Subscriber access provided by UNIV OF LOUISIANA
Remediation and Control Technologies
Reductive debromination of polybrominated diphenyl ethers: Dependence on Br number of the Br-rich phenyl ring Shun Guo, Lihua Zhu, Tetsuro Majima, Ming Lei, and Heqing Tang Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 26 Mar 2019 Downloaded from http://pubs.acs.org on March 27, 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 28
Environmental Science & Technology
1
Reductive debromination of polybrominated diphenyl ethers:
2
Dependence on Br number of the Br-rich phenyl ring
3
Shun Guoa, Lihua Zhua*, Tetsuro Majimaa, Ming Leib, Heqing Tangb,*
4
a Key
5
of Education), School of Chemistry and Chemical Engineering, Huazhong University
6
of Science and Technology, Wuhan 430074, PR China
7
b
8
Nationalities, Wuhan 430074, PR China
Laboratory of Material Chemistry for Energy Conversion and Storage (Ministry
College of Resources and Environmental Science, South-Central University for
9 10
Graphic abstract Reactive ring
Br m
m = 1~5 n = 0~5 m≥n
+
1
O
eBr m
Br n
Br n
··
- Br -
10-6
+H Br m-1
Br n
10-2
10-4
O
O
Relative rate constant (kR)
O
Br m-1
Br n
10-8
11 12
1
ACS Paragon Plus Environment
1
2
3
m
4
5
Environmental Science & Technology
13
ABSTRACT: Reductive debromination has been widely studied for the degradation
14
of polybrominated diphenyl ethers (PBDEs), although the reaction mechanisms are
15
not so clear. In the present study, the photocatalytic degradation and debromination of
16
ten PBDEs were carried out with CuO/TiO2 nanocomposites as the photocatalyst
17
under anaerobic condition. The pseudo-first-order rate constants were obtained for the
18
photocatalytic debromination of PBDEs, and their relative rate constants (kR) were
19
evaluated against kR= 1 for BDE209. Unlike the generally accepted summary that kR
20
is dependent on total Br number (N) of PBDEs, kR is found to depend on Br number
21
on a phenyl ring with more Br atoms than the other one. In other word, a phenyl ring
22
substituted by more Br is more reactive for the reductive debromination. The
23
calculated LUMO energies (ELUMO) of all PBDEs are well correlated to more reactive
24
phenyl ring with more Br, compared with N of two phenyl rings. The result was
25
explained by LUMO localization on the Br-rich phenyl ring, suggesting the reductive
26
debromination occurs on the phenyl ring.
2
ACS Paragon Plus Environment
Page 2 of 28
Page 3 of 28
Environmental Science & Technology
28
INTRODUCTION
29
Polybrominated diphenyl ethers (PBDEs) are well known as brominated flame
30
retardants (BFRs) to be extensively used in various consumer products over the past
31
four decades.1 It is estimated that approximately 46 000, 25 000, and 380 000 tons of
32
commercial penta-BDE, octa-BDE, and deca-BDE, respectively, were and will be
33
consumed in the United States and Canada during 1970~2020.2 The widespread use of
34
PBDEs makes them ubiquitous in the environment all over the world.3-5 It is known
35
that PBDEs are bio-accumulative to cause thyroid dysfunctions and developmental
36
neurotoxicity.6,7 However, as a class of persistent organic pollutants (POPs), the
37
natural degradation of PBDEs is very slow. Therefore, much attention has been paid
38
to develop degradation of such pollutants.
39
The oxidative degradation of PBDEs is only possible under very extreme
40
conditions: prolonged UV light irradiation8 or concentrated sulfuric acid as solvent9.
41
On the other hand, many reductive processes such as zerovalent iron (ZVI)
42
reduction,10,11
43
biodegradation15 have been studied, and a stepwise debromination mechanism is
44
suggested. As PBDEs, penta-BDE, octa-BDE, and deca-BDE have been listed as
45
POPs by Stockholm Convention.16 The persistence of these POPs in natural
46
environment is resulted from their low reactivity. More seriously, during the reductive
47
debromination of highly brominated PBDEs (for example, deca-BDE, i.e. BDE209),
48
low-brominated PBDEs are gradually accumulated to be less reactive for the
49
subsequent reductive debromination,10,12,17 but such low-brominated PBDEs are more
photocatalytic
reduction,12,13
electrochemical
3
ACS Paragon Plus Environment
reduction,14
and
Environmental Science & Technology
50
toxic than highly brominated PBDEs.18
51
The dependence of the reductive debromination rate on total Br number of
52
PBDEs has attracted much attention. It has been shown that the reactivity of the
53
debromination of PBDEs decreases with decreasing Br number in PBDEs.10,11,17,19
54
However, Granelli et al. studied the relative reduction rates of fifteen PBDEs by
55
sodium borohydride,17 and found that BDE181 (a hepta-BDE with total Br number
56
(N) of 7) was more reactive than two oct-BDEs, BDE201 and BDE202 (N= 8),
57
suggesting that the reactivity related to the substitution pattern. This experimental
58
result initiates the present study.
59
Generally, the electron-transfer initiated reductive dehalogenation of aromatic
60
halides includes two successive electron transfer processes.20-22 Firstly, electron is
61
transferred to the lowest unoccupied molecular orbital (LUMO) of aromatic halides,
62
giving the radical anion. Then, the electron of the radical anion transfers to σ
63
anti-bonding orbital of carbon-halogen bond (σC-X*), leading to the cleavage of a C-X
64
bond to give aromatic carbon radical and halide ion. The carbon radical abstracts a
65
hydrogen atom from solvent molecules to produce an aromatic compound with less
66
halogen atoms. If the reductive debromination of PBDEs precedes via this reaction
67
mechanism, their LUMO energies (ELUMO) can be a good descriptor for predicting the
68
reduction rates of PBDEs. Indeed, a good correlation has been observed for the
69
relation between the reaction rates and ELUMO of PBDEs.10,11 However, the relation
70
between ELUMO of PBDEs and their chemical structures is still unclear.
