MC-PS, while the degradation to
329
CO2 was dominant in MC-PS-Al2O3. Moreover, the TOC removal efficiency of PFOA in
330
MC-PS-Al2O3 was 98%, being 5.4 and 8.9 times higher than those of MC-Al2O3 (18%) and
ACS Paragon Plus Environment
16
Page 16 of 37
Page 17 of 37
331
Environmental Science & Technology
MC-PS (11%), respectively. 100
(a)
100
(b)
PFOA
C5F11CF=CFH
50
25
CO2
F2 =C CF
100
0
40 80 Milling time /min
Carbon distribution /%
0
332
CO2
50 PFHpA PFHxA PFCs
25
0
25
CO2 0 0 100
TC
75
PFOA
50
120
(c) PFOA
75
C3F7CF=CFH
F 1 C5 1
0
Carbon distribution /%
75
40 80 Milling time /min
TC
TOC
TOC removal /%
Carbon distribution /%
total carbon (TC)
40 80 Milling time /min
120
(d)
75 PS Al2O3
50
Al2O3-PS 25
0
120
0
40 80 Milling time /min
120
333
Figure 3 Distributions and balance of carbon during MC degradation of PFOA in (a)
334
MC-Al2O3, (b) MC-PS and (c) MC-Al2O3-PS. (d) Removal efficiency of total organic carbon
335
(TOC) during MC degradation of PFOA. The measured TOC in (b) includes that in the
336
generated organic products (such as fluoroketones, fluoroethers, and fluoroesters) and residual
337
PFOA.
338
Interaction of PFOA with Comilling Agents and the Transformations. After mixing
339
PFOA with PS and/or Al2O3 by manually grinding in an agate mortar, a small portion of the
340
sample was compressed into a KBr pellet for the FT-IR measurement. Table 1 summarizes
341
the dominant absorption peaks together with their assignments. In FT-IR spectrum of
342
PFOA-PS, a new and weak peak appeared at 1723 cm−1 in addition to the strong absorption
ACS Paragon Plus Environment
17
Environmental Science & Technology
343
peak for the carboxyl group of PFOA (Table 1, Figure S10a). Meanwhile, the symmetric
344
stretching (νs) peak of SO3− at 1108 cm−1 significantly decreased in intensity. This indicates
345
that most PFOA molecules exist in the form of carboxylic acid, while a small amount of
346
PFOA is bound to PS through the intermolecular hydrogen bond between the SO3− of PS and
347
the COOH of PFOA.32 For PFOA-Al2O3, the stretching peak of C=O (1768 cm−1) in PFOA
348
disappeared, and νs and the asymmetric stretching peak (νas) of COO− appeared at 1413 and
349
1676 cm−1, respectively, with Δν (= νas−νs) of 263 cm−1. The stretching peaks of the C−F units
350
moved to higher wavenumbers by 4 ~ 7 cm−1 (Table 1, Figure S10b). These changes are
351
related to both the COO−Al coordination in a monodentate mode through the neutralization
352
between COOH and surface OH on Al2O3 (>AlOH) (eq. 4) and hydrogen bond between the F
353
of PFOA and >AlOH (eq. 5).25,33,34 After mixing the three components, the spectral changes
354
related to the carboxyl group and C−F units were similar to those in PFOA-Al2O3 but different
355
from those in PFOA-PS (Table 1, Figure S10c). This indicates that PFOA is bound to the
356
Al2O3 surface rather than that of PS. The deprotonation of PFOA causes the dissociation of
357
hydrogen bonding between PFOA and PS, leading to the recovery of the νs peak of SO3− at
358
1108 cm−1. Unlike the expectation, the weak peak at 1108 cm−1 in PFOA-Al2O3-PS was
359
observed to indicate the presence of other hydrogen bond donors. After mixing PS and Al2O3,
360
the νs peak of SO3− at 1108 cm−1 weakened (Table 1, Figure S10d), implying that the surface
361
OH of Al2O3 and the SO3− of PS act as hydrogen bond donors and acceptors, respectively (eq.
362
6). Thus, it is concluded that Al2O3 anchors the PFOA and PS in ternary mixtures and the two
363
interaction modes including PFOA-Al2O3 and Al2O3-PS do not influence each other. The
364
characteristic absorption peaks of PFOA and PS in MC-Al2O3-PS gradually decreased as the
365
ball milling proceeded. Moreover, the rates of decrease were faster than those for MC-PS
366
(Figure S11), which suggests that the coexistence of Al2O3 promotes the conversion of PFOA
367
and PS.
ACS Paragon Plus Environment
18
Page 18 of 37
Page 19 of 37
Environmental Science & Technology
368
AlOH C 7 F15 COOH Al 3
369
AlOH Al 3
370
AlOH OS(O 2 )OOSO 3 AlOH
371
OOCC 7 F15 H 2 O
OOCC 7 F15 Al 3
(4)
OOCC 7 F15 HOAl
OS(O 2 )OOSO 3
(5) (6)
Table 1. Characteristic FT-IR peaks of PFOA, PS, Al2O3, and their mixtures before milling. Samples
Absorption peaka /cm−1 ν(C=O/COO−)
ν(CF)
νas(SO3−)
νs(SO3−)
PFOA
1768
1236, 1204, 1147
−
−
PS
−
−
1303, 1276
1108
Al2O3
−
−
−
−
PFOA-PS
1768, 1722(w)
1236, 1204, 1147
1303, 1276
1108(w)
PFOA-Al2O3
1676, 1413 (w)
1243, 1211, 1151
−
−
PS-Al2O3
−
−
1303, 1276
1108(w)
PFOA-Al2O3-PS
1681, 1415 (w)
1243, 1211, 1151
1303, 1276
1108(w)
372
aThe
373
XPS measurements were conducted to check the elemental changes of Al and S in
374
MC-Al2O3-PS. Before the ball milling, the Al 2p peak approximaterly 74.8 eV was
375
deconvoluted into two peaks at 73.9 and 75.0 eV corresponding to Al−O and Al−OH in Al2O3,
376
respectively (Figure 4a).35 After ball milling for 15 min, the Al−O peak gradually decreased
377
in intensity and a new peak appeared at 76.4 eV, which was assigned to F– bonded to Al
378
(F−Al).36 With an increase in the milling time to 60 min, the Al−O peak became weaker and
379
the intensity of F−Al peak increased. This indicated that more F atoms were eliminated from
380
the PFOA and then bonded to the Al2O3 surface. The sample before milling showed an S 2p
381
peak at 170.3 eV, which was assigned to S2O82− (Figure 4b).37 After ball milling for 1 h, the S
382
2p peak gradually moved to 169.3 eV to be in the proximity of the binding energy of
383
inorganic sulfates.38 This supports the hypothesis the conversion of S2O82– to SO42– took place
384
during the MC degradation of PFOA.
letter “w” in parenthesis represents the weak absorption peak.
