Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

9
Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A Bin Yang, Guang-Guo Ying * , Zhi-Feng Chen, Jian-Liang Zhao, Fu-Qiang Peng, Xiao-Wen Chen State Key Laboratory of Organic Geochemistry, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, Guangzhou 510640, China article info Article history: Received 4 November 2013 Received in revised form 19 May 2014 Accepted 30 May 2014 Available online 11 June 2014 Keywords: Ferrate(VI) Tetrabromobisphenol A Bisphenol A Oxidation Suspended particles abstract Ferrate(VI) (Fe(VI)) oxidative removal of various organic micropollutants mainly depends on the reactivity of Fe(VI) to target micropollutants and coexisting constituents present in source water. This study evaluated the potential of Fe(VI) oxidation of the brominated flame retardant tetrabromobisphenol A (TBBPA) by using reaction kinetics, products identification and toxicity evaluation, and investigated the influencing effects of humic acid and clay particles on Fe(VI) removal of TBBPA in comparison with bisphenol A (BPA). The obtained apparent second-order rate constants (k app ) for Fe(VI) reaction with TBBPA ranged from 7.9(±0.3) 10 3 M 1 s 1 to 3.3(±0.1) 10 1 M 1 s 1 with the half-life (t 1/2 ) ranging from 1.7 s to 419.3 s at pH 7.0-10 for an Fe(VI) concentration of 10 mg L 1 . Easier oxidation by Fe(VI) was observed for TBBPA than for BPA. Fe(VI) can destroy and transform the TBBPA molecule through b-scission reaction, yielding the chemical species of low bromine- substituted products. More importantly, the oxidation of TBBPA by Fe(VI) led to the loss of its multiple hormonal activities (androgenic, antiestrogenic and antiandrogenic activ- ities). The organic component humic acid decreased the TBBPA and BPA reactions with Fe(VI), while the inorganic component montmorillonite had no effect on their removal within the tested concentrations. Increasing the Fe(VI) dosage can reduce the effects of soluble organic matter and clay particles present in source waters on the degradation process, leading to the complete removal of target micropollutants. © 2014 Elsevier Ltd. All rights reserved. 1. Introduction Ferrate(VI) (Fe(VI)) is an emerging environmentally friendly water treatment chemical due to its dual functions as an oxidant and a subsequent coagulant/precipitant as ferric hy- droxide (Jiang and Lloyd, 2002; Sharma, 2002, 2011; Lee et al., 2004; Sharma et al., 2011). Fe(VI) can achieve a complete removal of various emerging organic micropollutants during water and wastewater treatment (Lee et al., 2009; Sharma, 2013). The removal efficiency mainly depends on the reac- tivity of Fe(VI) to target micropollutants and coexisting con- stituents present in source water (Lee and von Gunten, 2010). Fe(VI) has been known to react with electron-rich organic moieties, such as phenols, anilines, amines, and olefins through electrophilic oxidation mechanism (Hu et al., 2009; Lee et al., 2009; Yang et al., 2012; Sharma, 2010, 2013). The corresponding second-order reaction rate constants range from >1 to 10 5 M 1 s 1 in aqueous solution (Lee et al., 2009; Sharma, 2013). However, the coexisting constituents present * Corresponding author. Tel./fax: þ86 20 85290200. E-mail addresses: [email protected], [email protected] (G.-G. Ying). Available online at www.sciencedirect.com ScienceDirect journal homepage: www.elsevier.com/locate/watres water research 62 (2014) 211 e219 http://dx.doi.org/10.1016/j.watres.2014.05.056 0043-1354/© 2014 Elsevier Ltd. All rights reserved.

Transcript of Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

Page 1: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

ww.sciencedirect.com

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9

Available online at w

ScienceDirect

journal homepage: www.elsevier .com/locate/watres

Ferrate(VI) oxidation of tetrabromobisphenol A incomparison with bisphenol A

Bin Yang, Guang-Guo Ying*, Zhi-Feng Chen, Jian-Liang Zhao,Fu-Qiang Peng, Xiao-Wen Chen

State Key Laboratory of Organic Geochemistry, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences,

Guangzhou 510640, China

a r t i c l e i n f o

Article history:

Received 4 November 2013

Received in revised form

19 May 2014

Accepted 30 May 2014

Available online 11 June 2014

Keywords:

Ferrate(VI)

Tetrabromobisphenol A

Bisphenol A

Oxidation

Suspended particles

* Corresponding author. Tel./fax: þ86 20 852E-mail addresses: guangguo.ying@gmail.

http://dx.doi.org/10.1016/j.watres.2014.05.0560043-1354/© 2014 Elsevier Ltd. All rights rese

a b s t r a c t

Ferrate(VI) (Fe(VI)) oxidative removal of various organic micropollutants mainly depends

on the reactivity of Fe(VI) to target micropollutants and coexisting constituents present in

source water. This study evaluated the potential of Fe(VI) oxidation of the brominated

flame retardant tetrabromobisphenol A (TBBPA) by using reaction kinetics, products

identification and toxicity evaluation, and investigated the influencing effects of humic

acid and clay particles on Fe(VI) removal of TBBPA in comparison with bisphenol A (BPA).

The obtained apparent second-order rate constants (kapp) for Fe(VI) reaction with TBBPA

ranged from 7.9(±0.3) � 103 M�1 s�1 to 3.3(±0.1) � 101 M�1 s�1 with the half-life (t1/2) ranging

from 1.7 s to 419.3 s at pH 7.0-10 for an Fe(VI) concentration of 10 mg L�1. Easier oxidation

by Fe(VI) was observed for TBBPA than for BPA. Fe(VI) can destroy and transform the TBBPA

molecule through b-scission reaction, yielding the chemical species of low bromine-

substituted products. More importantly, the oxidation of TBBPA by Fe(VI) led to the loss

of its multiple hormonal activities (androgenic, antiestrogenic and antiandrogenic activ-

ities). The organic component humic acid decreased the TBBPA and BPA reactions with

Fe(VI), while the inorganic component montmorillonite had no effect on their removal

within the tested concentrations. Increasing the Fe(VI) dosage can reduce the effects of

soluble organic matter and clay particles present in source waters on the degradation

process, leading to the complete removal of target micropollutants.