71
To clarify the reductive debromination mechanism, we obtained the relative rate 4
ACS Paragon Plus Environment
Page 4 of 28
Page 5 of 28
Environmental Science & Technology
72
constant of the photocatalytic reductive debromination of PBDEs. We confirmed that
73
the relative rate constant is mainly dependent on Br number of one phenyl ring with
74
more Br than the other ring. The dependence of Br number of one phenyl ring on the
75
reductive debromination is explained by LUMO distributions of PBDEs to be more
76
localized on it than on the other one.
77
EXPERIMENTAL SECTION
78
Chemicals and materials. BDE209 (purity >98%), BDE47 (>98.5%), and BDE15
79
(>98%) were purchased from J&K Chemical, China. BDE99 (>95%) was supplied by
80
Wuhan Kaymke Chemical, China. TiO2 powders (P25, ca. 80% anatase, 20% rutile;
81
surface area, ca. 50 m2 g-1) were purchased from the Degussa,Germany. All reagents
82
were analytical grade and used without further purification. In addition, six PBDEs
83
including
84
2-methyl-3,4,5,6-tetrabromo-diphenyl ether (2-CH3-BDE61, as an alternative of
85
BDE61) were synthesized and characterized in the present study. The details are
86
available in the Supporting Information.
87
Photocatalytic degradation of PBDEs. Photocatalytic debromination of PBDEs was
88
conducted in a quartz vessel under Ar atmosphere at 30 oC by using CuO/TiO2
89
nanocomposites as a photocatalyst. The photocatalyst was prepared according to the
90
reported method to show a good photocatalytic activity for the reductive
91
debromination of low-brominated PBDEs as reported in our recent work.23 Typically,
92
CuO/TiO2 (10 mg) was suspended in a 50 mL of a given PBDE solution (10-5 mol L-1)
93
in methanol. After the suspension was ultrasonically dispersed for 2 min, it was
BDE166,
BDE116,
BDE75,
5
ACS Paragon Plus Environment
BDE30,
BDE7,
and
Environmental Science & Technology
94
purged with Ar for 30 min under magnetic stirring to remove dissolved O2. The
95
photocatalytic debromination was started by switching on the light, and during the
96
experiment the suspension was magnetically stirred under the protection of Ar
97
atmosphere. A PLS-SXE300 Xe lamp (Beijing Perfect Light, China) equipped with a
98
365 nm bandpass filter (365 ± 5 nm) was used as the light source. At given reaction
99
time intervals, aliquots (1 mL) were sampled, immediately centrifuged, and filtered
100
through a 0.22 μm membrane to remove the catalyst. The filtrates were analyzed by
101
HPLC (Agilent 1260 series, USA) equipped with a diode array and an SB-C18
102
column (4.6 × 150 mm). The mobile phase was 90% acetonitrile and 10% water at a
103
rate of 1.0 mL min-1, and the detection wavelength was set at 240 nm. Each of all the
104
experiments were usually conducted in triplicate, and the average values were
105
obtained throughout the present work. It should be noted that we do not need the same
106
irradiation intensity, because only relative debromination rate constant of PBDE is
107
used in the present work. Because the relative reactivity of PBDEs varies over 7
108
orders of magnitude as demonstrated in the present work, the irradiation intensity is
109
required to be adjusted in a wide range by changing irradiation path length and
110
irradiation area (light spot area). As a reference, the irradiation at 365 nm was
111
typically conducted with 4 mW cm-2 and an irradiation area of 1 cm-2. The irradiation
112
intensity in other cases was expressed by transforming it to an equivalent value with
113
an irradiation area of 1 cm-2.
114
GC-MS analysis method. The intermediate products were identificated by GC-MS
115
(TRACE 1300 GC coupled to ISQ system, Thermo Fisher Scientific, USA.) equipped 6
ACS Paragon Plus Environment
Page 6 of 28
Page 7 of 28
Environmental Science & Technology
116
with a TR-5MS capillary column (30 m × 0.25 mm × 0.25 μm). The initial oven
117
temperature was 40 oC. After maintaining for 1 min, the oven temperature was
118
increased by 26 oC min-1 to a final temperature 300 oC, and then held for 10 min. The
119
injector, transfer line and ion source temperatures were set at 280 oC. And the full
120
scan mode (m/z 50-700) was used for mass analysis.
121
Computation methods. All theoretical calculations were carried out with Gaussian09
122
program.24 The molecular geometries and molecular orbital properties of 209 PBDEs
123
were obtained with the aid of density functional theory (DFT) calculations using the
124
Becke’s
125
gradient-corrected correlation functional (B3LYP hybrid functional),25,26 which were
126
widely used in various studies27-29 with 6-31+G(d) basis sets.30 No symmetry
127
restriction was imposed during the optimization process. Because the reductive
128
debromination of PBDEs was carried in methanol, methanol was chosen as a model
129
solvent using the most widely used conductor polarized continuum model (CPCM).
three-parameter
hybrid
exchange
function
with
Lee-Yang-Parr
130
RESULTS AND DISCUSSION
131
Weak dependence of the reductive debromination of PBDEs on total Br
132
number. All PBDEs have a symmetric skeletal structure with two separated phenyl
133
groups connected by an ether bond (Figure 1a), in which a torsional angle exists
134
between two planes of phenyl rings. Five sites of each phenyl ring can be replaced by
135
Br. In the present study, the two phenyl rings were assigned to “m ring” and “n ring”,
136
representing the ring with more and less Br substitutions, respectively, m = 1~5, n =
137
0~5 and m ≥ n as shown in Figure 1a. 7
ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 28
138
To investigate the dependence of the reductive debromination of PBDEs on N,
139
we conducted the CuO/TiO2 photocatalytic degradation of nine PBDEs and
140
2-methyl-3,4,5,6-tetrabromo-diphenyl
141
atmosphere. Here, 2-CH3-BDE61 with m= 4 and a methyl group on m ring was
142
synthesized instead of BDE61 because we could not synthesize BDE61. Since the
143
ELUMO= -1.912 eV of 2-CH3-BDE61 is similar to ELUMO= -1.873 eV of BDE61, it is
144
rational to use 2-CH3-BDE61 as BDE61 in the present work. The ten PBDEs covered
145
N= 2, 3, 4, 5, 6, and 10, and also covered several groups of m ring having m = 1, 2, 3,
146
and 5. The photocatalytic reductive degradation of PBDEs was analyzed by a
147
pseudo-first-order rate equation (eq. 1), ln (c/c0) = - kt
148 149 150
ether
(2-CH3-BDE61)
under
anaerobic
(1)
where k is the pseudo-first-order rate constant (min-1), t is the reaction time (min), c0 and c are the concentrations of PBDE at 0 and t min, respectively.