ACS Paragon Plus Environment
19
Environmental Science & Technology
(a)
(b) Al-OH
Al-O
Al-F
81
Page 20 of 37
78
0 min
0 min
15 min
15 min
60 min
60 min
75 72 Binding energy /eV
385
69
174
172
170 168 Binding energy /eV
166
164
386
Figure 4. High resolution XPS spectra of (a) Al 2p and S 2p for PFOA-Al2O3-PS before and
387
after the ball milling for 15 and 30 min.
388
Identification of Reactive Species. During the ball milling, PS is activated to produce
389
SO4•−,18 and the metal oxide generates oxygen vacancies and free electrons due to a collateral
390
effect of losing lattice oxygen.25, 39, 40 To confirm such reactive species during the milling of
391
Al2O3 and PS, we first employed DPPH• as a probe of free electrons and radicals. When the
392
mixture of DPPH• and Al2O3 is subjected to the ball milling, the absorption peak of DPPH• at
393
525 nm significantly decreased with the appearance of a new peak approximately 425 nm
394
corresponding to DPPH− (Figure 5a),25, 39 which indicates the generation of free electrons on
395
the milled Al2O3 (eq. 7). However, no DPPH− was observed by using PS and Al2O3 as
396
comilling agents, which suggests that the adsorbed PS as a strong oxidant effectively traps the
397
electrons generated on Al2O3 surface (eq. 8). In addition, MC-Al2O3-PS showed a broad
398
absorption peak approximately 440 nm, which is different from that of MC-PS (Figure 5a),
399
but similar to that of the coupling product between DPPH• with •OH.41
400
Al O milling Al VO (2e - ) 0.5O 2
401
e S2 O8
402
The radical intermediates were also investigated by ESR spectroscopic measurement of
2
surf
SO 4
surf
SO 4
2
(7) (8)
ACS Paragon Plus Environment
20
Page 21 of 37
Environmental Science & Technology
403
solids recovered from the milled samples in the presence of DMPO as a spin trapping reagent.
404
As shown in Figure 5b, MC-Al2O3 showed no detectable signal in either solid or solution
405
samples. The milled sample of PS (1.0 g) in the presence of DMPO exhibited the strong
406
characteristic peaks of a DMPO-SO4•− adduct with hyperfine splitting constants of aN= 13.2 G,
407
aH= 9.6 and 1.48G (Figure 5b), which confirms the decomposition of PS into SO4•− under the
408
ball milling conditions (eq. 9).42,
409
DMPO-SO4•− signals became weaker due to the decreased amount of PS (Figure 5b).
410
Moreover, an approximately 1:2:2:1 quartet pattern (aH= aN= 14.9 G) was observed, which
411
was assigned to the DMPO-•OH adduct,43 confirming that surface OH group on Al2O3
412
promotes the conversion of SO4•− into •OH (eq. 10). After extracting the milled sample with
413
moisture, formation of the DMPO-•OH adduct was observed in the two mixtures and
414
MC-Al2O3-PS exhibited a stronger DMPO-•OH signal than MC-PS (Figure S12), thereby
415
supporting the conversion of SO4•− to •OH on the Al2O3 surface (•OHsurf). Moreover, the ESR
416
signals for the milled samples of PS-DMPO and Al2O3-PS-DMPO remain considerably
417
intense after the milled samples were placed at room temperature for 22 h and 4 h,
418
respectively (Figure S13). The long-lived DMPO-radical adducts on the milled samples are
419
similar to “environmentally persistent free radicals” that formed on the surface of engineered
420
nanomaterials during combustion or thermal treatments,44 probably because the association of
421
the free radical formed in situ with the surface of milled solids stabilizes the radical. 2
milling 2SO 4
422
S2 O8
423
OH surf SO 4
surf
surf
43
For the mixture of Al2O3 (0.7 g) and PS (0.3 g), the
(9)
surf
OH surf SO 4
2
(10)
ACS Paragon Plus Environment
21
Environmental Science & Technology
(a)
Normalized Abs
1.2
Page 22 of 37
(b)
Al2O3+DPPH Al O +PS+DPPH 2 3
DMPO+Al2O3+PS
0.8 PS+DPPH DMPO+PS
0.4
DPPH DMPO+Al2O3
0.0 300
424
400
500 Wavelength /nm
600
700 3360
3390
3420 3450 Magnetic field /G
3480
3510
425
Figure 5. Normalized UV-vis absorption spectra of DMSO-extracted solutions from the
426
milled samples in the presence of DPPH• (a), where the absorbances of the peaks in the
427
region of 200 ~ 600 nm are normalized to 1.0. ESR spectra of milled samples in the presence
428
of DMPO (b).
429 430
To evaluate the contributions of SO4•−surf and •OHsurf, the effect of surface OH on Al2O3
431
was studied. We preheated Al2O3 powders at 100 °C for 24 h. This preheating decreased the
432
surface water and/or OH− on the Al2O3 by approximately 50% and consequently the
433
defluorination efficiency of PFOA in MC-Al2O3-PS after 2 h milling decreased from 100% to
434
64% (Figure S14). The MC degradation of PFOA was further performed by using 18O-labeled
435
Al2O3 as a comilling agent. Prior to the degradation experiments, Al2O3 powders were
436
dispersed in H218O for 1 h to ensure full oxygen exchange between H218O and the surface OH
437
on Al2O3, following by vacuum drying at 60 °C for 1 h to remove the surface water on Al2O3.