© 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Ferrate(VI) (Fe(VI)) is an emerging environmentally friendly

water treatment chemical due to its dual functions as an

oxidant and a subsequent coagulant/precipitant as ferric hy-

droxide (Jiang and Lloyd, 2002; Sharma, 2002, 2011; Lee et al.,

2004; Sharma et al., 2011). Fe(VI) can achieve a complete

removal of various emerging organic micropollutants during

water and wastewater treatment (Lee et al., 2009; Sharma,

90200.com, guang-guo.ying@gig

rved.

2013). The removal efficiency mainly depends on the reac-

tivity of Fe(VI) to target micropollutants and coexisting con-

stituents present in source water (Lee and von Gunten, 2010).

Fe(VI) has been known to react with electron-rich organic

moieties, such as phenols, anilines, amines, and olefins

through electrophilic oxidation mechanism (Hu et al., 2009;

Lee et al., 2009; Yang et al., 2012; Sharma, 2010, 2013). The

corresponding second-order reaction rate constants range

from >1 to 105 M�1 s�1 in aqueous solution (Lee et al., 2009;

Sharma, 2013). However, the coexisting constituents present

.ac.cn (G.-G. Ying).

Page 2: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9212

in source water are also responsible for a rapid Fe(VI) con-

sumption, which determine its ability to remove organic

micropollutants. For example, Zhang et al. (2012) have shown

that humic acid and SiO2�3 notably reduced the bisphenol A

(BPA) removal, tert-butanol slightly decreased the BPA

removal, and the existence of HCO�3 slightly enhanced the BPA

removal by Fe(VI). Yang and Ying (2013) have shown that

humic acid, Mn2þ and NaCl significantly decreased the

removal efficiency of sunscreen agent benzophenone-3 by

Fe(VI), while Cu2þ enhanced the removal. However, NHþ4 , NO�

3 ,

Fe3þ and Fe2þ had no effects on benzophenone-3 removal

within the tested concentrations. Lee and von Gunten (2010)

also found that ammonia, nitrite, and bromide have little ef-

fects on the 17a-ethinylestradiol removal during Fe(VI) treat-

ment of a secondary wastewater effluent at pH 8. Thus, the

coexisting constituents present in source water play an

important role in the removal of organic micropollutants by

Fe(VI) treatment.

Suspended particleswith a particle size ranging from1mm

down to 1 mm are also an important coexisting constituent

ubiquitous in surface water and wastewater. Suspended par-

ticles are an important carrier for some pollutants and path-

ogens. The smaller the particle size, the greater the total

surface area per unit mass of particle, and resulting in the

higher pollutant load that is likely to be carried (Gregory, 2006;

Templeton et al., 2008). For example, the brominated flame

retardant tetrabromobisphenol A (TBBPA) was detected in

quantifiable levels in 15 of the 17 wastewater treatment

plants' sewage sludge samples, in a concentration range of not

detected (nd)-472 ng/g dw (Gorga et al., 2013). Thus, activated

sludge-associated TBBPA may still penetrate to the disinfec-

tion stage of tertiary treatment under certain circumstances.

The detection of dissolved and particle-bound taste and odor

compounds such as geosmin, b-ionone, 2-isobutyl-3-

methoxypyrazine, 2-isopropyl-3-methoxypyrazine and 2-

methyl-isoborneol in the Swiss lake waters were found to

affect the oxidation pretreatment for source water in the

drinking water treatment process (Peter et al., 2009). There-

fore, it is necessary to investigate the effects of suspended

particles on the fate of emerging organic micropollutants

during the oxidation treatment processes.

In this study, TBBPA was selected as the target compound

as it is the most widely used brominated flame retardant in

the world and has also been widely detected in surface water

(Covaci et al., 2009; He et al., 2013). TBBPA is synthesized from

BPA through bromination reaction, so they have very similar

molecular structures and functional groups. Compared with

BPA, TBBPA is characterized by a low to moderate water

solubility, and a moderately high octanol/water partition

coefficient which is dependent on ionization state at

different pH solutions (Kuramochi et al., 2008). When

released into the environment, TBBPA will have a tendency

to partition into sediment and soil, to bind to suspended

particles and to the lipid fraction of biota (He et al., 2013).

Besides, humic acid (HA) and montmorillonite (MMT) are

often used as the model organic and inorganic suspended

particles in experiments (Templeton et al., 2008). Therefore,

the objective of this study was to assess the potential of

Fe(VI) oxidation of TBBPA by using reaction kinetics, prod-

ucts identification and toxicity evaluation. The reactivity of

TBBPA with Fe(VI) was simultaneously compared with BPA,

which has been in-depth researched by Lee et al. (2005a) and

Li et al. (2008). Besides, the influences of suspended particles

such as HA and MMT on Fe(VI) removal of TBBPA and BPA

were investigated both in phosphate buffer solutions and

natural waters.