151
Figure 1b showed the debromination of BDE166, BDE116, and BDE47 in a
152
quartz vessel under irradiation at 365 nm with 4 mW cm-2. After irradiation for 12.5
153
min, the removal of BDE166 (N= 6) and BDE116 (N= 5) were 85 and 70%,
154
respectively. k= 0.15 and 0.099 min-1 were obtained for BDE166 and BDE116,
155
respectively. When PBDE is changed from BDE166 to BDE116 and then to BDE47,
156
N decreases by 1 in each step. Both BDE166 and BDE116 have similar k, while
157
almost no debromination of BDE47 (N= 4) was observed under the same conditions.
158
The difference cannot be explained by N, although N is important for the
159
debromination of PBDEs. 8
ACS Paragon Plus Environment
Page 9 of 28
Environmental Science & Technology
160
2
(a) Br m
3
3' Br n
6
4
(b)
O 1' 2'
1
6'
5
4' 5'
1.0
c/c0
BDE47 BDE116 BDE166
0.5
0.0
(c)
0
6
t/min
9
12
0 k= k=
-1
0 .4 9
-2 -3 -4
0.0 99 k min -1 = 0. 15 m in -1
k = 0.0 k=0
35 min -1
.0 5 5
min -1
-1 -1 in m in = 0 .7 3 m k
ln(c/c0)
3
0
5
10
15
20
25
161
t/min
162
Figure 1. (a) Structure of PBDEs with the positions. The two phenyl rings are defined
163
as m and n rings with m= 1~5, n= 0~5, and m≥ n. (b) Photodegradation of BDE166,
164
BDE116, and BDE47 in a quartz vessel under irradiation at 365 nm with 4 mW cm-2.
165
(c) A comparison between the degradation of BDE116 (dash line) and BDE166 (solid
166
line) under irradiation at 365 nm with three intensities: 2 (squares), 4 (triangles), and
167
25 mW cm-2 (circles). 9
ACS Paragon Plus Environment
Environmental Science & Technology
168
169
Since it is not easy to perform photocatalytic debromination of all PBDEs under
170
the same reaction conditions, we obtained the relative k (kR) from the following
171
procedure. Firstly, BDE209 was selected as the reference with kR = 1, because
172
BDE209 is the most easily degradable PBDE. Then, some PBDEs such as BDE209
173
and BDE166 were used to simultaneously obtain k in one experiment under lower
174
irradiation intensity. The kR value of each PBDE was obtained from the ratio of k for
175
PBDE and BDE209. Secondly, BDE166 having smaller k was selected as the
176
reference and photocatalytic degradation of more difficult degradable PBDEs was
177
performed with higher irradiation intensity. This step was repeated several times, kR
178
for all BPDEs were obtained. To confirm independence of kR of PBDEs on the
179
irradiation intensity, the photocatalytic degradation of BDE166 and BDE116 was
180
examined at three irradiation intensities as shown in Figure 1c, where k were 0.055,
181
0.15, and 0.73 min-1 for BDE166 and 0.036, 0.099, and 0.49 min-1 for BDE116. The
182
ratios of k for BDE166 to BDE116 were constantly 1.5 at three irradiation intensities,
183
showing kR can be correctly obtained for PBDEs even at different irradiation
184
intensities. The photocatalytic reductive debromination of ten PBDEs was conducted
185
by dividing them into several pairs (Figure S2). The ratios of k were evaluated to be
186
kBDE209 = (3.4 ± 0.2)kBDE166, kBDE166 = (1.5 ± 0.1)kBDE116, kBDE166 = (11 ± 1)k2-CH3-BDE61,
187
k2-CH3-BDE61 = (47 ± 5)kBDE99, kBDE99 = (2.4 ± 0.2)kBDE30, kBDE75 = (2.3 ± 0.2)kBDE30,
188
kBDE30 = (11 ± 1)kBDE47, kBDE47 = (7.8 ± 0.8)kBDE7, and kBDE7 = (55 ± 10)kBDE15. By this
189
way, kR of ten PBDEs was obtained as illustrated in Table 1. 10
ACS Paragon Plus Environment
Page 10 of 28
Page 11 of 28
Environmental Science & Technology
190 191
Table 1. Relative rate constant (kR) of ten PBDEs. PBDEs
N
m, n
kR
BDE209
10
5, 5
1
BDE166
6
5, 1
(2.9 ± 0.2)×10-1
BDE116
5
5, 0
(2.0 ± 0.2)×10-1
2-CH3-BDE61
4
4(CH3), 0
(2.7 ± 0.3)×10-2
BDE99
5
3, 2
(5.7 ± 0.9)×10-4
BDE75
4
3, 1
(5.5 ± 1.1)×10-4
BDE30
3
3, 0
(2.4 ± 0.4)×10-4
BDE47
4
2, 2
(2.2 ± 0.4)×10-5
BDE7
2
2, 0
(2.5 ± 0.5)×10-6
BDE15
2
1, 1
(4.5 ± 1.3)×10-8
192 193
Among the ten PBDEs, BDE209 is the most highly brominated with N= 10 and
194
the most easily degradable with the largest kR, while BDE15 is the most lowly
195
brominated with N= 2 and has the smallest kR. As shown in Table 1, kR of BDE209 is
196
2×107 times larger than that of BDE15. This is in good agreement with reported
197
results showing that the reductive debromination of PBDEs becomes more difficult
198
with decreasing N.10,11,17,19 As shown in Table 1, kR of BDE7 is 55 times larger than
199
that of BDE15, although both BDE7 and BDE15 have N= 2; kR of BDE75 is 25 times
200
larger than that of BDE47, although both BDE75 and BDE47 have N= 4; kR of
201
BDE116 is 350 times larger than that of BDE99, although both BDE116 and BDE99
202
have N= 5. As shown in Figure 2a, no relation between logkR and N was observed,
203
suggesting that kR of PBDE is not dependent on N. Therefore, kR was compared for 11
ACS Paragon Plus Environment
Environmental Science & Technology
Page 12 of 28
204
PBDEs with same m value. For m= 3, BDE99, BDE75 and BDE30 show similar kR in
205
the order of 10-4. For m= 5, BDE209 has the largest kR, while BDE116 has the
206
smallest kR. The kR difference was only 5 times, while Br number on n ring (n)
207
decreases from 5 to 0. This strongly suggested that kR of PBDE is mainly dependent
208
on m, which is confirmed by the clear correlation between log kR and m as shown in
209
Figure 2b. (b)
(a)
(c)
1
kR
10-2
10-4
10-8
210
log kR = -4.54ELUMO -10.63
log kR = 1.53m - 8.04
10-6
R2= 0.96
2
R = 0.97
2
4
6 N
8
10
1
2
3 m
4
5
-1.0
-1.5 -2.0 ELUMO/eV
-2.5
211
Figure 2. Plots of kR of PBDEs and (a) N, (b) m, and (c) ELUMO. The data represented
212
by open triangles and squares were from ref. 17 and 11, respectively.
