Subscriber access provided by BOSTON UNIV
Article
Rapid Destruction of Tetrabromobisphenol A by Iron(III)-Tetraamidomacrocyclic Ligand/Layered Double Hydroxide Composite/H2O2 system Chao Wang, Juan Gao, and Cheng Gu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b04294 • Publication Date (Web): 02 Dec 2016 Downloaded from http://pubs.acs.org on December 2, 2016
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 39
Environmental Science & Technology
1
Rapid Destruction of Tetrabromobisphenol A by
2
Iron(III)-Tetraamidomacrocyclic Ligand/Layered Double
3
Hydroxide Composite/H2O2 System Chao Wang1, Juan Gao2, Cheng Gu*1
4 5 1
6
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210023, P.R. China
7 8 9
2
Key Laboratory of Soil Environment and Pollution Remediation, Institute of Soil Science, Chinese Academy of Sciences, Nanjing, Jiangsu 210008, P. R. China
10 11 12
*To whom correspondence should be addressed
13
Cheng Gu
14
Professor
15
School of the Environment
16
Nanjing University
17
Nanjing, Jiangsu, 210023
18
P. R. China
19
Phone/Fax: +86-25-89680636
20
E-mail:
[email protected] 21 22 1
ACS Paragon Plus Environment
Environmental Science & Technology
Page 2 of 39
23
Abstract
24
Iron(III)-tetraamidomacrocyclic ligand (Fe(III)-TAML) activators have received
25
widespread attentions for their abilities to activate hydrogen peroxide to oxidize many
26
organic pollutants. In this study, Fe(III)-TAML/layered double hydroxide (LDH)
27
composite was developed by intercalating Fe(III)-TAML into the interlayer of LDH.
28
Electrostatic interaction and hydrogen bonding might account for the adsorption of
29
Fe(III)-TAML on LDH. The newly synthesized Fe(III)-TAML/LDH composite
30
showed
31
tetrabromobisphenol A (TBBPA) in the presence of hydrogen peroxide, which can be
32
fully degraded within 20 s, and the degradation rate increased up to 8 times compared
33
to free Fe(III)-TAML. In addition, the toxicity of the system was significantly reduced
34
after the reaction. The higher reactivity of Fe(III)-TAML/LDH system is attributed to
35
the enhanced adsorption of TBBPA on LDH, which could increase the contact
36
possibility between Fe(III)-TAML and TBBPA. Based on the analysis of reaction
37
intermediates, β-scission at the middle carbon atom and C-Br bond cleavage in phenyl
38
ring of TBBPA were involved in the degradation process. Furthermore, our results
39
demonstrated that the Fe(III)-TAML/LDH composite can be reused several times,
40
which could lower the overall cost for environmental implication and render
41
Fe(III)-TAML/LDH as the potential environmental-friendly catalyst for future
42
wastewater treatment under mild reaction conditions.
43
Keywords:
44
hydroxide, tetrabromobisphenol A, waste remediation, green catalyst
superior
reactivity
as
indicated
iron(III)-tetraamidomacrocyclic
by
ligand
efficient
decomposition
activator,
45 46 47 2
ACS Paragon Plus Environment
layered
of
double
Page 3 of 39
Environmental Science & Technology
48 49 50 51
52
TOC Art:
53 54 55 56 57 58 59 60 61 62 63 64 65 66
3
ACS Paragon Plus Environment
Environmental Science & Technology
67
Introduction
68
Tetrabromobisphenol A [4,4’-isopropylidenebis (2,6-dibromophenol), TBBPA],
69
accounting for nearly 60% of the total brominated flame retardant (BFR) market
70
share1,2, is primarily used as the flame retardant in printed circuit boards and also as
71
additives in many consumer products, e.g. textiles and plastics3-5. Recent studies
72
showed that TBBPA has been released into the environment and frequently detected in
73
various environmental matrices, including air, soil, water and sediment. Zweidinger et
74
al.6 reported a level of 1.8 µg TBBPA m-3 in the air near the production site. Studies
75
found that the concentrations of TBBPA in contaminated soil and sediment ranged
76
from 0.5 to 140 µg kg-1 (dry weight) and from 2 to 150 µg kg-1 (dry weight),
77
respectively7,8. It was also reported that the TBBPA concentration could reach 4.87 µg
78
L-1 in water from Lake Chaohu in China9. Based on the biodegradation results of
79
TBBPA in different environmental media, including soil, river sediment and water, the
80
estimated half-life of TBBPA in the environment was ~2 months3. TBBPA was even
81
detected in serum, adipose tissue and breast milk samples of humans at the level of ng
82
kg-1 and has exerted severe health effect10-12. In vitro studies had confirmed that
83
TBBPA could exhibit biological activity, including the disruption of thyroid
84
hormone13,14 and high lethal toxicity towards cerebellar granule cells15. The research
85
by Sjödin et al.16 showed a close correlation between air concentrations of TBBPA
86
and human serum levels, indicating that occupational exposure might be the important
87
route for human exposure.
88
Considering the recalcitrant nature of TBBPA in the environment and its potential 4
ACS Paragon Plus Environment
Page 4 of 39
Page 5 of 39
Environmental Science & Technology
89
risk, it is greatly needed to develop efficient techniques to eliminate TBBPA. It is
90
reported that TBBPA cannot be effectively degraded by traditional methods, such as
91
δ-MnO217, UV irradiation18 and photosensitized oxidation19, and the toxic byproducts,
92
e.g. brominated isopropylphenol derivatives were generated during the degradation
93
process. Recently, some advanced strategies were tested to fully degrade TBBPA.
94
Results from Ding et al. showed that TBBPA (18.4 µM) could be degraded in 30 min
95
to achieve 99% of TBBPA degradation, 67% of debromination and 56% of TOC
96
removal by sulfate radicals, which were produced from 1.5 mM peroxymonosulfate
97
catalyzed by 0.1 g L-1 CuFe2O4 nanoparticles20. However, during the reaction, Cu2+
98
was released and the concentration of Cu2+ could reach 1.3 mg L-1, that would
99
potentially lead to secondary pollution20. Ozonation was also utilized to achieve rapid
100
removal of TBBPA, but the generation of carcinogenic bromate ions retarded its
101
application21. Therefore, processes that can achieve rapid destruction and don’t
102
present any toxicity concerns during the degradation of TBBPA have still remained
103
elusive.
104
Iron(III)-tetraamidomacrocyclic ligand (Fe(III)-TAML, structure shown in Figure 1)
105
activators have been classified as the new family of green catalysts. Prior studies
106
showed that Fe(III)-TAML/H2O2 can completely mineralize chlorophenols in a few
107
minutes22,23. Further research indicated that Fe(III)-TAML activator/H2O2 systems
108
also exhibited high reactivity for degradation of fenitrothion and other
109
organophosphorus pesticides24, dibenzothiophene25, natural and synthetic estrogens26,
110
azo- and ruthenium-dyes27,28, pharmaceutical ingredient (e.g., sertraline29), and 5
ACS Paragon Plus Environment
Environmental Science & Technology
111
explosives (e.g., 2,4,6-trinitrotoluene and 1,3,5-trinitrobenzene30). Although the
112
reactivity of Fe(III)-TAML catalyst is strongly pH dependent, it decreases
113
significantly under neutral or acidic conditions31,32. However, even at pH 7,
114
Fe(III)-TAML still showed high reactivity compared to other advanced oxidation
115
techniques33. Previous studies have shown that immobilization of metalloporphyrin
116
molecules onto various supports, including graphene, single-walled carbon nanotubes
117
and natural montmorillonite clay mineral would significantly increase the stability,
118
catalytic activity and reusability34-36. However, only limited studies were conducted
119
for immobilization of Fe(III)-TAML activators. It was reported that the deposition of
120
Fe(III)-TAML on electrode surface would significantly increase its turnover
121
numbers37. So far, there is still lack of information to use immobilized Fe(III)-TAML
122
for environmental applications.