438
For convenience, the obtained
439
the defluorination efficiency of PFOA was reduced from 100% in MC-Al2O3-PS to 70.6% in
440
MC-(18OH-Al2O3)-PS at a 2 h milling (Figure 6a). Meanwhile, the isotope distribution of 18O
441
in another dominant product, CO2, was significantly changed. As shown in Figure 6b, the
442
collected gaseous product shows characteristic FT-IR absorptions of νas(C16O2) at 2361 and
18O-labeled
Al2O3 was named
ACS Paragon Plus Environment
22
18OH-Al O . 2 3
It was found that
Page 23 of 37
Environmental Science & Technology
443
2337 cm-1 in MC-PFOA-Al2O3-PS. These two peaks not only shift to lower wavenumbers by
444
approximately 17 cm-1 but also weaken in MC-PFOA-(18OH-Al2O3)-PS at the same milling
445
time, thus indicating that 16O in CO2 exchanged by 18O and the production of CO2 is slower
446
than in MC-PFOA-Al2O3-PS. Accordingly, the mass spectrum of CO2 produced in
447
MC-PFOA-Al2O3-PS exhibits a dominant m/z of 44 (C16O16O, 99.5%), being similar to
448
natural isotope abundance of CO2 in air (99.6%) and obtained from the reaction of Na2CO3
449
with HNO3 (99.3%) (Figures 6c and S15). In contrast, the ratios of CO2 with m/z of 44
450
(C16O16O), 46 (C16O18O) and 48 (C18O18O) in MC-PFOA-(18OH-Al2O3)-PS were changed to
451
63.6%, 32.0% and 4.4%, respectively (Figures 6c). Because the background influence of air
452
may cause a negative deviation in the evaluation of the absolute isotope ratio of
453
( R 18 O/16 O ), we did not calculate R 18 O/16 O . However, the mass area ratio of C16O18O to C18O18O
454
was 7.3, which suggests that one of the two oxygen atoms in CO2 comes from
455
surface OH on Al2O3. After collecting CO2 gas in NaOH solution and converting it to MnCO3
456
precipitant, the stable isotope ratio of
457
(VSMOW) was measured on an elemental analyzer coupled to an isotope ratio mass
458
spectrometer. The δ18O values for the MnCO3 obtained from MC-PFOA-Al2O3-PS and the
459
two reference compounds CO and ref-MnCO3 were 16.08‰, 9.07‰, and 38.02‰,
460
respectively, all of which are close to those of Pee Dee Belemnite, a carbonate isotope
461
standard (30‰ vs. VSMOW, Table S4). In contrast, the δ18O of MnCO3 obtained from
462
MC-PFOA-(18OH-Al2O3)-PS was 3251.50‰, which is 154 times higher than the average
463
value of the abovementioned three controls (21.05‰). The R 18 O/16 O values of CO32− and its
464
precursor CO2 in the case of MC-PFOA-(18OH-Al2O3)-PS were calculated to be 0.32 and 0.61,
465
respectively (Table S4). These results clearly demonstrated that the O atom in the surface OH
466
on Al2O3 can be incorporated into PFOA and finally transformed into CO2. By combining the
467
fact that the conversion of PFOA to CO2 in MC-Al2O3 and MC-PS was less than 20%, it is
18O/16O
18O/16O
18O-labeled
(δ18O, vs. Vienna standard mean ocean water
ACS Paragon Plus Environment
23
Environmental Science & Technology
Page 24 of 37
468
reasonable to propose that the surface OH− on Al2O3 is oxidized by SO4•−surf to •OHsurf, which
469
is then predominant species in the mineralization of PFOA.
Defluorination efficiency /%
100
(b)
(a)
15 min 60 min
75
900
18
MC-( OH-Al2O3)-PS
50
15 min
18
( OH-Al2O3)-PS Al2O3-PS
25
0
60 min MC-Al2O3-PS
0
60
120 180 Milling time /min
240
2500
2400 2300 -1 Wavenumber /cm
2200
(c) 8.0x10
6
CO2 obtained from
Intensity
Na2CO3+HNO3
6.0x10
6
4.0x10
6
2.0x10
6
MC-PFOA-Al2O3-PS 18
MC-PFOA-( OH-Al2O3)-PS Air
0.0 38
40
42
44
46
m/z
48
50
52
54
470 471
Figure 6. (a) Time profiles of defluorination efficiency of PFOA in MC-Al2O3-PS and
472
MC-PFOA-(18OH-Al2O3)-PS. (b) FT-IR and (c) mass spectra of collected gaseous products
473
from MC degradation of PFOA. (d) Isotopic oxygen composition (δ18O) of MnCO3 obtained
474
from degradation products of PFOA in MC-Al2O3-PS at a 2 h milling (S1-MnCO3) and
475
MC-(18OH-Al2O3)-PS at a 4 h milling (S2-MnCO3). CO2 obtained from the reaction Na2CO3
476
with HNO3 and indoor air, commercial CO gas and MnCO3 obtained from the reaction
477
Na2CO3 with MnSO4 (ref-MnCO3) were given as references in (c) and (d).
478
Reaction Mechanism. Based on the above results, the reaction mechanism of degradation of
479
PFOA using both Al2O3 and PS under the conditions of ball milling is proposed in Scheme 1.