2. Experimental section

2.1. Standards and materials

Chemical standards tetrabromobisphenol A (TBBPA, 99%) and

bisphenol A (BPA, 99%) were purchased from Dr. Ehrenstorfer

GmbH. Diammonium 2,20-azinobis-(3-ethylbenzothiazoline-6-sulfonate) (ABTS, 98%), and humic acid (HA) were purchased

from SigmaeAldrich; while montmorillonite K 10 (MMT) was

from Acros Organics. Potassium ferrate (Fe(VI), >95%) was

prepared by wet chemical synthesis (Delaude and Laszlo,

1996). Buffer chemicals and all other reagents used in the

experiments were of analytical grade. All reaction solutions

were prepared with Milli-Q water (>18 MU cm) from a Milli-

pore Water Purification System. The stock solutions of Fe(VI)

(0.5e1.0 mM) were freshly prepared by dissolving solid Fe(VI)

in Milli-Q water (pH z 9.2). The Fe(VI) stock solution was

quickly filtered through a 0.45 mm hydrophilic poly-

ethersulfone (PES) syringe filter (Shanghai ANPEL, China) and

then standardized spectrophotometrically at 510 nm

(ε ¼ 1150 M�1 cm�1). The stock solutions of TBBPA and BPA

were prepared inmethanol at the concentration of 100mg L�1.

The stock solutions of HA and MMT were prepared in 10 mM

phosphate buffer solutions (pH ¼ 8.0) with their concentra-

tions of 500 mg L�1 and 1000 mg L�1 respectively.

Surface water used in the experiments was sampled from

the Zhujiang River in Guangzhou, South China. The initial pH,

UV254, UV400, total organic carbon (TOC), conductivity, alka-

linity and turbidity of the natural water sample were 6.98,

0.093, 0.016, 3.43 mg L�1, 429 mS cm�1, 3.15 mM (HCO�3 ) and 7.0

NTU, respectively. Part of the above water sample was filtered

through a 0.45 mm cellulose-nitrate membrane to remove the

suspended particles. The dissolved organic carbon (DOC) and

turbidity of the filtered water sample was 3.04 mg L�1 and 5.4

NTU. All the water samples were used within 24 h. Besides,

the trace target pollutants in surface water did not interfere

with the experiments because of the high spiked concentra-

tions for TBBPA and BPA (2 mM).

2.2. Removal experiments of TBBPA and BPA at variousFe(VI) doses

The removal experiments of TBBPA and BPA with various

Fe(VI) doses were performed in 10 mM phosphate buffer so-

lutions (pH 7.0 and 8.0) and room temperature (24 ± 1 �C). In a

series of 25 mL amber volumetric flasks, a target compound

(2 mM) was firstly spiked into the buffer solution, and then the

filtered and standardized stock solution of Fe(VI) was added or

not added to yield various concentrations of 0e30 mM. The

mixed solutions were shaken in order to have sufficient re-

action. After 3 h reaction in the darkness, 1 mL of the reaction

solution was sampled and quenched with 0.1 mL of a

Page 3: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9 213

thiosulfate solution (5 mM). The residual concentrations for

TBBPA and BPA were then determined by the high perfor-

mance liquid chromatography (HPLC) which was described

later in Section 2.4. All removal experiments were performed

in duplicate.

The effect of selected particles on Fe(VI) oxidation of TBBPA

and BPA was evaluated by dosing various concentrations of

HA or MMT (0e40 mg L�1) to the reaction solutions used in the

above removal experiments. When HA and MMT were spiked

into the buffer solutions containing the target compounds of

TBBPA or BPA, the experiments were further divided into two

treatment groups: one followed with immediate addition of

the Fe(VI) stock solutions to oxidation; the other with mixing

of the solutions in a shaking incubator (SKY-211BG,

150 r min�1, 24 ± 1 �C) for 12 h, and followed with addition of

Fe(VI) stock solutions to reaction. The mixing process by the

shaking incubator was aimed to achieve the interaction of the

selected particles with the target compounds as occurred in

the actual water environment. After 3 h reaction in the dark-

ness, the residual concentrations of TBBPA and BPA were

determined by HPLC. The pH variation of all the experiments

was found below 0.05 units.

2.3. Kinetic experiments of TBBPA oxidation by Fe(VI)

The kinetics of TBBPA oxidation by Fe(VI) were studied under

pseudo-first-order conditions in the pH range of 7.0e10

(10 mM phosphate buffer solutions). In the 150 mL reaction

mixture solutions, the initial concentration of Fe(VI) was

25 mM, and the concentration for TBBPA was 2 mM. The ex-

perimentswere performed in a 200mL beaker equippedwith a

magnetic stirrer (700 r min�1) at the room temperature

(24 ± 1 �C). The reactions were initiated by adding an aliquot of

Fe(VI) stock solution to the suspensions containing TBBPA

under rapid mixing. At certain time intervals, 5 mL of the re-

action solution was sampled for the measurement of residual

Fe(VI) concentrations using a ABTS method at 415 nm (Lee

et al., 2005b), and 1 mL of the reaction solution was also

sampled and quenched with a thiosulfate solution (5 mM,

0.1mL) tomeasure residual concentrations of TBBPA by HPLC,

then the apparent second-order rate constants (kapp) of Fe(VI)

reaction with TBBPA were calculated by plotting the natural

logarithm of the TBBPA concentrations with the Fe(VI) expo-

sure (Yang et al., 2011a; Yang and Ying, 2013). All kinetic ex-

periments were carried out in duplicate and the pH variation

was below 0.1 units. Besides, the 150 mL kinetic reaction so-

lutions were also quenched with 5 mL thiosulfate solution

(5 mM) at a setting reaction time (0, 5, 10, 30, 50, 70 and 90 s) to

identify the reaction products and evaluate the changes of

hormonal activities.

Kinetic reaction experiments were also performed in nat-

ural water samples with and without filtration at pH 8.0

(20 mM borate buffer) to assess the effect of suspended par-

ticles present in natural waters on Fe(VI) removal of TBBPA

and BPA. Nevertheless, the difference with the above kinetic

process was that the target compound was firstly spiked into

the river water samples with and without filtration. Then the

solutions were mixed in a shaking incubator for 12 h to have

sufficient contact, and followed by addition of Fe(VI) to start

the kinetic reaction. Preliminary experiments showed that the

concentrations of the target compounds did not change dur-

ing mixing (Data not shown).