123
The layered double hydroxides (LDHs), also called anionic clays, consist of
124
brucite-like layers. Due to isomorphic substitutions in LDH structure, LDHs possess
125
positive charges, which are commonly compensated by exchangeable anions, e.g., Cl-,
126
CO32- et al.38,39 With the pronounced versatility and low cost, LDHs have been widely
127
utilized as the supports for synthesis of heterogeneous catalysts40-42 and for drug
128
delivery43,44. Given that Fe(III)-TAML is negatively charged, we hypothesize that
129
Fe(III)-TAML can be intercalated into the interlayer of LDH to form
130
Fe(III)-TAML/LDH composite, which would increase the stability and catalytic
131
reactivity of Fe(III)-TAML.
132
The objective of this study was to develop a novel material by immobilization of 6
ACS Paragon Plus Environment
Page 6 of 39
Page 7 of 39
Environmental Science & Technology
133
Fe(III)-TAML on the surface of LDH to effectively degrade TBBPA. Our results
134
showed that Fe(III)-TAML can form strong interaction with LDH through
135
electrostatic attraction and hydrogen bonding, which was also supported from our
136
spectroscopic analysis. The newly formed material exhibited extraordinary reactivity
137
as indicated by the degradation of TBBPA, the degradation rate increased up to 8
138
times compared to free Fe(III)-TAML, especially at lower pH. More importantly, after
139
reaction, less toxic byproducts are formed. The experimental results further
140
demonstrated that Fe(III)-TAML/LDH composite can be reused, that would extend its
141
potential application. The role of LDH in the reaction, in addition to protecting
142
Fe(III)-TAML from suicidal inactivation, is to provide a unique matrix to increase the
143
contact possibility between Fe(III)-TAML and TBBPA.
144
Materials and methods
145
Chemicals. Details of chemicals are listed in Text S1.
146
Preparation and characterization of LDH. Detailed information on preparation
147
and characterization of LDH is provided in Text S2.
148
Preparation and characterization of Fe(III)-TAML/LDH composite material.
149
The details of preparation and characterization of Fe(III)-TAML/LDH are given in
150
Text S3.
151
Degradation of TBBPA by Fe(III)-TAML and Fe(III)-TAML/LDH composite.
152
Degradation experiment was conducted in a 200 mL Erlenmeyer flask. For each
153
experiment, TBBPA stock solution was added to Fe(III)-TAML/LDH suspension with
154
the initial concentrations for TBBPA and Fe(III)-TAML of 18.4 and 1.0 µM, 7
ACS Paragon Plus Environment
Environmental Science & Technology
155
respectively. To investigate the pH effect on the reactivity of Fe(III)-TAML/LDH, the
156
degradation experiment was conducted at pH 8, 9 and 10, respectively. The reaction
157
was initiated by addition of 30 µL of H2O2 (30%) stock solution with the molar ratio
158
of H2O2/Fe(III)-TAML = 110/1. To avoid any pH fluctuation, both the pHs of
159
TBBPA/Fe(III)-TAML/LDH solution and H2O2 stock solution were pre-adjusted to
160
desired pH by addition of 0.1 M NaOH or 0.1 M HClO4 solution. The pH was stable
161
and the variation was within ±0.1 during the 120 s reaction period. At predetermined
162
time interval, an aliquot of 500 µL sample was withdrawn and mixed with 5 µL
163
concentrated perchloric acid (7 M) to terminate the reaction by demetalation of
164
Fe(III)-TAML45, and 1 mL methanol was added to extract TBBPA adsorbed by
165
Fe(III)-TAML/LDH composite. Then the sample was filtered through a 0.22 µm
166
polytetrafluoroethylene syringe filter and the concentrations of TBBPA were analyzed
167
using a high pressure liquid chromatography (HPLC) (Waters Alliance 2695, Milford,
168
MA). The released Br- concentrations during the reaction were determined by an ion
169
chromatography (IC) (Dionex Co., USA). In addition, the total organic carbon (TOC)
170
contents before and after the reaction were measured using a TOC-5000A analyzer
171
(Shimadzu, Japan) to quantify the mineralization of TBBPA during the reaction. For
172
comparison, similar experiment was conducted for degradation of TBBPA by 1.0 µM
173
free Fe(III)-TAML. The adsorption of TBBPA on Fe(III)-TAML/LDH was also
174
studied at different pHs. The adsorption kinetic experiment was conducted in a 0.22
175
µm polytetrafluoroethylene syringe filter to ensure the quick termination of the
176
adsorption process. The control experiments with full recovery (> 98%) demonstrated 8
ACS Paragon Plus Environment
Page 8 of 39
Page 9 of 39
Environmental Science & Technology
177
that either Fe(III)-TAML/LDH or LDH+H2O2 showed no effect on the degradation of
178
TBBPA (Fig. S3). The information for quantification of TBBPA and bromide ion,
179
mass spectrometric measurement and minimization of time lag in degradation
180
experiment is summarized in Text S4, S5 and S6, respectively.
181 182
Reusability of Fe(III)-TAML/LDH composite. The detailed information for reusability test of Fe(III)-TAML/LDH composite is provided in Text S7.
183
Computational methods. Gaussian 09W program46 with DFT method (B3LYP)47
184
was used to investigate the structures of Fe(III)-TAML and TBBPA, the infrared
185
spectrum of Fe(III)-TAML was also calculated. The 6-311G(d,p) basis set was
186
applied to all atoms of Fe(III)-TAML and TBBPA (i.e. C, H, O, N, Br and Fe). The
187
calculated infrared spectrum of Fe(III)-TAML was analyzed using Gview program
188
based on the optimized results. The frontier electron densities (FEDs) of the highest
189
occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital
190
(LUMO) of TBBPA were determined based on Gaussian output files, and the values
191
of 2FED2HOMO and (FED2HOMO + FED2LUMO) were calculated to predict the possible
192
reaction sites for the redox reactions.
193 194
Toxicity assay. The detailed information on toxicity assay is provided in Text S8.
Results and discussion
195
Characterization of Synthesized LDH. Based on the XRD and TEM
196
measurements (Fig. 2 and Fig. S1), the synthesized nanosheets of LDH show good
197
crystallinity, the crystallite is partially overlapped and shaped in a hexagonal form
198
with lateral size of 60-120 nm (Fig. S1), which is consistent with the results reported 9
ACS Paragon Plus Environment
Environmental Science & Technology
199
by Xu et al.48
200
Adsorption of Fe(III)-TAML on LDH. Due to the electrostatic interaction
201
between negatively charged Fe(III)-TAML and positively charged LDH layers, high
202
adsorption of Fe(III)-TAML on LDH was observed with fitted parameters
203
KL=0.00011 L µmol-1 and Qmax=320.72 µmol g-1 for Langmuir isotherm (Fig. S2 and
204
Table S1). The XRD patterns show that as more Fe(III)-TAML molecules are loaded,
205
the basal spacings of LDH increase from 7.6 to 10.9 Å (Fig. 2), confirming the
206
intercalation of Fe(III)-TAML into the hydrotalcite interlayer. Taking into account the
207
thickness of hydroxide layer sheet of ~4.8 Å49, the interlayer distance can be
208
estimated as 6.1 Å when the adsorption reaches the plateau, which is smaller than the
209
length of Fe(III)-TAML molecule (7.7 Å) as indicated by the distance between the
210
two oxygen atoms in the carbonyl groups (22O and 24O, Fig. 1(b)). To demonstrate
211
the conformation change of Fe(III)-TAML molecule in clay interlayer, the orientation
212
of Fe(III)-TAML was simulated at different interlayer expansion of 3.5, 5.4, and 6.1 Å,
213
respectively (Fig. 3). It is clearly shown that at higher intercalation loading, the
214
Fe(III)-TAML molecule tends to arrange into tilted configuration to accommodate
215
more Fe(III)-TAML molecules (Fig. 3), similar results were also observed for
216
biomolecules adsorbed on clay minerals36,50,51.