213 214
The relation between kR and m is also supported by the experimental results of kR
215
for a series of PBDEs. Granelli et al. reported kR by sodium borohydride for fifteen
216
PBDEs with N= 6~10 such as BDE209, BDE208, BDE207, BDE206, BDE204,
217
BDE203, BDE202, BDE201, BDE198, BDE196, BDE184, BDE183, BDE181,
218
BDE154, and BDE153, showing two exceptions from the relation between kR and N
219
(Table S1).17 For N= 8, the congers are divided into two groups: the first group 12
ACS Paragon Plus Environment
Page 13 of 28
Environmental Science & Technology
220
includes BDE204, BDE203, and BDE198, while the second group has BDE202,
221
BDE201, and BDE196. The kR values of the first group are 4 times larger than those
222
of the second group. BDE181 with N= 7 shows a much larger kR value than those of
223
PBDEs with N= 8 such as BDE202, BDE201, and BDE196. These exceptions from
224
the relation between kR and N are reasonably explained by the relation between kR and
225
m. For N= 8, the first and second groups have m= 5 and m=4, and the difference in
226
the m value causes that kR values of the first group are 4 times larger than those of the
227
second group. Similarly, BDE181 is a member of the N = 7 group with m = 5, while
228
BDE202, BDE201, and BDE196 have m= 4. It is certain that the kR value of an m = 5
229
member is considerably higher than that of the m = 4 members.
230
Weak dependence of ELUMO of PBDEs on the total Br number. From the
231
calculation as shown in the Experimental section, ELUMO of 209 PBDEs were obtained
232
(Table S2). For N= 10, deca-PBDE (BDE209) has the lowest ELUMO (about -2.3 eV).
233
Plots of ELUMO vs. N of PBDEs are shown in Figure 3a, indicating no clear relation
234
with a wide bell-mouth shape. As indicated by a dashed rectangle at the bottom of
235
Figure 3a, a narrow ELUMO range centered at -2.2 ± 0.1 eV covers twenty PBDEs with
236
N= 5~10. Similarly, 38 PBDEs with N= 4~8 are squeezed into a narrow ELUMO range
237
centered at -1.9 eV. Because of no clear correlation between ELUMO and N, we
238
consider m and n of PBDEs. Twenty PBDEs with ELUMO centered at -2.2 ± 0.1 eV
239
have m= 5, while 38 PBDEs with ELUMO centered at -1.9 eV have m= 4. This suggests
240
that ELUMO of PBDEs are dependent on m. As shown in Figure 3b, a plot of ELUMO vs.
241
m shows a clear relation. It should be noted that kR has a good correlation with ELUMO 13
ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 28
242
(Figure 2c), suggesting the relation between kR and m for all 209 PBDEs. This can be
243
explained by the frontier molecular orbital theory. The frontier orbitals of a PBDE
244
molecule include antibonding orbital of C-Br (σC-Br*) and antibonding orbital of C=C
245
(πC=C*). Since C-Br bond is very long (about 1.9Å) and the strong stabilization the
246
phenyl ring, the energy of σC-Br* might be lower than that of πC=C*. According to
247
Schrodinger’s equation, the molecular orbital is denoted by the linear combination of
248
atomic orbitals. Therefore, the phenyl ring with more Br substitutions has more
249
contribution of σC-Br* and has a lower energy. However, two phenyl rings of PBDE
250
are separated by an ether bond, the energy of antibonding orbital of C-O (σC-O*) is
251
much higher than σC-Br* and πC=C*, and therefore, σC-O* has very little contribution to
252
LUMO, to have no combination of “LUMO” of two phenyl rings. Thus, E“LUMO” of m
253
ring is extremely close to E“LUMO” of PBDE, while the “LUMO” of n ring becomes a
254
higher unoccupied molecular orbital. (b)
-0.8
-0.8
-1.2
-1.2
ELUMO/eV
ELUMO/eV
(a)
-1.6 -2.0 -2.4
255 256
-1.6 -2.0
1
2
3
4
5
6
7
8
9
10
-2.4
1
N
2
3
4
5
m
Figure 3. Dependence of ELUMO of 209 PBDEs on (a) N and (b) m.