480
After mixing PFOA with one agent, PFOA was tightly bound to Al2O3 through the COO−Al
ACS Paragon Plus Environment
24
Page 25 of 37
Environmental Science & Technology
481
bond and >AlOHF−C hydrogen bonding (eqs. 4 and 5), while the O2− of PS acted as the
482
proton accepter to bind PFOA through the hydrogen bond in MC-PS. In the ternary mixture,
483
the interaction between PFOA and Al2O3 does not change, but Al2O3 replaces PFOA to bind
484
PS via hydrogen bonding of >AlOHO2− (PS) (eq. 6). That is, Al2O3 strongly anchors the
485
PFOA and PS (Scheme 1). During the MC treatment, the high-energy impact has multiple
486
roles in the activation of solid-to-solid reactions by producing a transient temperature and
487
initiating the rotation, alignment and reorientation of molecules. On one hand, ball milling
488
induces the decarboxylation reaction of PFOA, thereby yielding CO2 and an unstable
489
perfluoroheptyl anion (−C7F15) on the Al2O3 surface (eq. 11) which preferentially reacts with
490
the proton of an OH group on Al2O3 to yield C7F15H (eq. 12). On the other hand, the
491
high-energy impact facilitates the decomposition of PS into SO4•−, which consequently
492
oxidizes OH−surf on Al2O3 to •OHsurf. As demonstrated in Figures 5 and 6, the active species
493
for the degradation of PFOA is •OHsurf in MC-Al2O3-PS. Hydrogen abstraction from C7F15H
494
by •OHsurf is followed by the formation of •C7F15 (eq. 13), which couples with •OHsurf to yield
495
C7F15OH (eq. 14). C7F15OH transforms into C6F13COF through the elimination of HF (eq. 15)
496
and reacts with the surface water to yield C6F13COOH together with H+ and F− (eq. 16).
497
Similarly, C6F13COOH is bound on the Al2O3 surface via the COO−Al and AlF
498
interactions, which activates the C−COO bond and C−F bonds, respectively (Scheme 1). This
499
leads the decarboxylation, which yields –C6F13, to take place as a second reaction on the
500
Al2O3 surface.
501
C 7 F15 COO Al milling C 7 F15 Al CO 2
(11)
502
( C 7 F15 )
(12)
503
(C 7 F15 H) surf OH surf ( C 7 F15 ) surf H 2 O
(13)
504
( C 7 F15 ) surf OH surf (C 7 F15 OH) surf
(14)
surf
H surf (C 7 F15 H) surf
ACS Paragon Plus Environment
25
Environmental Science & Technology
Page 26 of 37
505
(C 7 F15 OH) surf (C 6 F13COF) surf (HF (H F )) surf
(15)
506
C 6 F13COFsurf H 2 O surf C 6 F13COOH surf (HF (H F )) surf
(16)
507 508
Scheme 1. A schematic representation for MC degradation of PFOA in the presence of both
509
Al2O3 and PS.
510 511
The
successive
degradation
including
decarboxylation,
protonation,
512
hydrogen-abstraction, radical coupling, elimination of HF and hydrolysis (eqs. 11 ~ 16), occur
513
to yield short-chain PFCA products until CF3COOH is reached, which degrades to −CF3 on
514
the Al2O3 surface (eq. 11). Degradation and defluorination are repeated to yield CO2, H+, and
515
F−as the final products (eqs. 12 ~ 15). Thus, complete degradation and defluorination are
516
achieved. It is noted that the mineralization of 1 mole PFOA consumes 14 moles •OHsurf, of
517
which 7 moles •OHsurf are converted to CO2, and the others to H2O (Scheme S1). Moreover,
518
the mineralization of 1 mole PFOA generates 8 moles CO2, of which 7 moles come from the
519
produced short-chain PFCAs (Scheme S1). Each short-chain PFCA molecule (CnF2n+1COOH)
520
contains two external oxygen atoms: one from the addition of •OHsurf (eqs. 14 and 15, Scheme
521
S1), and the other from surface water via the hydrolysis reaction (eq. 16, Scheme S1). In
522
18O-labeling
experiments, all the surface OH− groups on Al2O3 were exchanged by the 18OH
ACS Paragon Plus Environment
26
Page 27 of 37
Environmental Science & Technology
523
of H218O, followed by drying at 60 °C for 1 h to remove surface water on the Al2O3. During
524
the MC treatment, the oxidation of surface 18OH− on Al2O3 by SO4•−surf yields •18OHsurf, while
525
surface water comes from moisture at natural isotope abundance (H216O), giving a theoretical
526
R 18 O/16 O in the produced CO2 of 0.78 as shown in eq. 17. The expected observations of
527
C16O18O as a main product and an
528
demonstrate that the MC degradation of PFOA in MC-Al2O3-PS follows a •OHsurf involved
529
pathway (eqs. 11 ~ 16).
R 18 O/16 O (CO2) = 0.6 (Figures 6b and 6c) clearly
C7F15COOHsurf + 718OH−surf + 7H+surf + 14•18OHsurf + 7H216Osurf 530
→ CO2 + 7C16O18O + 7H218O + 7H218O + 15HFsurf
(17)
531
From eq. 17, a complete defluorination of 1 mole PFOA may yield 15 moles F− and consumes
532
14 moles •OHsurf, being equivalent to 7 moles PS as calculated from eqs. 9 and 10. According
533
to eq. 3, the utilization efficiency of PS (PS) was calculated from the ratio of the
534
stoichiometrically required PS for generating F− to the actual consumption. For the MC
535
degradation of PFOA (0.25 g) at nAl O /nPS /nPFOA of 42: 8.4: 1, the residual PS decreased to
536
3.7% when the defluorination efficiency of PFOA was 100% (Figure S16a). The value of PS
537
in MC-Al2O3-PS remained at 87% during the defluorination of PFOA (Figure S16b). This
538
value is much higher than those for the defluorination of PFOA in aqueous solution (8% ~
539
22%) and aqueous soil slurry (< 1%).45−48
2
3
540
It should be noted that the main active species responsible for the degradation of PFOA
541
are the “free” electrons in MC-Al2O3 and SO4•− in MC-PS, respectively. This results in
542
different degradation pathways for PFOA from •OHsurf involved reactions in MC-Al2O3-PS.
543
In MC-Al2O3, the primary step of degradation of PFOA is the decarboxylation reaction (eq.