217
The strong interaction between Fe(III)-TAML and LDH was further supported
218
from FTIR spectroscopic analysis (Fig. 4). To help interpret the IR peak shift,
219
theoretical spectrum of Fe(III)-TAML was calculated and compared with our
220
experimental results (Table S2). In accordance with the calculated IR bands, peak 10
ACS Paragon Plus Environment
Page 10 of 39
Page 11 of 39
Environmental Science & Technology
221
assignments for Fe(III)-TAML are listed in Table S2. The synchronous vibration of
222
both C-N in-plane stretching (14C-11N and 13C-16N) is assigned to the peak at 1393
223
cm-1. The peak at 1455 cm-1 corresponds to the synchronous bending of hydrogen at
224
20C, 21C, 37C and 41C, while the 1475 cm-1 peak for the same bending at 46C and
225
50C. In addition, the peaks at 1575 and 1626 cm-1 are attributed to the C=O
226
(14C=22O and 13C=45O) in-plane stretching and HOH bending of water molecule
227
axially coordinated to Fe atom, respectively. Compared to the IR spectrum of free
228
Fe(III)-TAML, clear peak shifts were observed as Fe(III)-TAML is adsorbed on LDH.
229
The 1562 cm-1 peak results from the red shift of 1575 cm-1 peak, which is attributed to
230
the electrostatic interaction between the negatively charged oxygen atoms of carboxyl
231
groups and the positively charged LDH layers. This electrostatic effect indirectly
232
influences the C-N in-plane stretching, leading to the red shift from 1393 cm-1 in
233
Fe(III)-TAML to 1386 cm-1 in Fe(III)-TAML/LDH. The 1459 and 1477 cm-1 peaks,
234
corresponding to C-H bending, appear in Fe(III)-TAML/LDH with less obvious shifts.
235
The 1626 cm-1 peak in Fe(III)-TAML is not observable due to the overlap with the
236
HOH bending band for water molecule associated with LDH. Compared to the 1630
237
cm-1 peak in LDH, a blue shift occurs for the HOH bending of water molecule
238
adsorbed on LDH (1641 cm-1), which could be accounted for the interactions of H2O
239
molecule in LDH interlayer with O and N atoms of acylamino groups in
240
Fe(III)-TAML via hydrogen bonding. The occurrence of new peaks and the IR shifts
241
demonstrate the interaction between Fe(III)-TAML anions and positively charged
242
LDH layers, e.g. electrostatic and hydrogen bonding interaction. 11
ACS Paragon Plus Environment
Environmental Science & Technology
243
Degradation of TBBPA by Fe(III)-TAML and Fe(III)-TAML/LDH at different
244
pHs. The degradation of TBBPA catalyzed by Fe(III)-TAML and Fe(III)-TAML/LDH
245
shows strong pH dependent behavior (Fig. 5). As listed in Table S3, the fitted
246
pseudo-first-order rate constant is significantly enhanced as pH increases to 9 or 10
247
compared to that at pH 8. The highest degradation rate was observed at pH 10 with
248
complete degradation of TBBPA in less than 20 s, similar results were also observed
249
in the literature22,26,29. The high reactivity of Fe(III)-TAML under alkaline condition
250
can be explained by the increasing proportion of deprotonated species of
251
Fe(III)-TAML at higher pH. During the catalytic process, Fe(III)-TAML is initially
252
oxidized by H2O2 to form Fe(IV)-TAML52 or Fe(V)-TAML53, which can be then
253
quickly reduced by organic pollutants. However, different species of Fe(III)-TAML
254
shows distinct reactivity to form oxidized Fe(IV)-TAML or Fe(V)-TAML. As listed in
255
Scheme S1, the deprotonated species [Fe(III)-TAML(OH)]2- with higher electron
256
density compared to [Fe(III)-TAML(OH2)]- would act as a better electron donor to be
257
oxidized by H2O2 at a much faster reaction rate (k2 ≫ k1)28. With the similar
258
explanation, [Fe(III)-TAML(OH)]2- is more prone to react with H2O2 than the
259
deprotonated HO2- (k4 < k2), in addition, this reaction might also be affected by
260
electrostatic repulsion28. Therefore, the ionization for both the axial water molecule in
261
the coordination sphere of Fe(III)-TAML and H2O2 would play vital roles for the
262
generation of oxidized Fe(III)-TAML and subsequent degradation of pollutants. As
263
reported in prior studies, the dissociation constants (pKa) for water associated with
264
Fe(III)-TAML and H2O2 are ~10 and >11, respectively28,32, which indicates that the 12
ACS Paragon Plus Environment
Page 12 of 39
Page 13 of 39
Environmental Science & Technology
265
highest reactivity for Fe(III)-TAML would be in a narrow range between pH 10 and
266
11.
267
Compared to Fe(III)-TAML, the pseudo-first-order rate constants for TBBPA
268
degradation increase in the presence of Fe(III)-TAML/LDH, especially at pH 8 (Table
269
S3). To further investigate the underlying mechanism, the adsorption of TBBPA on
270
Fe(III)-TAML/LDH was studied. As illustrated in Figure S4, the TBBPA adsorption
271
on Fe(III)-TAML/LDH can be fitted with Langmuir isotherm model (Table S4), with
272
the maximum adsorption of 76.38, 107.76 and 157.88 µmol g-1 for pH at 8, 9 and 10,
273
respectively. The pH-dependent adsorption is attributed to the extent of ionization
274
forms for TBBPA at different pHs. As simulated by Visual MINTEQ, with the two
275
pKa values of 7.4 and 8.5, the proportion of TBBPA2- increases from 20.96% to 97.35%
276
as increase of pH from 8 to 10, whereas the percentage of TBBPA- decreases from
277
60.49% to 2.65% (Table S5). It indicates that at higher pH, TBBPA exists as a more
278
negatively charged ion, which would be more facile to be adsorbed by LDH through
279
an ion exchange process. The strong adsorption of TBBPA on LDH could explain the
280
enhancement of the degradation rate catalyzed by Fe(III)-TAML/LDH. Furthermore,
281
as shown in Figure S5, the adsorption of TBBPA on LDH is fast, after 4 s, the ratios
282
of adsorption amount (Qt) to the equilibrium adsorption amount (Qe) could reach 22%,
283
29% and 32% for pH 8, 9 and 10, respectively. According to Fick’s law, the diffusate
284
concentration gradient is the driving force for mass transfer process54. Due to the high
285
reaction rate between TBBPA and active Fe(III)-TAML, the concentration of TBBPA
286
in LDH interlayer decreases so fast that a great concentration difference is generated 13
ACS Paragon Plus Environment
Environmental Science & Technology
287
between LDH and homogeneous solution, therefore, a high diffusion rate would be
288
expected during the reaction period. To further quantitatively investigate the diffusion
289
process, the total amount of TBBPA remained (including both TBBPA in solution and
290
adsorbed on LDH surface) and the amount of TBBPA in solution were separately
291
measured by addition of methanol to extract the adsorbed TBBPA and by simple
292
centrifugation, respectively. The amount of TBBPA adsorbed on LDH could be
293
obtained by calculating the difference between these two measurements. As shown in
294
Figure S6, the amount of TBBPA adsorbed on LDH maintains in the range of 33-55
295
µmol g-1 in the first 30 s reaction period at pH 8, which is even higher than that (32
296
µmol g-1) after 120 s for adsorption kinetics of TBBPA on Fe(III)-TAML/LDH (inset
297
of Figure S6). The fast TBBPA adsorption on LDH provides direct evidence that the
298
diffusion of TBBPA onto LDH may not limit the reaction between TBBPA and
299
Fe(III)-TAML. In addition, Fe(III)-TAML/LDH with different Fe(III)-TAML loadings
300
of 4.42, 6.96, 8.36, 11.92, and 22.38 µmol g-1 were prepared to degrade the same
301
amount of TBBPA (18.4 µM). As illustrated in Figure S7, the pseudo-first-order
302
reaction rate constants are linearly proportional to the amount of Fe(III)-TAML
303
loaded on LDH, indicating that the amount of Fe(III)-TAML rather than mass transfer
304
dominated the surface reaction. So, it can be concluded that the adsorption of TBBPA
305
on LDH would not be a rate limiting step to slow down the degradation reaction,
306
instead it would enhance the dissipation of TBBPA by increasing the contact
307
possibility between TBBPA and TAML activator.