257 258
The relation between kR and m. As discussed above, the Br-richer phenyl ring
259
is more reactive than the other ring for all PBDEs, because this ring mainly 14
ACS Paragon Plus Environment
Page 15 of 28
Environmental Science & Technology
260
determines kR of the reductive debromination. To understand the relation between kR
261
and m, we analyzed the LUMO distributions of 23 PBDEs with m= 1~5 as shown in
262
Figure 4. A large change of LUMO from πC=C* to σC-Br* is observed as m increases
263
from 1, 2 to 4, 5, consistent with above demonstrated explanation that more σC-Br*
264
character would be involved in the LUMO as the Br substitution increases.
265
Furthermore, BDE209 has two phenyl rings fully substituted by Br (N= 10, m= n= 5),
266
and LUMO is distributed on both rings. All other PBDEs with m= 5 have LUMO
267
distribution on m ring. A similar LUMO distribution on m ring is observed for PBDEs
268
with m= 4, while LUMO is distributed on both rings for BDE202 and BDE201 with
269
m= n= 4. For m= 3, LUMO is mainly distributed on m ring, but also on n ring with
270
increasing n for 0 ~ 3. When m= 2 or 1, LUMO is distributed on two rings, consistent
271
with our results as mentioned above.
15
ACS Paragon Plus Environment
Environmental Science & Technology
209 (5, 5)
203 (5, 3)
202 (4, 4)
154 (3, 3)
47 (2, 2)
272
208 (5, 4)
207 (5, 4)
206 (5, 4)
181 (5, 2)
198 (5, 3)
201 (4, 4)
153 (3, 3)
166 (5, 1)
184 (4, 3)
99 (3, 2)
7 (2, 0)
75 (3, 1)
15 (1, 1)
Page 16 of 28
204 (5, 3)
116 (5, 0)
183 (4, 3)
30 (3, 0)
3 (1, 0)
273
Figure 4. LUMO distributions of 23 PBDEs. The number is the serial number of
274
PBDEs (for example, 209 denotes BDE209), and the numbers in the parenthesis is m,
275
n.
276 277
During the electron-transfer initiated reductive debromination of PBDEs, an
278
electron is firstly entered to the LUMO of PBDEs. To quantify the selectivity of
279
LUMO, the atomic dipole moments corrected Hirshfeld population (ADCH) on three
280
components such as m ring, O, and n ring of PBDEs and PBDEs.- with the same
281
geometry as the neutral PBDEs are calculated by using Multiwfn software. 31,32 The 16
ACS Paragon Plus Environment
Page 17 of 28
Environmental Science & Technology
282
ADCH differences of three components between PBDE and PBDE.- were used to
283
simulate the distribution of LUMO. As shown in Table S3, LUMO is mainly localized
284
on m ring for all PBDEs with m> n. When an electron is attached to a neutral
285
molecule, the charge increase on m ring is in the range from 0.776 (BDE99) to 0.917
286
(BDE116), but that on n ring is less than 0.1. For PBDEs with m= n such as BDE209,
287
BDE153, BDE47, and BDE15, the charge increases on two rings are similar. Both
288
LUMO localization and ADCH differences show preferential reduction of m ring,
289
which is responsible for the relation between kR and m.
290
Because both kR and LUMO distribution of PBDEs show clear dependences on
291
m, we conclude that m is important but n is not for reductive debromination. As
292
shown in Table 2, polybrominated phenols (C6HxBr5-xOH), polybrominated phenyl
293
methyl ethers (C6HxBr5-xOCH3), and polybrominated phenyl trifluoromethyl ethers
294
(C6HxBr5-xOCF3), with the same chromophore (i.e. C6Br5O-), have similar ELUMO.
295
ELUMO of BDE116 (5, 0) is calculated to be -2.209 eV, while ELUMO increases to
296
-2.151 eV for phenyl ring (n= 0) change to an electron-donating methyl group (CH3),
297
and decreases to -2.314 eV for phenyl ring (n= 0) change to an electron-withdrawing
298
trifluoromethyl group (CF3). The same tendency is observed for all PBDEs containing
299
different m=1~5. Increasing electron-withdrawing group leads to lower ELUMO and
300
larger kR according to Figure 2c. However, the influence of the substituted group on
301
ELUMO is often less than 0.2 eV. In contrast, ELUMO increases by 0.2~0.5 eV for each
302
addition of a Br on m ring. Therefore, m has stronger influence on ELUMO than the
303
substituted groups on polybrominated phenoxyl analogues where ELUMO are mainly 17
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 28
304
dependent on m ring. With increasing m, the eletrophilicity of m ring increases
305
through inductive and conjugative effects. The substituted group influences slightly
306
the eletrophilicity of m ring by indirect inductive effect via ether bond. As the result,
307
even CF3 has little influence, smaller than 0.2 eV, on ELUMO. It should be mentioned
308
that the change of ELUMO for PBDEs with m= 5 induced by Br number substituted on
309
n ring (n= 1~4) is less than 0.1 eV. Similarly, the change of ELUMO for PBDEs with
310
m= 4 induced by Br number substituted on n ring (n= 1~3) is small, and so on.
311
Table 2. ELUMO of bromophenoxyl analogues with different substituted groups
312
(bromophenoxyl-Y) Bromophenoxyl
ΔELUMO, 1 b /eV
ELUMO of bromophenoxyl-Y/eV
ΔELUMO,maxa
-C6H5
-CF3
-H
-CH3
/eV
-2.209
-2.314
-2.161
-2.151
0.163
-1.876
-1.985
-1.870
-1.791
0.194
-1.436
-1.565
-1.372
-1.360
0.205
-1.123
-1.312
-1.107
-1.051
0.261
-0.977
-1.093
-0.859
-0.845
0.248
0.31 ~ 0.44
0.25 ~ 0.42
0.27 ~ 0.50
0.31 ~ 0.42
The maximum difference between the different columns in the same row.
313
a
314
b
315
row.