544
11) and then the “free” electrons on Al2O3 (eq. 7) enable the conversion of the perfluoroalkyl
545
anion to C5F11CF=CFH through protonation and reductive defluorination.25 In MC-PS, SO4•−
ACS Paragon Plus Environment
27
Environmental Science & Technology
546
oxidizes the PFOA adsorbed on the surface of PS powders to form unstable •C7F15 through a
547
one-electron oxidation, deprotonation, and decarboxylation (Text S2, eqs. S1 ~ 3). Further
548
reactions between •C7F15 and PFOA yield C7F15COOC7F15, C7F15COC7F15, C7F15OC7F15, and
549
C7F15COC6F13 as the main products (Text S2, eqs. S4 ~ 7, Figure S7). Thus, the TOC removal
550
and defluorination ratios were as low as 11% and 6.4%, respectively (Figures 1b and 2b),
551
although 61% of the PFOA was degraded in MC-PS after a 2 h milling. It is also noted that
552
the defluorination efficiency of PFOA in MC-PS (< 10%) is lower than those (> 70%)
553
induced by photolysis or thermolysis of PS in aqueous solution,45−47 due to the different
554
degradation pathways of PFOA. Upon treatment with PS in water, SO4•− can be converted
555
into •OH, and both of them contribute to the degradation of PFOA. The unstable •C7F15
556
generated from oxidation by SO4•− usually reacts with •OH, dissolved O2 and/or sufficient
557
H2O to yield C6F13COOH, which follows an analogous mechanism to produce short-chain
558
PFCAs products until complete mineralization and defluorination occurs.45−48 The fast
559
diffusion of chemical species in water will not cause PFCA molecules to remain close to the
560
PS species until the parent PFCA molecule is completely defluorinated. As a result, quite
561
large amounts of short-chain PFCAs were detected as major products during the
562
defluorination of PFOA in PS-involved aqueous solution.45−48
563
Applications and Implications. We found complete degradation and defluorination of
564
PFOA through activation of C−COO− and C−F bonds and generation of •OH by the MC
565
treatment using Al2O3 and PS as comilling agents, respectively. The MC treatment was also
566
used to degrade the homologues of PFOA such as PFPA, PFHxA, and PFHpA under the same
567
conditions (i.e., mAl O /mPS /mPFCA , 2.58: 1.37: 0.25; mb/ms, 50:1). After ball milling for 2 h,
568
such PFOA homologues are degraded with a defluorination efficiency of 100% (Figure S17).
569
Perfluorooctane sulfonic acid (PFOS) is another important perfluorinated pollutant. In general,
570
PFOS is much more resistant to oxidative degradation by SO4•− than PFCAs, and hence, no
2
3
ACS Paragon Plus Environment
28
Page 28 of 37
Page 29 of 37
Environmental Science & Technology
571
transformation of PFOS was observed in aqueous solution or soil slurry in the presence of
572
heat-activated PS.47,48 We used the developed MC method to degrade PFOS, and found that
573
56% of the added PFOS was slowly degraded in MC-Al2O3-PS with a defluorination
574
efficiency of 9.7% after a 2 h milling (Figure S18). The considerably slower degradation of
575
PFOS than PFOA was in good agreement with the abovementioned opinion that PFOS is
576
much more resistant to oxidative degradation by SO4•−. Furthermore, the enhanced
577
degradation of PFOS in MC-Al2O3-PS relative to that in MC-PS may also support the
578
important role of the Al2O3 surface in the activation of the perfluorinated molecules. The
579
lower extent of defluorination of PFOS in MC-Al2O3-PS was due to the C−SO3− bond in
580
PFOS being cleaved by the oxidation to produce SO3 and •C7F15 (eq. 18) with the generated
581
SO3 reacting with OH−surf (eq. 19), which suppressed the transformation of SO4•− to •OHsurf
582
(eq. 10), significantly decreasing the generation of •OH for its coupling with •C7F15. Finally,
583
the main fortune of •C7F15 is to form dimers with itself without considerable elimination of F
584
atoms.
C 7 F15 SO 3 SO 4
585
C 7 F15SO 3 SO 4
586
SO3 OH surf SO 4 H
surf
2
2
(18) (19)
587
The advanced oxidation processes (AOPs) using •OH or SO4•− have drawn interest in the
588
degradation of organic pollutants.8−12 However, there are limited AOPs capable of the
589
complete defluorination and mineralization of PFOA pollutants in solids. The present study of
590
MC treatment using both Al2O3 and PS provides valuable information for the design of AOPs
591
for the complete defluorination and mineralization of refractory PFCAs. Moreover, the
592
normalized energy consumption for releasing one micro-molar fluorine ions from the
593
degradation of PFOA (E, kJ μmol−1) was evaluated to be approximately 0.15 kJ μmol−1 in
594
MC-Al2O3-PS. This is the same order of magnitude in previously reported MC treatments,21
595
but higher energy efficiency is obtained using this method than using other existing methods
ACS Paragon Plus Environment
29
Environmental Science & Technology
Page 30 of 37
596
(Table 2). It was noted that the degradation and defluorination efficiencies of PFOS in
597
MC-Al2O3-PS were higher than those (44% and 3.4%, respectively) in MC-PS, but lower than
598
those in MC-Al2O3 (96% and 21%) (Figure S18). This inspires us to develop in the future an
599
appropriate MC treatment method to degrade PFOS.