308
The most significant enhancement for TBBPA degradation efficiency by 14
ACS Paragon Plus Environment
Page 14 of 39
Page 15 of 39
Environmental Science & Technology
309
Fe(III)-TAML/LDH occurred at pH 8 (Table S3), which is not consistent with the
310
adsorption results (Fig. S4). At pH 10, the predominant species of TBBPA is TBBPA2-,
311
and the increase of degradation rate for Fe(III)-TAML/LDH is 1.58 compared to free
312
Fe(III)-TAML (Table S3), so the contribution of adsorption for the increase of
313
reactivity is < 2. Whereas, when pH increases from 8 to 9, the degradation rate
314
increases 15 times for free Fe(III)-TAML (Table S3). Therefore, the pH effect on the
315
reactivity is more pronounced. It has been reported that the basicity in clay interlayer
316
may increase resulting from the bicarbonate hydrolysis and the presence of Mg-O
317
Lewis base sites on MgAl LDH materials55. In the pH range between 8 and 10,
318
bicarbonate is the predominant species56. The presence of bicarbonate can be
319
evidenced by the strong band at 1370 cm-1 (Fig. 4), which is assigned to the
320
antisymmetric v3 vibration of bicarbonate anion57. The exposure to air during LDH
321
preparation inevitably introduces carbon dioxide into the system. On the other hand,
322
the Mg-O basic sites on MgAl LDH could also enhance the basicity55. Therefore, it
323
can be concluded that the pH in clay interlayer is higher than that in homogeneous
324
solution58. Compared to pH 9 and 10, the contributions from bicarbonate and Mg-O
325
sites are more pronounced at pH 8, since the change of OH- level would be more
326
significant at neutral or near neutral pH, which might account for the higher increase
327
for interlayer pH and higher reactivity for Fe(III)-TAML/LDH system at pH 8. It was
328
summarized by DeNardo et al.59 that the rate constants for both Fe-TAML activation
329
by H2O2 and substrate oxidation by reactive Fe-TAML are positively relevant to
330
solution pH, indicating that both rate constants increase with the increase of pH levels. 15
ACS Paragon Plus Environment
Environmental Science & Technology
Page 16 of 39
331
Therefore, it is likely that the main utility of Fe(III)-TAML/LDH is to increase the
332
activity at lower solution pH by raising the interlayer pH.
333
To investigate the degradation efficiency of TBBPA by Fe(III)-TAML/LDH, both
334
mineralization and debromination during the degradation process were studied. A
335
significant increase of debromination rate was observed for Fe(III)-TAML/LDH
336
system compared to free Fe(III)-TAML (Fig. S8 and Table S6), which can also be
337
explained by the strong interaction between TBBPA and LDH. As reported by Gupta
338
et al.22, Fe(III)-TAML could achieve nearly complete mineralization and
339
dechlorination for chlorophenols. However, only limited debromination was observed
340
in our experiment. As shown in Figure S8, at 120 s, the debromination ratios are
341
15.66, 23.07 and 30.92% for Fe(III)-TAML at pH 8, 9, and 10, respectively, while the
342
respective
343
Fe(III)-TAML/LDH. When the reaction is extended to 1800 s, similar debromination
344
ratios are also achieved for Fe(III)-TAML (21.1%) and Fe(III)-TAML/LDH (22.1%)
345
at pH 8 (Fig. S8(a)). The low oxidative debromination for TBBPA might be attributed
346
to the generation of the brominated organic carboxylic acids (Fig. 6), these
347
compounds are reported to be resistant to various oxidants (e.g. hydroxyl radicals and
348
ozone)21. The strong pH dependence for debromination rate and efficiency also results
349
from varied concentration of the most reactive species [Fe(III)-TAML(OH)]2- under
350
different pH values.
debromination
values
are
21.85,
23.41,
and
30.19%
for
351
Due to the low aqueous solubility of TBBPA under acidic condition60, after addition
352
of perchloric acid to terminate the reaction, TBBPA would precipitate, which might 16
ACS Paragon Plus Environment
Page 17 of 39
Environmental Science & Technology
353
interfere with TOC measurement. Therefore, the mineralization efficiency during the
354
reaction was determined by calculating the difference between the initial and final
355
TOC (0 and 120 s) in reaction solution. As illustrated in Figure S9(a), TOC removals
356
after 120 s reaction are 29.22, 38.81 and 43.90% for Fe(III)-TAML under pH 8, 9,
357
and
358
Fe(III)-TAML/LDH. Similar results were also observed when the reaction was fully
359
completed (1800 s) (Fig. S9(b)). The results from debromination and mineralization
360
experiments show that in Fe(III)-TAML/LDH system, even there is no significant
361
increase for total debromination, higher TOC removal was observed compared to
362
Fe(III)-TAML. The discrepancy might be explained that during the experiment, the
363
reaction intermediates can be adsorbed on LDH, which could promote the reaction to
364
proceed, resulting in higher mineralization efficiency. On the other hand, these
365
brominated byproducts may accumulate on LDH external surfaces to retard further
366
debromination reaction. The adsorption of final reaction products on LDH could also
367
partially account for TOC removal. Therefore, the adsorption of TBBPA and the
368
reaction intermediates on LDH might affect the extent of degradation for TBBPA by
369
Fe(III)-TAML/LDH/H2O2 system, further study is needed to reveal the detailed
370
degradation mechanism by measuring the intermediates during the reaction and under
371
different reaction conditions.
10,
respectively,
and
increase
to
46.05,
54.33
and
62.30%
for
372
TBBPA degradation pathway. Based on the reaction products identified in our
373
experiment (Figs. S10 and S11) and previous studies17,21,61, the degradation pathway
374
of TBBPA is shown in Figure 6. The degradation is initiated by an electron 17
ACS Paragon Plus Environment
Environmental Science & Technology
Page 18 of 39
375
deprivation reaction occurring on the hydroxyl group of TBBPA, generating the
376
phenoxy radical R1, which is then transformed to a more stable radical R2 by
377
resonance and leads to the relocation of the unpaired electron to the aromatic ring.
378
Subsequently, the carbocation intermediate radical R3 and a new radical R4 are
379
released by β-scission (cleavage between the isopropyl group and the benzene ring) of
380
R2. Similar results were also observed by previous researchers for TBBPA
381
degradation by manganese dioxide17, UV/Fenton reaction62 and ozonation21, they
382
proposed that the primary reaction includes the bond cleavage at the middle carbon
383
atom of TBBPA21. Substitution and deprotonation can transform the carbocation R3 to
384
4-(2-hydroxyisopropyl)-2,6-dibromophenol
385
4-isopropylene-2,6-dibromophenol (R5), respectively. Meanwhile, P1 can be
386
dehydrated
387
1-(3,5-dibromo-4-hydroxyphenyl)ethanone (P2, MW = 294). Although R5 was not
388
detected in our study, probably due to the rapid transformation upon contact with
389
Fe(III)-TAML, it has been proposed as the intermediate during the oxidation of
390
TBBPA by Lin et al.17 Our results further demonstrate that the reaction intermediate
391
P2 completely disappears after 600 s (Figs. S10 and S11), with the final products of
392
(Z)-4-acetyl-2-bromopent-2-enedioic acid (P3, MW = 251), CO2 or CO, and bromide
393
ion. Since product P3 has not been reported in the literature, the structure of P3 can be
394
only deduced from the information of mass fragments (Figs. S10 and S11). Similarly,
395
R4 may also be transformed to 2-bromosuccinic acid (P4, MW = 197), CO2 or CO,
396
and bromide ion by the oxidation of Fe(III)-TAML. In summary, the oxidative
to
form
R5,
which
(P1,
is
MW
further
18
ACS Paragon Plus Environment
=
310)
oxidized
and
to
Page 19 of 39
Environmental Science & Technology
397
degradation of TBBPA is assumed to proceed via the cleavage of the middle carbon
398
atom of TBBPA, generating a carbocation intermediate R3 and a radical R4, which
399
are finally transformed to P3 and P4. The proposed degradation pathway is further
400
supported from the theoretical calculations. The calculated FED result for each atom
401
of TBBPA is listed in Figure S12(a), and the three-dimensional iso-surfaces of HOMO
402
(Fig. S12(b)) and LUMO (Fig. S12(c)) frontier orbitals are also exhibited for
403
visualization of the electron densities. As shown in Figure S12(a), 20O and 21O of
404
TBBPA have the highest 2FED2HOMO values (0.1558, 0.1558), suggesting that 20O or
405
21O should be the most possible reaction site, where one electron is extracted by
406
Fe-TAML, forming the TBBPA radical. It is consistent with the initial step in the
407
degradation pathway (Fig. 6). In addition to 20O and 21O, high 2FED2HOMO values
408
for 3C (0.1377) and 10C (0.1376) are also observed among all the C atoms of TBBPA,
409
indicating that these two sites are susceptible for attack by the reactive species of
410
Fe-TAML, which is confirmed by the reaction products, e.g., P1 and P2.