The value differences between two adjacent upper and down rows except for the last
316
As mentioned above, m of PBDEs is important for kR and ELUMO, while n has
317
less influence. As shown in Table S3, the m ring acts as the electron acceptor in the 18
ACS Paragon Plus Environment
Page 19 of 28
Environmental Science & Technology
318
reduction process, causing the debromination on m ring. To prove this mechanism, we
319
analyzed
320
debromination of BDE99 (m, n= 3, 2) and BDE166 (5, 1). As shown in Figure 5a, the
321
debromination of parent BDE99 produces five tetra-BDEs as the intermediates, three
322
of which (BDE47, BDE49, and BDE66) are resulted from the debromination on m
323
ring, while other two (BDE48 and BDE74) are from the debromination on n ring.
324
According to their relative abundance, the average concentrations of former three
325
intermediates are about three times higher than those of latter two as shown in Figure
326
S3. This means that total amount of intermediates from m-ring-debromination are
327
more than 80% for the debromination of BDE99, demonstrating clearly that the
328
debromination occurs mainly on m ring.
the
intermediates
generated
the
CuO/TiO2
photocatalytic
(b) 74 Br
66 Br
49 Br
Br 48 Br 47
20
329
48 49
47 74 66
3 Br
22
24
26
114 Br
117
99
O
Relative abundance
Relative abundance
(a)
from
24
28
115 Br
O
117 Br
Br
Br 116
Br
115
4 Br 114
26
28
166
30
32
Retention time/min
Retention time/min
330
Figure 5. GC-MS chromatograms of the solutions after the photodegradation of (a)
331
BDE99 and (b) BDE166 on CuO/TiO2 under Ar atmosphere. The numbers represent
332
the serial number of the corresponding PBDE. For example, the number 47 means
333
BDE47. The numbers beside the chemical structures of BDE99 and BDE166
334
represent the serial number of the corresponding PBDE generated from eliminating Br
335
at the position. 19
ACS Paragon Plus Environment
Environmental Science & Technology
336
The m-ring debromination process becomes much clear for BDE166. The two
337
phenyl rings of BDE166 have a large difference of Br number (5 against 1). As shown
338
in Figure 5b, the debromination of BDE166 gives three penta-BDE products
339
(BDE114, BDE115, and BDE117) together with several tetra-BDE products. If the
340
debromination of the n ring (n= 1) could occur competitively, BDE116 would be
341
obtained. However, it was not detected even after a more careful analysis to eliminate
342
any possible interference from BDE115 (Figure S3b).
343
Environmental Implications. PBDEs are ubiquitous pollutants being considered as
344
the first generation of “emerging contaminants”33. The present study provides an
345
understanding of basic environmental chemistry of these pollutants in terms of the
346
relation between kR and m. When m > n, the m ring is more reactive than the n ring of
347
a PBDE in the electron-transfer initiated reductive debromination. Because the
348
debromination preferably occurs on the Br-rich phenyl ring, the debromination of
349
BDE209 proceeds in a stepwise mechanism, i.e., N= 10 to 9, 9 to 8, and then 8 to 7,
350
and so on. This indicates that the reductive debromination proceeds in a “cross
351
debromination” mechanism in which two phenyl rings in BDE209 cause the
352
debromination to give intermediates with two phenyl rings having same Br ((m, n)=
353
(5, 5) to (4, 5), (4, 5) to (4, 4), then (4, 4) to (3, 4), and so on). The present work
354
experimentally confirms that kR of BDE209 (5, 5) decreases by a magnitude of 107
355
compared with kR of BDE15 (1, 1). Such a large decrease in the reactivity for PBDEs
356
with small N indicates that the reductive debromination is not suitable for the
357
complete debromination of PBDEs when their debromination is governed by 20
ACS Paragon Plus Environment
Page 20 of 28
Page 21 of 28
Environmental Science & Technology
358
reductive debromination involving electron transfer. This makes us easily understand
359
that reductive debromination of highly brominated PBDEs was observed to be
360
stopped at the accumulation of low-brominated PBDEs in the anaerobic
361
biotransformation34 and ZVI transformation35. Therefore, we have to develop
362
alternative methods for achieving complete debromination of PBDEs. Indeed,
363
metal-induced catalytic debromination of PBDEs in the presence of hydrogen sources
364
have been reported recently,36-38 which could be a candidate for the complete
365
debromination of PBDEs.
366
ASSOCIATED CONTENT
367
Supporting Information. Synthesis of PBDEs (Text S1 and Scheme S1); 1H NMR
368
spectrums of synthesized PBDEs (Figure S1); Photocatalytic reductive degradation of
369
10 congeners of PBDEs (Figure S2); GC-MS chromatograms of the solutions
370
obtained from the photodegradation of (a) BDE99 and (b) BDE 166 on CuO/TiO2
371
under Ar atmosphere (Figure S3); Relative rate of reductive debromination of PBDEs
372
by sodium borohydride (Table S1); ELUMO values and substituted patterns (m and n)
373
of all 209 PBDE congeners (Table S2); ADCH populations of 23 PBDEs and their
374
vertical anion radicals who are divided into three fragments: m ring, O atom and n
375
ring ( Table S3).
376
AUTHOR INFORMATION
377
Corresponding Authors
378
* Tel/Fax: +86 27 87543632 (L. Zhu), +86 27 67843990 (H. Tang); E-mail: 21
ACS Paragon Plus Environment
Environmental Science & Technology
379
[email protected] (L. Zhu),
[email protected] (H. Tang).
380
Notes
381
The authors declare no competing financial interest.
382
ACKNOWLEDGMENTS
383
The authors acknowledge the financial supports from the National Natural Science
384
Foundation of China (Grants Nos. 21707170 and 21777194). The authors would
385
thank all the reviewers for their greatly helpful suggestions and discussions.
386
REFERENCES
387
(1) Besis, A.; Samara, C. Polybrominated diphenyl ethers (PBDEs) in the indoor and
388
outdoor environments - A review on occurrence and human exposure. Environ.
389
Pollut. 2012, 169, 217-229.
390
(2) Abbasi, G.; Buser, A. M.; Soehl, A.; Murray, M. W.; Diamond, M. L. Stocks and
391
flows of PBDEs in products from use to waste in the US and Canada from 1970 to
392
2020. Environ. Sci. Technol. 2015, 49, 1521-1528.
393
(3) Li, Y.; Qiao, L.; Ren, N.; Sverko, E.; Mackay, D.; Macdonald, R. W. Macdonald.