600 601
Table 2. Comparison of energy consumption for the degradation of PFOA by different
602
treatment methods. Treatment method MC-Al2O3-PS MC-KOH
MC-NaOH
UV/In2O3 UV/BiOCl UV/PS UV/H2O2 US+SnO2-Sb/ carbon aerogel electrode
Conditions
Power
Deflurination
Energy
(W)
efficiency (%)
(kJ mol−1)a
750
100 (2 h)
0.15 b
750
97 (4 h)
0.39 b
21
750
19 (4 h)
1.99 b
21
23
33.7 (4 h)
1.64
49
10
59.3 (24 h)
24.3
50
200
59.1 (4 h)
10.8
45
9
46 (24 h)
28.2
51
> 50 c
81 (5 h)
> 5.10
52
0.604 mmol PFOA, 2.58 g Al2O3, 1.37 g PS 0.483 mmol PFOA, 4.6 g KOH 0.483 mmol PFOA, 4.6 g NaOH 0.1 mM PFOA, 0.4 L 0.5 g L−1 In2O3 0.02 mM PFOA, 0.2 L, 0.5 g L-1 BiOCl 29.6 mol PFOA, 1.10 mmol PS 0.02 mM PFOA, 0.2 L, 40 mM H2O2, 4 mM Fe2+ 0.1 mg L−1 PFOA, 0.06 L, immersed area 5 cm2
Ref. This work
603
a
Energy consumed to release one micro-molar F− ions from the degradation PFOA, which is
604
calculated from eq. S11 (Text S3). b Planetary ball mill is a robust floor model with 4 grinding
605
stations, and thus the treated PFOA should be 4 times of the added pollutants.
606
power includes the ultrasonic irradiation (50 W) and the employed power in electrochemical
607
experiment, but the latter of them is unknown.
ACS Paragon Plus Environment
30
c
The total
Page 31 of 37
Environmental Science & Technology
608
ASSOCIATED CONTENT
609
Supporting Information
610
Supporting texts (Text S1−S3), figures (Figure S1−S18), and table (Table S1−S4), as
611
mentioned in the supporting information. This material is available free of charge via the
612
Internet at http://pubs.acs.org.
613 614
615
Corresponding Author
616
** Prof. Lihua Zhu, Fax: +86 27 8754 3632; E-mail:
[email protected].
617
Prof. HeqingTang, Fax: +86 27 6784 3990; E-mail:
[email protected].
618
Notes
619
The authors declare no competing financial interest.
AUTHOR INFORMATION
620 621
622
This research was financially supported by the National Natural Science Foundation of China
623
(Grants Nos. 21777194 and 21477043). We thank to Prof. Anmin Zhen and, Dr. Shenhui Li,
624
Wuhan Institute of Physics and Mathematics of the Chinese Academy of Science, and Dr.
625
Zhiwu Yu, High Magnetic Field Laboratory of the Chinese Academy of Science for their
626
help.
ACKNOWLEDGEMENTS
627 628
629
(1) Zareitalabad, P.; Siemens, J.; Hamer, M.; Amelung, W. Perfluorooctanoic acid (PFOA)
630
and perfluorooctanesulfonic acid (PFOS) in surface waters, sediments, soils and
631
wastewater-A review on concentrations and distribution coefficients. Chemosphere 2013, 91,
632
725−732.
633
(2) Houde, M.; De Silva, A. O.; Muir, D. C.; Letcher, R. J. Monitoring of perfluorinated
634
compounds in aquatic biota: an updated review. Environ. Sci. Technol. 2011, 45, 7962−7973.
REFERENCES
ACS Paragon Plus Environment
31
Environmental Science & Technology
635
(3) Xiao, F.; Simcik, M. F.; Halbach, T. R.; Gulliver, J. S. Perfluorooctane sulfonate (PFOS)
636
and perfluorooctanoate (PFOA) in soils and groundwater of a U.S. metropolitan area:
637
migration and implications for human exposure. Water Res. 2015, 72, 64−74.
638
(4) Cagnetta, G.; Robertson, J.; Huang, J.; Zhang, K.; Yu, G. Mechanochemical destruction of
639
halogenated organic pollutants: A critical review. J. Hazard. Mater. 2016, 313, 85−102.
640
(5) Song, Z.; Tang, H.; Wang, N.; Zhu, L. Reductive defluorination of perfluorooctanoic acid
641
by hydrated electrons in a sulfite-mediated UV photochemical system. J. Hazard. Mater.
642
2013, 262, 332−338.
643
(6) Gu, Y.; Dong, W.; Luo, C.; Liu, T. Efficient reductive decomposition of perfluorooctane
644
sulfonate in a high photon flux UV/sulfite system. Environ. Sci. Technol. 2016, 50,
645
10554−10561.
646
(7) Tang, H.; Xiang, Q.; Lei, M.; Yan, J.; Zhu, L.; Zou, J. Efficient degradation of
647
perfluorooctanoic acid by UV-Fenton process. Chem. Eng. J. 2012, 184, 156−162.
648
(8) Liu, Y.; Chen, S.; Quan, X.; Yu, H.; Zhao, H.; Zhang, Y. Efficient mineralization of
649
perfluorooctanoate by electro-Fenton with H2O2 electro-generated on hierarchically porous
650
carbon. Environ. Sci. Technol. 2015, 49, 13528−13533.
651
(9) Song, C.; Chen, P.; Wang, C.; Zhu, L. Photodegradation of perfluorooctanoic acid by
652
synthesized TiO2-MWCNT composites under 365nm UV irradiation. Chemosphere 2012, 86,
653
853−859.
654
(10) Sun, B.; Ma,J.; Sedlak, D. L. Chemisorption of perfluorooctanoic acid on powdered
655
activated carbon initiated by persulfate in aqueous solution. Environ. Sci. Technol. 2016, 50,
656
7618−7624.
657
(11) da Silva-Rackov, C. K. O.; Lawal, W. A.; Nfodzo, P. A.; Vianna, M. M. G. R.; do
658
Nascimento, C. A. O.; Choi, H. Degradation of PFOA by hydrogen peroxide and persulfate
659
activated by iron-modified diatomite. Appl. Catal. B Environ. 2016, 192, 253−259.
ACS Paragon Plus Environment
32
Page 32 of 37
Page 33 of 37
Environmental Science & Technology
660
(12) Qian, Y.; Guo, X.; Zhang, Y.; Peng, Y.; Sun, P.; Huang, C. H.; Niu, J.; Zhou, X.;
661
Crittenden, J. C. Perfluorooctanoic acid degradation using UV-persulfate process: Modeling
662
of the degradation and chlorate formation. Environ. Sci. Technol. 2016, 50, 772−781.