2.4. Chemical analysis

The water samples were characterized as follows. The UV

absorbance was measured with a Helios Alpha spectropho-

tometer (Thermo Spectronic, Cambridge, UK). The conduc-

tivity and pH values were determined using a Thermo Orin 5

star pH meter (Thermo Fisher Scientific, USA), which was

calibrated using standard reagents (1314 mS cm�1 and pH 4.0,

7.0, and 10.0, Thermo China). TOC and DOC were measured

using a Shimadzu TOC-V and TNM-1 analyzer (Shimadzu

Scientific Instruments, Columbia, MD) (Yang et al., 2012; Yang

and Ying, 2013). Alkalinity and turbidity were measured by

standardized titrimetric analysis and spectrophotometric

method respectively.

For the determination of the target compounds TBBPA and

BPA, the quenched BPA reaction solutions were directly

analyzed by HPLC, while TBBPA in the quenched reaction so-

lutions was analyzed after extraction. Briefly, the quenched

TBBPA reaction solutions were firstly extracted with

dichloromethane (3 � 0.3 mL), then the extracts were

concentrated under a gentle nitrogen, and re-dissolved in

0.5 mL acetonitrile for HPLC analysis. The instrument used for

analysis of TBBPA and BPA was an Agilent 1200 series HPLC

equippedwith a diode array detector (DAD) and a fluorescence

detector (FLD). Chromatographic separation was performed

on a Zorbax Eclipse XDB-C18 column (150 � 4.6 mm, 5 mm).

The column temperature was set at 30 �C. The mobile phase

consisted of acetonitrile and Milli-Q water (80:20 for TBBPA

and 60:40 for BPA) at a flow rate of 1 mL min�1. The injection

volume was 100 mL. The UV wavelength for TBBPA detection

was 210 nm. The excitation wavelength and emission wave-

length of FLD for BPA detection was 230 nm and 305 nm

respectively. The limit of quantitation was 50 mg L�1 for

TBBPA, and 10 mg L�1 for BPA. The recovery of extractions

ranged between 80.4% and 91.5% independent of the solution

pH values (7.0e10) (Fig. S1 and Fig. S2).

For reaction products identification and hormonal activ-

ities transformation, 150 mL quenched reaction solutions

were adjusted to pH 2 with 1 M HCl. Reaction products were

extracted by vigorous shaking with 3 � 20 mL dichloro-

methane. Each extract was passed through an anhydrous

Na2SO4 column to remove water. The extract was concen-

trated under a rotary evaporator and re-dissolved in 1 mL

methanol, then filtered through a 0.22 mm nylon syringe filter

into a 2 mL amber glass vial which was kept at �20 �C until

analysis. The recovery rate for the TBBPA was about 73e82%

during this extraction process.

The reaction products of TBBPA were analyzed by rapid

resolution liquid chromatographyetandem mass spectrom-

etry (RRLCeMS/MS). The RRLC-MS/MS instrument used in this

analysis was an Agilent 1200 series RRLC connected to an

Agilent 6460 triple quad mass spectrometer, with an Agilent

Zorbax SB-C18 column (100� 3mm, 1.8 mm). Themobile phase

consisted of (A)Milli-Qwater and (B)methanol, whichwas run

at a flow rate of 0.3 mL min�1. The gradient was programmed

as follows: 30% B at 0 min, increased to 80% at 25 min, kept in

80% B to 45 min, and the post time was 5 min. The column

Page 4: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

Fig. 1 e Apparent second-order rate constants and

associated model simulation for the reaction of TBBPA,

BPA and 4-bromophenol with Fe(VI) as a function of pH

(7.0e10) at the room temperature (24 ± �C). Kinetic data for

BPA and 4-bromophenol are calculated from Lee et al.

(2005a, b).

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9214

temperature was set at 40 �C. The mass spectrometer was

operated under negative electrospray ionization with MS2

Scan mode and Product Ion mode. MS2 Scan mode had a

fragmentor voltage of 135 V with mass scan range of 50e1200

amu. Product Ion mode had a fragmentor voltage of 135 V and

collision energy of 24 V with mass scan range of 50e800 amu.

The ionization source conditions were listed as follows: the

drying gas flow 3 mL min�1 at 350 �C, sheath gas flow

12 mL min�1 at 350 �C and the nebulizer pressure 40 psi.

2.5. Bioassays

Multiple hormonal activities (estrogenic and androgenic ac-

tivities, antiestrogenic and antiandrogenic activities) were

measured for the extracts of reaction solutions during Fe(VI)

oxidation of TBBPA by using our previously developed in vitro

recombinant yeast bioassays (Zhao et al., 2011). In brief, the

yeast with human estrogen receptor (hER-a) gene integrated

into expression plasmid was used for yeast estrogen screen

(YES) and yeast antiestrogen screen (YAES) bioassays, and the

yeast with human androgen receptor (hAR-a) gene integrated

into expression plasmid was used for yeast androgen screen

(YAS) and yeast antiandrogen screen (YAAS) bioassays. Stan-

dard chemicals 17b-estradiol, dihydrotestosterone, tamoxifen

Table 1 e Second-order rate constants for the reaction of Fe(VI

Compound pKa Reactingspecies

k11/

Tetrabromobisphenol

A(TBBPA)

7.5

8.5

HFeO�4 þ XH2/

HFeO�4 þ XH�/

HFeO�4 þ X2�

1.1(±01.8(±01.9(±0

Bisphenol A(BPA) 9.6

10.2

HFeO�4 þ XH2/

HFeO�4 þ XH�/

HFeO�4 þ X2�

8.2(±08.0(±02.6(±0

4-bromophenol 9.34 HFeO�4 þ XH/

HFeO�4 þ X�

8.0(±01.2(±0

and flutamide were used as the positive control for YES, YAS,

YAES and YAAS (Fig. S3). The potency of hormonal activity of a

sample was calculated from the ratio of median effective

concentration (EC50) of the sample and the EC50 of the corre-

sponding positive control with a four-parameter log-logistic

model using SigmaPlot 10.0 software. The levels of these hor-

monal activities were then expressed as 17b-estradiol equiva-

lent (EEQ), dihydrotestosterone equivalent (DEQ), tamoxifen

equivalent (TEQ), and flutamide equivalent (FEQ), respectively.