394
Decabrominated diphenyl ethers (BDE-209) in Chinese and global air: levels,
395
gas/particle partitioning, and long-range transport: is long-range transport of BDE-209
396
really governed by the movement of particles? Environ. Sci. Technol. 2017, 51,
397
1035-1042.
398
(4) Li, W.; Ma, W.; Jia, H.; Hong, W.; Moon, H.; Nakata, H.; Minh, N. H.; Sinha, R.
399
K.; Chi, K. H.; Kannan, K.; Sverko, E.; Li, Y. Polybrominated diphenyl ethers
400
(PBDEs) in surface soils across five Asian countries: levels, spatial distribution, and 22
ACS Paragon Plus Environment
Page 22 of 28
Page 23 of 28
Environmental Science & Technology
401
source contribution. Environ. Sci. Technol. 2016, 50, 12779-12788.
402
(5) Khairy, M. A.; Lohmann, R. Using polyethylene passive samplers to study the
403
partitioning and fluxes of polybrominated diphenyl ethers in an urban river. Environ.
404
Sci. Technol. 2017, 51, 9062-9071.
405
(6) Leijs, M. M.; Tusscher, G. W.; Olie, K.; Teunenbroek, T.; Aalderen, W. M.;
406
Voogt, P.; Vulsma, T.; Bartonova, A.; Krauss, M. K.; Mosoiu, C.; Riojas-Rodriguez,
407
H.; Calamandrei, G.; Koppe, J. G. Thyroid hormone metabolism and environmental
408
chemical exposure. Environ. Health. 2012, 11, S10.
409
(7) Kiciński, M.; Viaene, M. K.; Hond, E. D.; Schoeters, G.; Covaci, A.; Dirtu, A. C.;
410
Nelen, V.; Bruckers, L.; Croes, K.; Sioen, I.; Baeyens, W.; Larebeke, N. V.; Nawrot,
411
T. S. Neurobehavioral function and low-level exposure to brominated flame retardants
412
in adolescents: A cross-sectional study. Environ. Health. 2012, 11, 86.
413
(8) Huang, A.; Wang, N.; Lei, M.; Zhu, L.; Zhang, Y.; Lin, Z.; Yin, D.; Tang, H.
414
Efficient oxidative debromination of decabromodiphenyl ether by TiO2-mediated
415
photocatalysis in aqueous environment. Environ. Sci. Technol. 2013, 47, 518-525.
416
(9) Shi, J.; Qu. R.; Feng, M.; Wang, X.; Wang, L.; Yang, S.; Wang. Z. Oxidative
417
degradation of decabromodiphenyl ether (BDE 209) by potassium permanganate:
418
reaction pathways, kinetics, and mechanisms assisted by density functional theory
419
calculations. Environ. Sci. Technol. 2015, 49, 4209-4217.
420
(10) Keum,Y.; Li, Q.; Reductive debromination of polybrominated diphenyl ethers by
421
zerovalent iron. Environ. Sci. Technol. 2005, 39, 2280-2286.
422
(11) Zhuang, Y.; Ahn, S.; Luthy, R. G. Debromination of polybrominated diphenyl 23
ACS Paragon Plus Environment
Environmental Science & Technology
423
ethers by nanoscale zerovalent iron: pathways, kinetics, and reactivity. Environ. Sci.
424
Technol. 2010, 44, 8236-8242.
425
(12) Sun, C.; Zhao, D.; Chen, C.; Ma, W.; Zhao J. TiO2-mediated photocatalytic
426
debromination of decabromodiphenyl ether: kinetics and intermediates. Environ. Sci.
427
Technol. 2009, 43, 157-162.
428
(13) Lei, M.; Wang, N.; Zhu, L.; Tang, H. Peculiar and rapid photocatalytic
429
degradation of tetrabromodiphenyl ethers over Ag/TiO2 induced by interaction
430
between silver nanoparticles and Br in the target. Chemosphere 2016, 150, 536-544.
431
(14) Liu, Y.; Chen, S.; Quan, X.; Fan, X.; Zhao, H.; Zhao, Q.; Yu, H. Nitrogen-doped
432
nanodiamond rod array electrode with superior performance for elecroreductive
433
debromination of polybrominated diphenyl ethers. Appl. Catal. B: Environ. 2014,
434
154-155, 206-212.
435
(15) Chen, J.; Wang, C.; Pan, Y.; Farzana, S. S.; Tam, N. F. Biochar accelerates
436
microbial reductive debromination of 2,2’,4,4’-tetrabromodiphenyl ether (BDE-47) in
437
anaerobic mangrove sediments. J. Hazard. Mater. 2018, 341, 177-186.
438
(16) Castro-Jiménz, J.; Barhoumi, B.; Paluselli, A.; Tedetti, M.; Jiménz, B.;
439
Muñoz-Arnanz, J.; Wortham, H.; Ridha, Driss, M.; Sempéré, R. Occurrence, loading,
440
and exposure of atmospheric particle-bound POPs at the African and European edges
441
of the western Mediterranean sea. Environ. Sci. Technol. 2017, 51, 13180-13189.
442
(17) Granelli, L.; Eriksson, J.; Bergman. Å. Sodium borohydride reduction of
443
individual polybrominated diphenyl ethers. Chemosphere 2012, 86, 1008-1012.
444
(18) Usenko, C. Y.; Robinson, E. M.; Usenko, S.; Brooks, B. W.; Bruce, E. D. PBDE 24
ACS Paragon Plus Environment
Page 24 of 28
Page 25 of 28
Environmental Science & Technology
445
developmental effects on embryonic zebrafish. Environ. Toxicol. Chem. 2011, 30,
446
1865-1872.
447
(19) Santos, M. S.; Alves, A.; Madeira, L. M. Chemical and photochemical
448
degradation of polybrominated diphenyl ethers in liquid systems – A review. Water
449
Res. 2016, 88, 39-59.