411
Reusability of Fe(III)-TAML/LDH. One of the advantages for immobilization of
412
bioactive materials is to increase their reutilization, hence in current study, we also
413
tested the reusability of the novel Fe(III)-TAML/LDH composite for degradation of
414
TBBPA. As illustrated in Figure S13(a), the activity of Fe(III)-TAML/LDH could be
415
retained for at least three cycles. The removal percentages of TBBPA are 100.0, 100.0,
416
44.6 and 0.4% for the first, second, third and fourth cycle, respectively, after 300 s
417
reaction for each cycle. The decrease of the reactivity for Fe(III)-TAML/LDH could
418
be attributed to the suicidal inactivation of Fe(III)-TAML58 and the release of 19
ACS Paragon Plus Environment
Environmental Science & Technology
419
Fe(III)-TAML from LDH. Based on the results of stability tests for both
420
Fe(III)-TAML and Fe(III)-TAML/LDH (Fig. S14), the high activity of free
421
Fe(III)-TAML could be retained for four cycles, then decreases significantly, whereas,
422
the Fe(III)-TAML/LDH composite shows high reactivity until six cycles. The
423
decrease of activity can be explained by the suicidal inactivation of Fe(III)-TAML as
424
the substrate was quickly consumed during the reaction59,63. However, the presence of
425
LDH could increase the concentration of TBBPA in clay interlayer, which may
426
prolong the lifetime of Fe(III)-TAML63. More importantly, we found that
427
Fe(III)-TAML was released from LDH during the reaction by detecting the Fe in
428
solution, and the sum of Fe after each reaction cycle was 49.16 µg L-1 (0.88 µM) (Fig.
429
S13(b)), which is close to the total molar concentration of Fe(III)-TAML (1.0 µM)
430
used in the system. The release of Fe(III)-TAML could be due to the strong
431
competition for the adsorption sites in LDH interlayer with bromide ion (Br-), reaction
432
intermediates, e.g., carboxylic acids (P3 and P4), hydroxyl ion (OH-) and
433
bicarbonate/carbonate ion (HCO3-/CO32-). For example, at pH 10, the concentration of
434
OH- is 100 µM, which is 2 order of magnitude higher than Fe(III)-TAML.
435
Acute Toxicity. To demonstrate the “green catalyst” nature of Fe(III)-TAML, the
436
toxicity test was also conducted to assess the potential risk of reaction intermediates
437
produced from TBBPA degradation. Photobacterium phosphoreum (P. phosphoreum)
438
was used for the acute toxicity assay. As indicated in Figure S15, before the reaction,
439
the initial inhibitions of TBBPA mixed with Fe(III)-TAML or Fe(III)-TAML/LDH are
440
over 30%, after 600 s reaction period, the inhibition values decrease to less than 0.9%. 20
ACS Paragon Plus Environment
Page 20 of 39
Page 21 of 39
Environmental Science & Technology
441
Significant decrease in the toxicity towards P. phosphoreum could be attributed to the
442
complete decomposition of TBBPA and the lower toxicity of the reaction products. In
443
addition, based on our observation, the toxicities of Fe(III)-TAML and
444
Fe(III)-TAML/LDH themselves towards P. phosphoreum are negligible (Fig. S15),
445
which is consistent with the results in the literatures.22,64 Gupta et al. reported that the
446
LONEC values (the highest observed concentrations of activator that show no
447
bacterial death) for Fe(III)-TAML activators were greater than 30 mg L-1 (~50 µM)
448
based on the luminescent bacteria test22. Truong et al.64 used the embryonic zebrafish
449
to assess the developmental toxicity of Fe(III)-TAML, observing no significant
450
mortalities and morphological malformations for embryonic zebrafish even exposed
451
to 250 µM Fe(III)-TAML.
452
Environmental Implications
453
In this study, a new Fe(III)-TAML/LDH composite was synthesized by
454
intercalating Fe(III)-TAML into the LDH interlayer via anion exchange. Both FTIR
455
and XRD spectra indicate that Fe(III)-TAML can form strong interaction with LDH,
456
and with the increase of adsorption, Fe(III)-TAML molecules tend to exist as the tilted
457
configuration. Compared to the degradation of TBBPA by Fe(III)-TAML,
458
Fe(III)-TAML/LDH showed significantly improved degradation efficiency, including
459
faster elimination of TBBPA and higher TOC removal. It could be attributed to the
460
strong adsorption of TBBPA onto LDH, which would increase the contact possibility
461
between TBBPA and Fe(III)-TAML.
462
Fe(III)-TAML has been considered as a green catalyst for rapid degradation of
463
many persistent organic contaminants with high inherent reactivity and low
464
environmental risks27,30,33. Our results indicate that the combination of Fe(III)-TAML 21
ACS Paragon Plus Environment
Environmental Science & Technology
465
with LDH provides a new attempt to not only promote the degradation rate, but also
466
improve its reusability, which would extend the application of Fe(III)-TAML in a
467
wider range of pH for advanced treatment of recalcitrant pollutants with lower cost.
468 469
470
Acknowledgements
471
This work was financially supported by the National Key Basic Research Program of
472
China (2014CB441102), National Science Foundation of China (grants 21222704,
473
21237002 and 21477051) and the Collaborative Innovation Center for Regional
474
Environmental Quality. We thank the Analytical Center and High Performance
475
Computing Center of Nanjing University for the characterization of samples and
476
computational study.
477 478
Supporting Information
479
This material is available free of charge via the Internet at http://pubs.acs.org.
480
Texts S1-S8, Tables S1-S6, Scheme S1, and Figures S1-S15
481 482
References:
483
(1) Covaci, A.; Voorspoels, S.; Abdallah, M. A.-E.; Geens, T.; Harrad, S.; Law, R. J.
484
Analytical and environmental aspects of the flame retardant tetrabromobisphenol A
485
and its derivatives. J. Chromatogr. A 2009, 1216 (3), 346-363.
486
(2) Alaee, M.; Arias, P.; Sjödin, A.; Bergman, Å. An overview of commercially used
487
brominated flame retardants, their applications, their use patterns in different 22
ACS Paragon Plus Environment
Page 22 of 39
Page 23 of 39
Environmental Science & Technology
488
countries/regions and possible modes of release. Environ. Int. 2003, 29 (6), 683-689.
489
(3) Birnbaum, L. S.; Staskal, D. F. Brominated flame retardants: Cause for concern?
490
Environ. Health Persp. 2004, 112 (1), 9-17.
491
(4) Guo, Y.; Zhou, J.; Lou, X.; Liu, R.; Xiao, D.; Fang, C.; Wang, Z.; Liu, J. Enhanced
492
degradation of tetrabromobisphenol A in water by a UV/base/persulfate system:
493
Kinetics and intermediates. Chem. Eng. J. 2014, 254, 538-544.
494
(5) Guo, Y.; Lou, X.; Xiao, D.; Xu, L.; Wang, Z.; Liu, J. Sequential reduction–
495
oxidation for photocatalytic degradation of tetrabromobisphenol A: Kinetics and
496
intermediates. J. Hazard. Mater. 2012, 241-242, 301-306.
497
(6) Zweidinger, R. A.; Cooper, S. D.; Erickson, M. D.; Michael, L. C.; Pellizzari, E. D.