3. Results and discussion

3.1. Kinetics for the reaction of TBBPA with Fe(VI)

Second-order reaction rate law (Eq. (1)) was used to describe

the Fe(VI) oxidation of TBBPA in the phosphate buffer solu-

tions (Lee et al., 2005a, 2009; Yang et al., 2011a; Yang and Ying,

2013). The second-order rate constants (kapp) were determined

under pseudo-first-order conditions with Fe(VI) concentra-

tions in excess to TBBPA and calculated by using Eq. (2). The

values of kapp for the reaction of Fe(VI) with TBBPA as a

function of pH (7.0e10) are presented in Fig. 1. The kapp of the

reaction decreased with increasing pH values. These pH-

dependent variations in kapp could be quantitatively

modeled by species-specific reactions between Fe(VI) species

and TBBPA species by using Eq. (3), whichwas programmed by

using the least-squares nonlinear regressions in the software

SigmaPlot 10.0. Consequently, the species-specific second-

order rate constants (k) of the reaction betweenHFeO�4 and the

neutral or the dissociated TBBPA are listed in Table 1 and

shown in Fig. 1 (R2 ¼ 0.99). Kinetic data for BPA and 4-

bromophenol were calculated from Lee et al. (2005a) as a

comparison to TBBPA (Fig. 1, Table 1).

�d½TBBPA�=dt ¼ kapp½FeðVIÞ�½TBBPA� (1)

ln�½TBBPA��½TBBPA�0� ¼ �kapp

Zt

0

½FeðVIÞ�dt (2)

kapp½FeðVIÞ�tot½TBBPA�tot ¼X

i ¼ 1;2j ¼ 1;2; 3

kijaibj½FeðVIÞ�tot½TBBPA�tot

(3)

The k of the reaction between HFeO�4 and the dissociated

TBBPA, BPA and 4-bromophenol was greater than the reaction

) with TBBPA, BPA and 4-bromophenol.

k12/k13 k(pH 7) k(pH 8) Ref.

.1) � 104/

.1) � 104/

.07) � 104

7.9 � 103 2.6 � 103 This study

.1) � 102/

.2) � 104/

.2) � 105

6.4 � 102 4.1 � 102 Lee et al. (2005a, b)

.4) � 101/

.04) � 1048.6 � 101 8.7 � 101 Lee et al. (2005a, b)

Page 5: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9 215

between HFeO�4 and the neutral compounds (Table 1), which

indicates the reaction between HFeO�4 and the dissociated

target compounds controls the overall reaction. The obtained

k value for the reaction of HFeO�4 with each of TBBPA's acid-

base species was 1.1(±0.1) � 104 M�1 s�1 (k11),

1.8(±0.1) � 104 M�1 s�1 (k12) and 1.9(±0.07) � 104 M�1 s�1 (k13)

respectively. The k12 and k13 of dissociated TBBPA with HFeO�4

are lower than the ones of BPA with HFeO�4 .Since an aromatic

ring is substituted with halogens, the reactivity of the sub-

stance decreases due to a decreased electron density of the

ring. However, the k11 of neutral TBBPA with HFeO�4 is higher

than that of BPA, this may be due to the oxidationmechanism

which is different from direct attack at the ring. Thus, the kappvalue of Fe(VI) reaction with TBBPA was greater than that of

the BPA in the pH range of 7.0e9.0, but TBBPA had a less kappvalue than the BPA in the pH range of 9.0e10. Moreover, the

kapp values for both TBBPA and BPA reaction with Fe(VI) are

greater than that of the 4-bromophenol, indicating that the

reactivity of these compounds with Fe(VI) can be enhanced by

the introduction of substituent groups into benzene ring.

TBBPA showed a more appreciable reactivity with Fe(VI).

The obtained kapp for Fe(VI) reaction with TBBPA ranged from

7.9(±0.3) � 103 M�1 s�1 to 3.3(±0.1) � 101 M�1 s�1 with the half-

life (t1/2) ranging from 1.7 s to 419.3 s at pH 7.0e10 for an Fe(VI)

concentration of 10 mg L�1 (Fig. 1). The t1/2 for Fe(VI) reaction

with TBBPA (1.7 s, pH 7.0) is magnitude lower than that of the

permanganate (29.8 s, pH 7.0) (Pang et al., 2014), manganese

dioxide (1.79 min, pH 4.5) (Lin et al., 2009) and photochemical

(30 min, pH 7.1) (Eriksson et al., 2004), indicating that Fe(VI)

treatment technology appears to be a promising tool for the

removal of brominated flame retardant TBBPA in water.

3.2. Removal of TBBPA and BPA by Fe(VI)

Removal of TBBPA and BPA at various Fe(VI) doses was

investigated in pH 7.0 and 8.0 phosphate buffer solutions.

Fig. 2 shows the relative changes in the residual concentra-

tions of TBBPA and BPA as a function of the Fe(VI) dose

Fig. 2 e Oxidation of TBBPA and BPA in 10 mM phosphate

buffer solutions as a function of the Fe(VI) doses (0e30 mM).