663
(13) Wang, F.; Lu, X.; Li, X. Y.; Shih, K. Effectiveness and mechanisms of defluorination of
664
perfluorinated alkyl substances by calcium compounds during waste thermal treatment.
665
Environ. Sci. Technol. 2015, 49, 5672−5680.
666
(14) Zhang, Q.; Matsumoto, H.; Saito, F.; Baron, M. Debromination of hexabromobenzene by
667
its co-grinding with CaO. Chemosphere 2002, 48, 787−793.
668
(15) Zhang, Q.; Saito, F.; Ikoma, T.; Tero-Kubota, S. Effects of quartz addition on the
669
mechanochemical dechlorination of chlorobiphenyl by using CaO. Environ. Sci. Technol.
670
2001, 35, 4933−4935.
671
(16) Zhang, W.; Huang, J.; Xu, F.; Deng, S.; Zhu, W.; Yu, G. Mechanochemical destruction
672
of pentachloronitrobenzene with reactive iron powder. J. Hazard. Mater. 2011, 198, 275−281.
673
(17) Zhang, K.; Huang, J.; Zhang, W.; Yu, Y.; Deng, S.; Yu, G. Mechanochemical
674
degradation of tetrabromobisphenol A: Performance, products and pathway. J. Hazard.
675
Mater. 2012, 243, 278−285.
676
(18) Huang, A.; Zhang, Z.; Wang, N.; Zhu, L.; Zou, J. Green mechanochemical oxidative
677
decomposition of powdery decabromodiphenyl ether with persulfate. J. Hazard. Mater. 2016,
678
302, 158−165.
679
(19) Liu, X.; Zhang, X.; Zhang, K.; Qi, C. Sodium persulfate-assisted mechanochemical
680
degradation of tetrabromobisphenol A: Efficacy, products and pathway. Chemosphere 2016,
681
150, 551−558.
682
(20) Yan, X.; Liu, X.; Qi, C.; Wang, D.; Lin, C. Mechanochemical destruction of a
683
chlorinated polyfluorinated ether sulfonate (F-53B, a PFOS alternative) assisted by sodium
684
persulfate. RSC Adv. 2015, 5, 85785−85790.
ACS Paragon Plus Environment
33
Environmental Science & Technology
685
(21) Zhang, K.; Huang, J.; Yu, G.; Zhang, Q.; Deng, S.; Wang, B. Destruction of
686
perfluorooctane sulfonate (PFOS) and perfluorooctanoic acid (PFOA) by ball milling.
687
Environ. Sci. Technol. 2013, 47, 6471−6477.
688
(22) Zhang, K.; Cao, Z.; Huang, J.; Deng, S.; Wang, B.; Yu, G. Mechanochemical destruction
689
of Chinese PFOS alternative F-53B. Chem. Eng. J. 2016, 286, 387–393.
690
(23) Chen, W.; Hooper, T.; Ng, J.; White, A. J. P.; Crimmin, M. R. Palladium-catalyzed
691
carbon-fluorine and carbon-hydrogen bond alumination of fluoroarenes and heteroarenes.
692
Angew. Chem. Int. Ed. 2017, 56, 12687−12691.
693
(24) Senaweera, S. M.; Singh, A.; Weaver, J. D. Photocatalytic hydrodefluorination: Facile
694
access to partially fluorinated aromatics. J. Am. Chem. Soc. 2014, 136, 3002−3005.
695
(25) Lv, H.; Wang, N.; Zhu, L.; Zhou, Y.; Li, W.; Tang, H. Alumina-mediated
696
mechanochemical method for simultaneously degrading perfluorooctanoic acid and
697
synthesizing a polyfluoroalkene. Green Chem. 2018, 20, 2526−2533.
698
(26) Cagnetta, G.; Zhang, Q., Huang, J.; Lu, M.; Wang, B.; Wang, Y.; Deng, S.; Yu, G.
699
Mechanochemical destruction of perfluorinated pollutants andmechanosynthesis of lanthanum
700
oxyfluoride: A waste-to-materialsprocess. Chem. Eng. J. 2017, 316, 1078−1090.
701
(27) Zhou, Q.; Chen, M.; Zhu, L.; Tang, H. Determination of perfluorinated carboxylic acids
702
in water using liquid chromatography coupled to a corona-charged aerosol detector. Talanta
703
2015, 136, 35−41.
704
(28) Liang, C. J.; Huang, C. F.; Mohanty, N.; Kurakalva, R.M. A rapid spectrophotometric
705
determination of persulfate anion in ISCO. Chemosphere 2008, 73, 1540−1543.
706
(29) Delogu, F. Molecular dynamics of atomic rearrangements at perturbed surfaces. Phys.
707
Rev. B 2009, 80, 014115-1−014115-8
708
(30) G. Cagnetta, K. Zhang, Q. Zhang, J. Huang, G. Yu. Mechanochemical pre-treatment for
709
viable recycling of plastic waste containing haloorganics. Waste Manage. 2018, 75, 181–186.
ACS Paragon Plus Environment
34
Page 34 of 37
Page 35 of 37
Environmental Science & Technology
710
(31) Cagnetta, G., Huang, J., Wang, B., Deng, S., Yu, G. A comprehensive kinetic model for
711
mechanochemical destruction of persistent organic pollutants. Chem. Eng. J. 2016, 291, 30–
712
38.
713
(32) Liu, Q.; Huang, A.; Wang, N.; Zheng, G.; Zhu, L. Rapid fluorometric determination of
714
perfluorooctanoic acid by its quenching effect on the fluorescence of quantum dots. J. Lumin.
715
2015, 161, 374–381.
716
(33) Nara, M.; Torii, H.; Tasumi, M. Correlation between the vibrational frequencies of the
717
carboxylate group and the types of its coordination to a metal ion: An ab Initiomolecular
718
orbital study. J. Phys. Chem. 1996, 100, 19812−19817.
719
(34) Wang, M.; Wang, N.; Tang, H.; Cao, M.; She, Y.; Zhu, L. Surface modification of
720
nano-Fe3O4 with EDTA and its use in H2O2 activation for removing organic pollutants. Catal.