498
Sampling and analysis for semivolatile brominated organics in ambient air. In
499
Monitoring Toxic Substances; Schuetzle, D., Eds.; American Chemical Society:
500
Washington, DC 1979; pp 217-231.
501
(7)
502
tetrabromobisphenol A and its metabolite found in river and marine sediments in
503
Japan. Chemosphere 1983, 12 (11-12), 1533-1539.
504
(8) Watanabe, I.; Kashimoto, T.; Tatsukawa, R. Identification of the flame retardant
505
tetrabromobisphenol A in the river sediment and the mussel collected in Osaka. Bull.
506
Environ. Contam. Toxicol. 1983, 31 (1), 48-52.
507
(9) Yang, S.; Wang, S.; Liu, H.; Yan, Z. Tetrabromobisphenol A: tissue distribution in
508
fish, and seasonal variation in water and sediment of Lake Chaohu, China. Environ.
509
Sci. Pollut. Res. 2012, 19 (9), 4090-4096.
Watanabe,
I.;
Kashimoto,
T.;
Tatsukawa,
R.
23
ACS Paragon Plus Environment
The
flame
retardant
Environmental Science & Technology
510
(10) Cariou, R.; Antignac, J.-P.; Marchand, P.; Berrebi, A.; Zalko, D.; Andre, F.; Le
511
Bizec, B. New multiresidue analytical method dedicated to trace level measurement
512
of brominated flame retardants in human biological matrices. J. Chromatogr. A 2005,
513
1100 (2), 144-152.
514
(11) Kuiper, R. V.; van den Brandhof, E. J.; Leonards, P. E. G.; van der Ven, L. T. M.;
515
Wester, P. W.; Vos, J. G. Toxicity of tetrabromobisphenol A (TBBPA) in zebrafish
516
(Danio rerio) in a partial life-cycle test. Arch. Toxicol. 2007, 81 (1), 1-9.
517
(12) Hu, J.; Liang, Y.; Chen, M.; Wang, X. Assessing the toxicity of TBBPA and
518
HBCD by zebrafish embryo toxicity assay and biomarker analysis. Environ. Toxicol.
519
2009, 24 (4), 334-342.
520
(13) Meerts, I. A. T. M.; van Zanden, J. J.; Luijks, E. A. C.; van Leeuwen-Bol, I.;
521
Marsh, G.; Jakobsson, E.; Bergman, Å.; Brouwer, A. Potent competitive interactions
522
of some brominated flame retardants and related compounds with human transthyretin
523
in vitro. Toxicol. Sci. 2000, 56 (1), 95-104.
524
(14) Decherf, S.; Seugnet, I.; Fini, J.-B.; Clerget-Froidevaux, M.-S.; Demeneix, B. A.
525
Disruption of thyroid hormone-dependent hypothalamic set-points by environmental
526
contaminants. Mol. Cell. Endocr. 2010, 323 (2), 172-182.
527
(15) Reistad, T.; Mariussen, E.; Ring, A.; Fonnum, F. In vitro toxicity of
528
tetrabromobisphenol A on cerebellar granule cells: Cell death, free radical formation,
529
calcium influx and extracellular glutamate. Toxicol. Sci. 2007, 96 (2), 268-278.
530
(16) Sjödin, A.; Carlsson, H. A.; Thuresson, K.; Sjölin, S.; Bergman, Å.; Östman, C.
531
Flame retardants in indoor air at an electronics recycling plant and at other work 24
ACS Paragon Plus Environment
Page 24 of 39
Page 25 of 39
Environmental Science & Technology
532
environments. Environ. Sci. Technol. 2001, 35 (3), 448-454.
533
(17) Lin, K.; Liu, W.; Gan, J. Reaction of tetrabromobisphenol A (TBBPA) with
534
manganese dioxide: Kinetics, products, and pathways. Environ. Sci. Technol. 2009, 43
535
(12), 4480-4486.
536
(18) Eriksson, J.; Rahm, S.; Green, N.; Bergman, Å.; Jakobsson, E. Photochemical
537
transformations of tetrabromobisphenol A and related phenols in water. Chemosphere
538
2004, 54 (1), 117-126.
539
(19) Han, S. K.; Bilski, P.; Karriker, B.; Sik, R. H.; Chignell, C. F. Oxidation of flame
540
retardant tetrabromobisphenol A by singlet oxygen. Environ. Sci. Technol. 2008, 42
541
(1), 166-172.
542
(20) Ding, Y.; Zhu, L.; Wang, N.; Tang, H. Sulfate radicals induced degradation of
543
tetrabromobisphenol A with nanoscaled magnetic CuFe2O4 as a heterogeneous
544
catalyst of peroxymonosulfate. Appl. Catal. B: Environ. 2013, 129, 153-162.
545
(21) Qu, R.; Feng, M.; Wang, X.; Huang, Q.; Lu, J.; Wang, L.; Wang, Z. Rapid
546
removal of tetrabromobisphenol A by ozonation in water: Oxidation products,
547
reaction pathways and toxicity assessment. PLoS ONE 2015, 10 (10), e0139580; DOI
548
10.1371/journal.pone.0139580.
549
(22) Gupta, S. S.; Stadler, M.; Noser, C. A.; Ghosh, A.; Steinhoff, B.; Lenoir, D.;
550
Horwitz, C. P.; Schramm, K.-W.; Collins, T. J. Rapid total destruction of
551
chlorophenols by activated hydrogen peroxide. Science 2002, 296 (5566), 326-328.
552
(23) Mierzwicki, K.; Berski, S.; Latajka, Z. AIM and ELF analysis of the H-, Me-,
553
and F-substituted FeIII–TAML complexes. Chem. Phys. Lett. 2011, 507 (1-3), 29-36. 25
ACS Paragon Plus Environment
Environmental Science & Technology
554
(24) Chanda, A.; Khetan, S. K.; Banerjee, D.; Ghosh, A.; Collins, T. J. Total
555
degradation of fenitrothion and other organophosphorus pesticides by catalytic
556
oxidation employing Fe-TAML peroxide activators. J. Am. Chem. Soc. 2006, 128 (37),
557
12058-12059.
558
(25) Mondal, S.; Hangun-Balkir, Y.; Alexandrova, L.; Link, D.; Howard, B.; Zandhuis,
559
P.; Cugini, A.; Horwitz, C. P.; Collins, T. J. Oxidation of sulfur components in diesel
560
fuel using Fe-TAML® catalysts and hydrogen peroxide. Catal. Today 2006, 116 (4),
561
554-561.
562
(26) Shappell, N. W.; Vrabel, M. A.; Madsen, P. J.; Harrington, G.; Billey, L. O.; Hakk,
563
H.; Larsen, G. L.; Beach, E. S.; Horwitz, C. P.; Ro, K.; Hunt, P. G.; Collins, T. J.
564
Destruction of estrogens using Fe-TAML/peroxide catalysis. Environ. Sci. Technol.
565
2008, 42 (4), 1296-1300.
566
(27) Chahbane, N.; Popescu, D.-L.; Mitchell, D. A.; Chanda, A.; Lenoir, D.; Ryabov,
567
A. D.; Schramm, K.-W.; Collins, T. J. FeIII-TAML-catalyzed green oxidative
568
degradation of the azo dye Orange II by H2O2 and organic peroxides: Products,
569
toxicity, kinetics, and mechanisms. Green Chem. 2007, 9, 49-57.
570
(28) Ghosh, A.; Mitchell, D. A.; Chanda, A.; Ryabov, A. D.; Popescu, D. L.; Upham,
571
E. C.; Collins, G. J.; Collins, T. J. Catalase−peroxidase activity of iron(III)-TAML
572
activators of hydrogen peroxide. J. Am. Chem. Soc. 2008, 130 (45), 15116-15126.
573
(29) Shen, L. Q.; Beach, E. S.; Xiang, Y.; Tshudy, D. J.; Khanina, N.; Horwitz, C. P.;
574
Bier, M. E.; Collins, T. J. Rapid, biomimetic degradation in water of the persistent
575
drug sertraline by TAML catalysts and hydrogen peroxide. Environ. Sci. Technol. 26
ACS Paragon Plus Environment
Page 26 of 39
Page 27 of 39
Environmental Science & Technology
576
2011, 45 (18), 7882-7887.