Experimental conditions: pH ¼ 7.0 and 8.0, T ¼ 24 ± 1 �C,[TBBPA/BPA] ¼ 2.0 mM, and contact time 3 h.

(0e30 mM). The residual concentrations of TBBPA and BPA

decreased with the increasing dose of Fe(VI). The dosed

amount of Fe(VI) for oxidation of TBBPAwasmuch lower than

that used for BPA either at pH 7.0 or 8.0. For example, 12.5 mM

of Fe(VI) can completely remove TBBPA in the pH 7.0 buffer

solution, but more than 30 mM of Fe(VI) was needed for BPA

removal, which illustrated the easier oxidation of TBBPA

molecule than BPA by Fe(VI). This may be because TBBPA had

a higher kapp value in pH 7.0 and 8.0 than that of BPA (Table 1).

Besides, the dosed amounts of Fe(VI) for removal of TBBPA and

BPA were higher at pH 7.0 more than at pH 8.0 (Fig. 2). For

example, 7.5 mM Fe(VI) was required to completely remove

TBBPA at pH 8.0, but 12.5 mMFe(VI) was required at pH 7.0. This

can be explained by the higher stability of the Fe(VI) in the

phosphate solution at pH 8.0, which resulted in a higher Fe(VI)

exposure, and a lower Fe(VI) dosage for the removal of TBBPA

and BPA in pH 8.0.

3.3. Oxidation products

The oxidation products of Fe(VI) reaction with TBBPA at

different sampling times (0e90 s) were tentatively identified

by using RRLC-MS/MS (Table 2, Fig. S4eS6 and Table S1). Table

2 lists the primary molecular ions, product ions, relative

abundances and proposed structures of the 2 products (A and

B) and parent compound TBBPA. The two products A and B

had amolecular ion peak atm/z of 308.9 and 322.9 and product

ions 290.9, 80.9 and 78.8 (Fig. S6), which were considered as 4-

(2-hydroxyisopropyl)-2,6-dibromophenol and 4-(2-

methoxylisopropyl)-2,6-dibromophenol, respectively (Feng

et al., 2013; Pang et al., 2014). But it should be noted that

products A and B were also presented at trace levels in the

initial reaction solution (0 s). During the Fe(VI) oxidation pro-

cess, A and B gradually increased and then decreased (Fig. S5).

Thus, A and B could be considered as the reaction products.

Besides, another 10 peaks were also presented in the chro-

matograms which have characteristic bromine isotope ions

(Fig. S4, Table S1), corresponding to dibromo-, tribromo- and

tetrabromo- compounds, but no meaningful chemical struc-

tures were proposed. These 10 peaks may be the oxidation

products of TBBPA or impurities and their potential oxidation

products. Consequently, the detection of a large number of

low bromine-substituted products that are important for

further Fe(VI) oxidative degradation.

The oxidation products of BPA by Fe(VI) have been studied

by Li et al. (2008) using LC-MS/MS and GC-MS/MS techniques.

In the previous study, nine products were identified as 4-

isopropanolphenol, p-isopropylphenol, p-isopropenylphenol,

oxalic acid, propanedioic acid, 4-isopropyl-cyclohexa-2,5-

dienone, p-hyroxyacetophenone, styrene, and (1-phenyl-1-

butenyl)benzene. Because the electron donating hydroxyl

group increases the electron density of each aromatic ring,

causing the bond connecting with two aromatic rings to be

more vulnerable. So, the initial oxidation step for BPA by Fe(VI)

was considered to be the bond cleavage of the two phenyl

groups by Fe(VI) attack, yielding p-isopropylphenol, phenol, 4-

isopropanolphenol, and (1-phenyl-1-butenyl)benzene. These

intermediate products can be further degraded by Fe(VI)

oxidation, producing styrene, p-hydroxyacetophenone, 4-

isopropyl-cyclohexa-2,5-dienone, propanedioic acid, and

Page 6: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

Table 2 eMain fragment ions (m/z) of TBBPA and products in total ion chromatogram identified bymeans of RRLC-ESI-MS/MS in negative mode.

Compound Retentiontime (min)

m/z valuesa (Relative abundance) Proposed structureb

MS2 scan Product ion

A 14.454 306.8(62.5%)/308.9c(100%)/

310.9(62.5%)

308.9(4.4%)/290.9(100%)/80.9(16.4%)/78.8(19.7%)

B 21.352 320.9(55.7%)/322.9(100%)/

324.9(54.8%)

322.9(0.5%)/290.9(100%)/80.9(23.2%)/78.8(21.2%)

3

TBBPA 28.720 538.7(23.8%)/540.7(83.3%)/

542.8(100%)/544.7(82.5%)/

546.7(20.6%)

542.8(100%)/527.6(1.8%)/462.4(0.3%)/460.7(0.3%)/

447.7(3.6%)/445.7(3.3%)/419.6(1.1%)/417.6(0.6%)/

290.7(1.8%)

/288.9(0.7%)/80.9(1.1%)/78.8(1.0%)

4

a Fig. S4 and Fig. S6.b The uncertain substituent positions of atoms are displayed in dotted line.c Molecular ion peak.

Fig. 3 e Proposed reaction schemes for oxidation of TBBPA

by Fe(VI).

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9216

oxalic acid. Although the parent compound BPA could be

completely removed within less than 5 min, the oxidation

products would still persist in the final reaction solutions (Li

et al., 2008).