450
(20) Isse, A. A.; Mussini, P. R.; Gennaro, A. New insights into electrocatalysis and
451
dissociative electron transfer mechanisms: the case of aromatic bromides. J. Phys.
452
Chem. C 2009, 113, 14983-14992.
453
(21) Rotko, G.; Romańczyk, P. P.; Andryianau, G.; Kurek, S. S. Stepwise and
454
concerted dissociative electron transfer onto a σ*-type orbital in polybrominated
455
aromatics. Electrochem. Commun. 2014, 43, 117-120.
456
(22) Houman, A. Electron transfer initiated reactions: bond formation and bond
457
dissociation. Chem. Rev. 2008, 108, 2180-2237.
458
(23) Lei, M.; Wang, N.; Zhu, L; Zhou, Q.; Nie, G.; Tang, H. Photocatalytic reductive
459
degradation of polybrominated diphenyl ethers on CuO/TiO2 nanocomposites: A
460
mechanism based on the switching of photocatalytic reduction potential being
461
controlled by the valence state of copper. Appl. Catal. B: Environ. 2016, 182,
462
414-423.
463
(24) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Frisch, M. J.; Trucks, G. W.;
464
Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.;
465
Barone, V.; Mennucci, B.; Petersson, G.A.; Nakatsuji, H.; Caricato, M.; Li, X.;
466
Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; 25
ACS Paragon Plus Environment
Environmental Science & Technology
467
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda,
468
Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A.; Peralta, J. E.; Ogliaro, F.;
469
Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi,
470
R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi,
471
J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken,
472
V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A.
473
J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.;
474
Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.;
475
Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.;
476
2009. Gaussian 09, Revision A. 02. Gaussian Inc., Wallingford, CT.
477
(25) Becke, A. D. Density-functional thermochemistry. Ⅲ . The role of exact
478
exchange. J. Chem. Phys. 1993, 98, 5648-5652.
479
(26) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti
480
correlation-energy formula into a function of the electron density. Phys. Rev. B 1988,
481
37, 785-789.
482
(27) Wang, Z.; Zeng, X.; Zhai, Z. Prediction of supercooled liquid vapor pressures
483
and n-octanol/air partition coefficients for polybrominated diphenyl ethers by means
484
of molecular descriptors from DFT method. Sci. Total. Environ. 2008, 389, 296-305.
485
(28) Wang, X.; Wang, Y.; Chen, J.; Ma, Y.; Zhou, J.; Fu, Z. Computation
486
toxicological investigation on the mechanism and pathways of xenobiotics
487
metabolized by cytochrome P450: a case of BDE-47. Environ. Sci. Technol. 2012, 46,
488
5126-5133. 26
ACS Paragon Plus Environment
Page 26 of 28
Page 27 of 28
Environmental Science & Technology
489
(29) Wei, X.; Yuan, Q.; Serge, B.; Xu, T.; Ma, G.; Yu, H. In silico investigation of
490
gas/particle partitioning equilibrium of polybrominated diphenyl ethers (PBDEs).
491
Chemosphere, 2017, 188, 110-118.
492
(30) Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self-consistent molecular orbital
493
method. XII. Further extensions of gaussion-type basis sets for use in molecular
494
orbital studies of organic molecules. J. Chem. Phys. 1972, 56, 2257-2261.
495
(31) Lu, T.; Chen, F. Atomic Dipole Moment Corrected Hirshfeld Population Method.
496
J. Theor. Comput. Chem. 2012, 11, 163-183.
497
(32) Lu, T.; Chen, F.; Multiwfn: A Multifunctional Wavefunction Analyzer. J.
498
Comput. Chem. 2012, 33, 580-592.
499
(33) Castro-Jiménez, J.; Eisenreich, S. J.; Ghiani, M.; Mariani, G.; Skejo, H.; Umlauf,
500
G.; Wollgast, J.; Zaldívar, J. M.; Berrojalbiz, N.; Reuter, H. I.; Dachs, J. Atmospheric
501
occurrence and deposition of polychlorinated dibenzo-p-dioxins and dibenzofurans
502
(PCDD/Fs) in the open Mediterranean Sea. Environ. Sci. Technol. 2010, 44,
503
5456-5463.
504
(34) Stapleton, H. M.; Brazil, B.; Holbrook, R. D.; Mitchelmore, C. L. In vivo and in
505
vitro debromination of decabromodiphenyl ether (BDE 209) by juvenile rainbow trout
506
and common carp. Environ. Sci. Technol. 2006, 40, 4653-4658.
507
(35) Li, A.; Tai, C.; Zhao, Z.; Wang, Y.; Zhang, Q.; Jiang, G.; Hu, J. Debromination
508
of decabrominated diphenyl ether by resin-bound iron nanoparticles. Environ. Sci.
509
Technol. 2007, 41, 6841-6846.
510
(36) Lei, M.; Guo, S.; Wang, Z.; Zhu, L.; Tang, H. Ultrarapid and deep debromination 27
ACS Paragon Plus Environment
Environmental Science & Technology
511
of tetrabromodiphenyl ether over noble-metal-free Cu/TiO2 nanocomposites under
512
mild conditions. Environ. Sci. Technol. 2018, 52, 11743-11751.
513
(37) Zahran, E. M.; Bedford, N. M.; Nguyen, M. A.; Chang, Y.; Guiton, B. S.; Naik,
514
R. R.; Bachas, L. G.; Knecht, M. R. Light-activated tandem catalysis driven by
515
multicomponent nanomaterials. J. Am. Chem. Soc. 2014, 136, 32-35.
516
(38) Wang, R.; Lu, G.; Lin, H.; Huang, K.; Tang, T.; Xue, X.; Yang, X.; Yin, H.;
517
Dang, Z. Relative roles of H-atom transfer and electron transfer in the debromination
518
of polybrominated diphenyl ethers by palladized nanoscale zerovalent iron. Environ.
519
Pollut. 2017, 222, 331-337.
28
ACS Paragon Plus Environment
Page 28 of 28