577
(30) Kundu, S.; Chanda, A.; Khetan, S. K.; Ryabov, A. D.; Collins, T. J. TAML
578
activator/peroxide-catalyzed facile oxidative degradation of the persistent explosives
579
trinitrotoluene and trinitrobenzene in micellar solutions. Environ. Sci. Technol. 2013,
580
47 (10), 5319-5326.
581
(31) Polshin, V.; Popescu, D.-L.; Fischer, A.; Chanda, A.; Horner, D. C.; Beach, E. S.;
582
Henry, J.; Qian, Y.-L.; Horwitz, C. P.; Lente, G.; Fabian, I.; Münck, E.; Bominaar, E.
583
L.; Ryabov, A. D.; Collins, T. J. Attaining control by design over the hydrolytic
584
stability of Fe-TAML oxidation catalysts. J. Am. Chem. Soc. 2008, 130 (13),
585
4497-4506.
586
(32) Popescu, D.-L.; Chanda, A.; Stadler, M. J.; Mondal, S.; Tehranchi, J.; Ryabov, A.
587
D.; Collins, T. J. Mechanistically inspired design of FeIII-TAML peroxide-activating
588
catalysts. J. Am. Chem. Soc. 2008, 130 (37), 12260-12261.
589
(33) Tang, L. L.; DeNardo, M. A.; Gayathri, C.; Gil, R. R.; Kanda, R.; Collins, T. J.
590
TAML/H2O2 oxidative degradation of metaldehyde: Pursuing better water treatment
591
for the most persistent pollutants. Environ. Sci. Technol. 2016, 50 (10), 5261-5268.
592
(34) Zhang, Y.; Xu, C.; Li, B. Self-assembly of hemin on carbon nanotube as highly
593
active peroxidase mimetic and its application for biosensing. RSC Adv. 2013, 3 (17),
594
6044-6050.
595
(35) Guo, Y.; Deng, L.; Li, J.; Guo, S.; Wang, E.; Dong, S. Hemin-graphene hybrid
596
nanosheets with intrinsic peroxidase-like activity for label-free colorimetric detection
597
of single-nucleotide polymorphism. ACS Nano 2011, 5 (2), 1282-1290. 27
ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 39
598
(36) Xiong, J.; Hang, C.; Gao, J.; Guo, Y.; Gu, C. A novel biomimetic catalyst
599
templated by montmorillonite clay for degradation of 2,4,6-trichlorophenol. Chem.
600
Eng. J. 2014, 254 (0), 276-282.
601
(37) Demeter, E. L.; Hilburg, S. L.; Washburn, N. R.; Collins, T. J.; Kitchin, J. R.
602
Electrocatalytic oxygen evolution with an immobilized TAML activator. J. Am. Chem.
603
Soc. 2014, 136 (15), 5603-5606.
604
(38) Pesic, L.; Salipurovic, S.; Markovic, V.; Vucelic, D.; Kagunya, W.; Jones, W.
605
Thermal characteristics of a synthetic hydrotalcite-like material. J. Mater. Chem. 1992,
606
2 (10), 1069-1073.
607
(39) Khan, A. I.; O'Hare, D. Intercalation chemistry of layered double hydroxides:
608
Recent developments and applications. J. Mater. Chem. 2002, 12 (11), 3191-3198.
609
(40) Wypych, F.; Bubniak, G. A.; Halma, M.; Nakagaki, S. Exfoliation and
610
immobilization of anionic iron porphyrin in layered double hydroxides. J. Colloid
611
Interf. Sci. 2003, 264, 203-207.
612
(41) Nakagaki, S.; Halma, M.; Bail, A.; Arízaga, G. G. C.; Wypych, F. First insight
613
into catalytic activity of anionic iron porphyrins immobilized on exfoliated layered
614
double hydroxides. J. Colloid Interf. Sci. 2005, 281, 417-423.
615
(42) Wypych, F.; Bail, A.; Halma, M.; Nakagaki, S. Immobilization of iron(III)
616
porphyrins
617
(3-aminopropyl)triethoxysilane. J. Catal. 2005, 234, 431-437.
618
(43) Ren, L.; He, J.; Zhang, S.; Evans, D. G.; Duan, X. Immobilization of penicillin G
619
acylase in layered double hydroxides pillared by glutamate ions. J. Mol. Catal. B:
on
exfoliated
MgAl
layered
double
hydroxide,
28
ACS Paragon Plus Environment
grafted
with
Page 29 of 39
Environmental Science & Technology
620
Enzym. 2002, 18 (1-3), 3-11.
621
(44) Kwak, S.-Y.; Jeong, Y.-J.; Park, J.-S.; Choy, J.-H. Bio-LDH nanohybrid for gene
622
therapy. Solid State Ionics 2002, 151 (1-4), 229-234.
623
(45) Ghosh, A.; Ryabov, A. D.; Mayer, S. M.; Horner, D. C.; Prasuhn, D. E.; Sen
624
Gupta, S.; Vuocolo, L.; Culver, C.; Hendrich, M. P.; Rickard, C. E. F.; Norman, R. E.;
625
Horwitz, C. P.; Collins, T. J. Understanding the mechanism of H+-induced
626
demetalation as a design strategy for robust iron(III) peroxide-activating catalysts. J.
627
Am. Chem. Soc. 2003, 125 (41), 12378-12379.
628
(46) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;
629
Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.;
630
Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.;
631
Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.;
632
Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.;
633
Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers,
634
E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.;
635
Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J.
636
M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.;
637
Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.;
638
Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.;
639
Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J.
640
B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision A. 02; Gaussian, Inc.:
641
Wallingford, CT, 2009. 29
ACS Paragon Plus Environment
Environmental Science & Technology
642
(47) Becke, A. D. Density - functional thermochemistry. III. The role of exact
643
exchange. J. Chem. Phys. 1993, 98 (7), 5648-5652.
644
(48) Xu, Z. P.; Stevenson, G. S.; Lu, C.-Q.; Lu, G. Q.; Bartlett, P. F.; Gray, P. P. Stable
645
suspension of layered double hydroxide nanoparticles in aqueous solution. J. Am.
646
Chem. Soc. 2006, 128, 36-37.
647
(49) Newman, S. P.; Jones, W. Synthesis, characterization and applications of layered
648
double hydroxides containing organic guests. New J. Chem. 1998, 22 (2), 105-115.
649
(50) Tagaya, H.; Ogata, A.; Kuwahara, T.; Ogata, S.; Karasu, M.; Kadokawa, J.-I.;
650
Chiba, K. Intercalation of colored organic anions into insulator host lattices of layered
651
double hydroxides. Microporous Mater. 1996, 7 (2), 151-158.
652
(51) Ukrainczyk, L.; Chibwe, M.; Pinnavaia, T. J.; Boyd, S. A. ESR study of cobalt(II)
653
tetrakis(N-methyl-4-pyridiniumyl)porphyrin and cobalt(II) tetrasulfophthalocyanine
654
intercalated in layered aluminosilicates and a layered double hydroxide. J. Phys.
655
Chem. 1994, 98 (10), 2668-2676.
656
(52) Kundu, S.; Annavajhala, M.; Kurnikov, I. V.; Ryabov, A. D.; Collins, T. J.
657
Experimental and theoretical evidence for multiple FeIV reactive intermediates in
658
TAML-activator catalysis: Rationalizing a counterintuitive reactivity order. Chem. Eur.
659
J. 2012, 18 (33), 10244-10249.
660
(53) Kundu, S.; Thompson, J. V. K.; Shen, L. Q.; Mills, M. R.; Bominaar, E. L.;
661
Ryabov, A. D.; Collins, T. J. Activation parameters as mechanistic probes in the
662
TAML iron(V)-oxo oxidations of hydrocarbons. Chem. Eur. J. 2015, 21 (4),
663
1803-1810. 30
ACS Paragon Plus Environment
Page 30 of 39
Page 31 of 39
Environmental Science & Technology
664
(54) Braida, W. J.; White, J. C.; Ferrandino, F. J.; Pignatello, J. J. Effect of Solute
665
Concentration on Sorption of Polyaromatic Hydrocarbons in Soil: Uptake Rates.