On the basis of the kinetic information, products identifi-

cation and the mechanism of Fe(VI) reaction with BPA, a

plausible reaction scheme for Fe(VI) oxidation of TBBPA was

proposed in Fig. 3. Initially, Fe(VI) oxidizes thephenolmoietyof

TBBPA by one electron transfer generating a phenoxyl radical

andFe(V) as thefirst step (Rushet al., 1995;Huanget al., 2001; Li

et al., 2008; Yanget al., 2011b). Thephenoxy radical is stabilized

byelectron resonancewithin thephenol ring and forms radical

R1, which supported by the calculation of charge distribution

and spin densities via molecular modeling (Lin et al., 2009;

Feng et al., 2013; Pang et al., 2014). Radical R1 undergoes b-

scission (cleavage between one of the benzene rings and the

isopropyl group) and releases a new radical 2,6-dibromo-4-

isopropylphenol carbocation (R2). R2 may react with water to

form the product A, or substitution with the methanol cosol-

vent (stock solution was prepared in methanol) to yield the

product B (Feng et al., 2013). The pathway of product A has

been extensively reported in the oxidative degradation of

TBBPA by permanganate, manganese dioxide, UV/Fenton, and

UV irradiation (Eriksson et al., 2004; Lin et al., 2009; Zhonget al.,

2012; Pang et al., 2014). For product B, this is probably not valid

in a potential realwater treatment. In addition, radicalsR1 and

R2 may be transformed to other low bromine-substituted

products through further Fe(VI) oxidation.

3.4. Changes in hormonal activities

The multiple hormonal activities (estrogenic and androgenic

activities, antiestrogenic and antiandrogenic activities) for the

extracts of reaction solutions during Fe(VI) oxidation were

determined by using in vitro recombinant yeast bioassays, as

shown in Table 3. TBBPA showed in vitro androgenic, anti-

androgenic and anti-estrogenic activities, but no estrogenic

activity as shown in Table 3, which is consistent with the

previous studies (Kitamura et al., 2005; Christen et al., 2010;

Page 7: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

Table 3 e The changes in hormonal activities during Fe(VI) oxidation process of TBBPA.

<Reactiontime(s)

TBBPA(mg L�1)

Estrogen(EEQ, ng L�1)a

Antiestrogen(TEQ, ng L�1)a

Androgen(DEQ, ng L�1)a

Antiandrogen(FEQ, ng L�1)a

0 105.1 NDb 176.5 ± 85.6 2286.9 ± 147.8 300.8 ± 24.5

5 73.6 ND ND 2194.2 ± 234.2 57.5 ± 9.7

10 57.4 ND ND 526.8 ± 157.3 76.2 ± 5.1

30 22.6 ND ND 276.9 ± 117.3 20.1 ± 8.4

50 10.9 ND ND 178.3 ± 62.0 14.6 ± 6.2

70 6.3 ND ND 5.3 ± 1.0 18.6 ± 14.3

90 0.3 ND ND ND 9.8 ± 4.2

a EEQ: 17b-estradiol equivalent; TEQ: tamoxifen equivalent; DEQ: dihydrotestosterone equivalent; FEQ: flutamide equivalent.b Not detected.

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9 217

Huang et al., 2013). During Fe(VI) oxidation process, the

determined androgenic and antiandrogenic activities

decreased with increasing reaction time, and the reduction of

the androgenic and antiandrogenic activities was in propor-

tion to the decreasing residual TBBPA concentrations (Fig. S7

and Fig. S8). This indicates that the determined androgenic

and antiandrogenic activities are related to the levels of

TBBPA, suggesting the reaction products exhibiting negligible

androgenic and antiandrogenic activities. Besides, no estro-

genic and antiestrogenic activities were found throughout the

Fe(VI) degradation process, which indicates that the reaction

products also have no estrogenic and antiestrogenic activities.

Consequently, the oxidation of TBBPA by Fe(VI) could lead to

the loss of its hormonal activities.

Fig. 4 e Effect of selected suspended particles (HA, MMT) on

the removal of TBBPA and BPA in 10 mM phosphate buffer

solutions by Fe(VI). HA: humic acid; MMT:

montmorillonite. Experimental conditions: [TBBPA/

BPA]¼ 2.0 mM, pH¼ 8.0, T¼ 24 ± 1 �C, and contact time 3 h.

They were 2.5 mM and 5 mM of Fe(VI) for TBBPA and BPA to

achieve a 70% removal.

3.5. The effect of suspended particles on Fe(VI) oxidation

The influence of selected suspended particles (HA and MMT)

on TBBPA/BPA removal during Fe(VI) treatment was investi-

gated in pH 8.0 phosphate buffer solutions. There were no

obvious difference in TBBPA/BPA concentrations pretreated

with andwithoutmixing process, indicating no sorption of HA

and MMT with TBBPA/BPA occurred. Fig. 4 shows the relative

changes in the residual concentrations of TBBPA/BPA as a

function of particles doses.With the increasing amount of HA,

the residual concentration of TBBPA/BPA was increased. For

example, when adding 20mg L�1 of HA, the removal efficiency

of TBBPA and BPA dropped from 70% to 33.2% and 28.4%

respectively, indicating the spiked HA decreased the Fe(VI)

oxidation of TBBPA/BPA. Nevertheless, the removal of TBBPA/

BPA was not affected by the presence of MMT within the

tested concentrations (0e40 mg L�1). Therefore, when

applying a lower dose of Fe(VI) for treatment, the dissolved

organic matter would decrease the removal of target pollut-

ants by Fe(VI), but clay minerals would have no significant

effect.

The influence of natural particles on Fe(VI) removal of

TBBPA/BPA was also investigated by kinetic oxidation exper-

iments with the filtered and unfiltered natural water samples.