721
Sci. Technol. 2012, 2, 187−194.
722
(35) Lu, W.; Iwasa, Y.; Ou, Y.; Jinno, D.; Kamiyama, S.; Petersen, P. M.; Ou, H. Effective
723
optimization of surface passivation on porous silicon carbide using atomic layer deposited
724
Al2O3. RSC Adv. 2017, 7, 8090−8097.
725
(36) Shimizu, K.; Driver, G. W.; Lucas, M.; Sparrman, T.; Shchukarev, A.; Boily, J. F.
726
Bifluoride ([HF2]−) formation at the fluoridated aluminium hydroxide/water interface. Dalton
727
Trans. 2016, 45, 9045−9050.
728
(37) Groveman, S.; Peng, J.; Itin, B.; Diallo, I.; Pratt, L. M.; Greer, A.; Biddinger, E. J.;
729
Greenbaum, S. G.; Drain, C. M.; Francesconi, L.; Vittadello, M. The role of ozone in the
730
formation and structural evolution of graphene oxide obtained from nanographite. Carbon
731
2017, 122, 411−421.
732
(38) Pany, S.; Naik, B.; Martha, S.; Parida, K. Plasmon induced nano Au particle decorated
733
over S,N-modified TiO2 for exceptional photocatalytic hydrogen evolution under visible light.
734
ACS Appl. Mater. Interfaces 2014, 6, 839−846.
ACS Paragon Plus Environment
35
Environmental Science & Technology
Page 36 of 37
735
(39) Cagnetta, G.; Huang, J.; Lu, M.; Wang, B.; Wang, Y.; Deng, S.; Yu,G. Defect
736
engineered oxides for enhanced mechanochemical destruction of halogenated organic
737
pollutants. Chemosphere 2017, 184, 879−883.
738
(40) Chai, H.; Zhang, Z.; Zhou, Y.; Zhu, L.; Lv, H.; Wang, N. Roles of intrinsic Mn3+ sites
739
and
740
decabromodiphenyl ether with manganese dioxide. Chemosphere 2018, 207, 41−49.
741
(41) Hristea, E. N.; Hillebrand, M.; Caproiu, M. T.; Caldararu, H.; Constantinescu, T.;
742
Balaban, A. T. Scavenging the hydroxyl radical by 2,2-diphenyl-1-picrylhydrazyl. ARKIVOC
743
2002, ii, 123−132.
744
(42) Huang, Z.; Bao, H.; Yao, Y.; Lu, W.; Chen, W. Novel green activation processes and
745
mechanism of peroxymonosulfate based on supported cobalt phthalocyanine catalyst. Appl.
746
Catal. B Environ. 2014, 154−155, 36−43.
747
(43) Fang, G.; Gao, J.; Dionysiou, D. D.; Liu, C.; Zhou, D. Activation of persulfate by
748
quinones: Free radical reactions and implication for the degradation of PCBs. Environ. Sci.
749
Technol. 2013, 47, 4605−4611.
750
(44) Vejerano, E. P.; Rao, G.; Khachatryan, L.; Cormier, S. A.; Lomnicki, S. Environmentally
751
persistent free radicals: Insights on a new class of pollutants. Environ. Sci. Technol. 2018, 52,
752
2468−2481.
753
(45) Hori, H., Yamamoto, A., Hayakawa, E., Taniyasu, S., Yamashita, N., Kutsuna, S.,
754
Kiatagawa, H., Arakawa, R. Efficient decomposition of environmentally persistent
755
perfluorocarboxylic acids by use of persulfate as a photochemical oxidant. Environ. Sci.
756
Technol. 2005, 39, 2383−2388.
757
(46) Hori, H.; Nagaoka, Y.; Murayama, M.; Kutsuna, S. Efficient decomposition of
758
perfluorocarboxylic acids and alternative fluorochemical surfactants in hot water. Environ.
759
Sci. Technol. 2008, 42, 7438−7443.
lattice
oxygen
in
mechanochemical
debromination
ACS Paragon Plus Environment
36
and
mineralization
of
Page 37 of 37
Environmental Science & Technology
760
(47) Bruton, T. A.; Sedlak, D. L. Treatment of perfluoroalkyl acids by heat-activated
761
persulfate under conditions representative of in situ chemical oxidation. Chemosphere 2018,
762
206, 457−464.
763
(48) Park, S.; Lee, L. S.; Medina, V. F.; Zull, A.; Waisner, S. Heat-activated persulfate
764
oxidation of PFOA, 6:2 fluorotelomer sulfonate, and PFOS under conditions suitable for
765
in-situ groundwater remediation. Chemosphere 2016, 145, 376−383.
766
(49) Li, X.; Zhang, P.; Jin, L.; Shao, T.; Li, Z.; Cao, J. Efficient photocatalytic decomposition
767
of perfluorooctanoic acid by indium oxide and its mechanism. Environ. Sci. Technol. 2012,
768
46, 5528−5534.
769
(50) Song, Z.; Dong, X.; Wang, N.; Zhu, L.; Luo, Z.; Fang, J.; Xiong, C. Efficient
770
photocatalytic defluorination of perfluorooctanoic acid over BiOCl nanosheets via a hole
771
direct oxidation mechanism, Chem. Eng. J. 2017, 317, 925−934.
772
(51) Tang, H.; Xiang, Q.; Lei M.; Yan, J.; Zhu, L.; Zou, J. Efficient degradation of
773
perfluorooctanoic acid by UV-Fenton process, Chem. Eng. J. 2012, 184, 156–162.
774
(52) Zhao, H.; Gao, J.; Zhao, G.; Fan, J.; Wang, Y.; Wang, Y. Fabrication of novel
775
SnO2-Sb/carbon
776
perfluorooctanoate with high catalytic efficiency. Appl. Catal. B Environ. 2013, 136,
777
278−286.
aerogel
electrode
for
ultrasonic
ACS Paragon Plus Environment
37
electrochemical
oxidation
of