666
Environ. Sci. Technol. 2001, 35 (13) 2765-2772.
667
(55) Chisem, I. C.; Jones, W.; Martin, I.; Martin, C.; Rives, V. Probing the surface
668
acidity of lithium aluminium and magnesium aluminium layered double hydroxides. J.
669
Mater. Chem. 1998, 8 (8), 1917-1925.
670
(56) Stefánsson, A.; Bénézeth, P.; Schott, J. Carbonic acid ionization and the stability
671
of sodium bicarbonate and carbonate ion pairs to 200 °C - A potentiometric and
672
spectrophotometric study. Geochim. Cosmochim. Ac. 2013, 120, 600-611.
673
(57) Lin, Y.-J.; Li, D.-Q.; Evans, D. G.; Duan, X. Modulating effect of Mg-Al-CO3
674
layered double hydroxides on the thermal stability of PVC resin. Polym. Degrad.
675
Stabil. 2005, 88 (2), 286-293.
676
(58) Constantino, V. R. L.; Pinnavaia, T. J. Basic properties of Mg2+1-xAl3+x layered
677
double hydroxides intercalated by carbonate, hydroxide, chloride, and sulfate anions.
678
Inorg. Chem. 1995, 34 (4), 883-892.
679
(59) DeNardo, M. A.; Mills, M. R.; Ryabov, A. D.; Collins, T. J. Unifying evaluation
680
of the technical performances of iron-tetra-amido macrocyclic ligand oxidation
681
catalysts. J. Am. Chem. Soc. 2016, 138 (9), 2933-2936.
682
(60) Kuramochi, H.; Kawamoto, K.; Miyazaki, K.; Nagahama, K.; Maeda, K.; Li,
683
X.-W.; Shibata, E.; Nakamura, T.; Sakai, S.-I. Determination of physicochemical
684
properties of tetrabromobisphenol A. Environ. Toxicol. Chem. 2008, 27 (12),
685
2413-2418. 31
ACS Paragon Plus Environment
Environmental Science & Technology
686
(61) Feng, Y.; Colosi, L. M.; Gao, S.; Huang, Q.; Mao, L. Transformation and
687
removal of tetrabromobisphenol A from water in the presence of natural organic
688
matter via laccase-catalyzed reactions: Reaction rates, products, and pathways.
689
Environ. Sci. Technol. 2013, 47 (2), 1001-1008.
690
(62) Zhong, Y.; Liang, X.; Zhong, Y.; Zhu, J.; Zhu, S.; Yuan, P.; He, H.; Zhang, J.
691
Heterogeneous UV/Fenton degradation of TBBPA catalyzed by titanomagnetite:
692
Catalyst characterization, performance and degradation products. Water Res. 2012, 46
693
(15), 4633-4644.
694
(63) Emelianenko, M.; Torrejon, D.; DeNardo, M. A.; Socolofsky, A. L.; Ryabov, A.
695
D.; Collins, T. J. Estimation of rate constants in nonlinear reactions involving
696
chemical inactivation of oxidation catalysts. J. Math. Chem. 2014, 52, 1460-1476.
697
(64) Truong, L.; DeNardo, M. A.; Kundu, S.; Collins, T. J.; Tanguay, R. L. Zebrafish
698
assays as developmental toxicity indicators in the green design of TAML oxidation
699
catalysts. Green Chem. 2013, 15 (9), 2339-2343.
700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 32
ACS Paragon Plus Environment
Page 32 of 39
Page 33 of 39
Environmental Science & Technology
716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752
Figure Legends Figure 1. The molecular structure (a) and optimized geometry (b) of the prototype Fe(III)-TAML activator. The optimized calculation was conducted at B3LYP/6-311G (d,p) level. All atoms except hydrogen are symbolled and labeled. Figure 2. X-ray diffraction patterns of LDH with Fe(III)-TAML loadings of (a) 0 µmol g-1, (b) 2.86 µmol g-1, (c) 12.32 µmol g-1, (d) 22.38 µmol g-1, (e) 87.15 µmol g-1 and (f) 176.82 µmol g-1, respectively. The basal spacings of LDH increase as 2 theta values decrease. Figure 3. Simulation of the existing configuration of Fe(III)-TAML anion in LDH interlayer with Fe(III)-TAML loadings of (a) 22.38 µmol g-1, (b) 87.15 µmol g-1 and (c) 176.82 µmol g-1. The Fe(III)-TAML molecule tends to arrange into tilted configuration as the increasing accommodation of more Fe(III)-TAML molecules. The distance between 22O and 24O is 7.7 Å. Figure 4. Calculated and experimental IR spectra: calculated IR spectrum of (a) Fe(III)-TAML; observed IR spectra of (b) Fe(III)-TAML in solid phase, (c) Fe(III)-TAML/LDH powder, and (d) dried synthesized LDH. The experimental IR spectra were all obtained using KBr pellet technique. Figure 5. Degradation kinetics of TBBPA catalyzed by Fe(III)-TAML and Fe(III)-TAML/LDH at (a) pH 8, (b) pH 9 and (c) pH 10. Lines represent the fitted pseudo-first-order kinetic curves for different pH levels. The determined coefficients of determination from the overall regression analysis R2 are greater than 0.96. Experimental conditions: the initial concentrations of TBBPA, Fe(III)-TAML, H2O2, and LDH were 18.4 µM, 1.0 µM, 2.0 mM and 0.045 g L-1, respectively. The Fe(III)-TAML/LDH with Fe(III)-TAML loading of 22.38 µmol g-1 was chosen to degrade TBBPA. To avoid any pH fluctuation, both the pHs of TBBPA/Fe(III)-TAML/LDH solution and H2O2 stock solution were pre-adjusted to desired pH by addition of 0.1 M NaOH or 0.1 M HClO4 solution. Error bars represent standard deviations (n = 3). Figure 6. Proposed catalytic pathway of TBBPA by Fe(III)-TAML/H2O2. Molecular ion clusters obtained from mass spectra are shown below their structures, respectively.
33
ACS Paragon Plus Environment
Environmental Science & Technology
Figure 1.
34
ACS Paragon Plus Environment
Page 34 of 39
Page 35 of 39
Environmental Science & Technology
Figure 2.
35
ACS Paragon Plus Environment
Environmental Science & Technology
Figure 3.
36
ACS Paragon Plus Environment
Page 36 of 39
Page 37 of 39
Environmental Science & Technology
(d) LDH 1630 1370
Relative Intensity
(c) Fe(III)-TAML/LDH 1370
1459
1386
1641
1477 1562
(b) Fe(III)-TAML
1393 1626 1455 1575 (a) DFT: Fe(III)-TAML 1475 1513 1439 1379
1000
1200
1400
1636 1597
1600
1800 -1
Wavenumber (cm )
Figure 4.
37
ACS Paragon Plus Environment
2000
Environmental Science & Technology
1.2
Fe(III)-TAML Fe(III)-TAML/LDH Model Fit for Fe(III)-TAML Model Fit for Fe(III)-TAML/LDH
1.0 0.8 Ct/C0
Page 38 of 39
pH = 8 0.6 0.4 0.2 0.0 0
20
40
1.2
100
120
Fe(III)-TAML Fe(III)-TAML/LDH Model Fit for Fe(III)-TAML Model Fit for Fe(III)-TAML/LDH
1.0 0.8 Ct/C0
60 80 Time (s) (a)
pH = 9 0.6 0.4 0.2 0.0 0
10
20
1.2
50
60
Fe(III)-TAML Fe(III)-TAML/LDH Model Fit for Fe(III)-TAML Model Fit for Fe(III)-TAML/LDH
1.0 0.8 Ct/C0
30 40 Time (s) (b)
pH = 10 0.6 0.4 0.2 0.0 0
5
10
15 20 Time (s) (c)
25
Figure 5.
38
ACS Paragon Plus Environment
30
Page 39 of 39
Environmental Science & Technology
Figure 6.
39
ACS Paragon Plus Environment