Compared to the relatively lower dose of Fe(VI) used in the

above batch experiments, the dosage of Fe(VI) (25 mM) was in

excess to TBBPA/BPA. Alongwith the increasing reaction time,

the residual concentrations of TBBPA/BPA by Fe(VI) in unfil-

tered and filtered water samples as well as phosphate buffer

solutions decreased, showing similar trends (Fig. 5). But, the

residual concentrations of TBBPA/BPA in the unfiltered water

samples were slightly more than in the filtered water sample,

indicating the presence of suspended particles in natural

water can reduce the removal efficiency of TBBPA/BPA by

Fe(VI) (Lee et al., 2009; Lee and von Gunten, 2010). Besides, the

unfiltered natural water had initial TOC and turbidity of

3.4mg L�1 and 7.0 NTU, but the filteredwater had the TOC and

turbidity of 3.0 mg L�1 and 5.4 NTU. So the residual concen-

trations of TBBPA and BPA in both filtered and unfilteredwater

samples were more than the phosphate buffer solutions

because of the high dissolved organic matter in natural wa-

ters. Besides, other coexisting constituent such as inorganic

ions andmetal cations present in filtered and unfiltered water

samples may be responsible for the above differences (Yang

and Ying, 2013). Finally, Fig. 5 also shows that dosing more

Fe(VI) can achieve complete removal of the target pollutants

in the natural water. This is because the competition between

target pollutants and dissolved organic matter can disappear

rapidly after the electron-rich organic moieties present in

dissolved organic matter are consumed by the selective

Page 8: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

Fig. 5 e - Removal of TBBPA and BPA by Fe(VI) during the

treatment of natural river water. Experimental conditions:

[TBBPA/BPA]0 ¼ 2 mM, [Fe(VI)]0 ¼ 25 mM, pH ¼ 8.0 (20 mM

borate buffer), T ¼ 24 ± �C.

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9218

oxidant Fe(VI) (Lee and von Gunten, 2010; Yang and Ying,

2013).

4. Conclusions

This study demonstrated that by Fe(VI) oxidation the bromi-

nated flame retardant TBBPA could be more easily removed

than BPA. The determined kapp of the Fe(VI) reaction with

TBBPA was greater than that of BPA in the studied pH range.

Based on the tentatively identified products, the Fe(VI)

oxidation of TBBPA generates the main products of 4-(2-

hydroxyisopropyl)-2,6-dibromophenol and 4-(2-

methoxylisopropyl)-2,6-dibromophenol. The Fe(VI) oxidation

process decreased the multiple hormonal activities (andro-

genic and antiandrogenic activities) of TBBPA, with no

detectable estrogenic and antiestrogenic activities. It was

shown that the natural organic matters such as HA in water

would decrease the Fe(VI) reactions with TBBPA and BPA,

while the inorganic particles such as clay minerals had no

effect on the removal within the tested concentrations.

Complete removal of organic compounds such as TBBPA can

be achieved by increasing the Fe(VI) doses in order to reduce

the effects of suspended particles present in natural waters.

Acknowledgments

The authors would like to acknowledge the financial support

from the National Natural Science Foundation of China (NSFC

U1133005, 41273119 and 41121063), CAS Key Projects (KZZD-

EW-09), Guangdong Natural Science Foundation

(S2012040007592), and China Postdoctoral Science Foundation

funded project (2012M511845). The authors would like to

thank Dr. Junyi Li (Lancaster Environment Centre) and Dr. Bo

Liu for their useful discussions. This is a Contribution No. 1920

from GIG CAS.

Appendix A. Supplementary data

Supplementary data related to this article can be found at

http://dx.doi.org/10.1016/j.watres.2014.05.056.

r e f e r e n c e s

Christen, V., Crettaz, P., Oberli-Schrammli, A., Fent, K., 2010.Some flame retardants and the antimicrobials triclosan andtriclocarban enhance the androgenic activity in vitro.Chemosphere 81 (10), 1245e1252.

Covaci, A., Voorspoels, S., Abdallah, M.A.-E., Geens, T., Harrad, S.,Law, R.J., 2009. Analytical and environmental aspects of theflame retardant tetrabromobisphenol-A and its derivatives. J.Chromatogr. A 1216, 346e363.

Delaude, L., Laszlo, P., 1996. A novel oxidizing reagent based onpotassium ferrate(VI). J. Org. Chem. 61 (18), 6360e6370.

Eriksson, J., Rahm, S., Green, N., Bergman, Å., Jakobsson, E.,2004. Photochemical transformations of tetrabromobisphenolA and related phenols in water. Chemosphere 54 (1),117e126.

Feng, Y., Colosi, L.M., Gao, S., Huang, Q., Mao, L., 2013.Transformation and removal of tetrabromobisphenol A fromwater in the presence of natural organic matter via laccase-catalyzed reactions: reaction rates, products, and pathways.Environ. Sci. Technol. 47 (2), 1001e1008.

Gorga, M., Martinez, E., Ginebreda, A., Eljarrat, E., Barcelo, D.,2013. Determination of PBDEs, HBB, PBEB, DBDPE, HBCD,TBBPA and related compounds in sewage sludge fromCatalonia (Spain). Sci. Total Environ. 444, 51e59.

Gregory, J., 2006. Particles in Water: Properties and Processes.Taylor & Francis Group, LLC, pp. 1e2.

He, M.J., Luo, X.J., Yu, L.H., Wu, J.P., Chen, S.J., Mai, B.X., 2013.Diasteroisomer and enantiomer-specific profiles ofhexabromocyclododecane and tetrabromobisphenol A in anaquatic environment in a highly industrialized area, SouthChina: Vertical profile, phase partition, and bioaccumulation.Environ. Pollut. 179, 105e110.

Hu, L., Martin, H.M., Arcs-Bulted, O., Sugihara, M.N., Keatlng, K.A.,Strathmann, T.J., 2009. Oxidation of carbamazepine by Mn(VII)and Fe(VI): reaction kinetics and mechanism. Environ. Sci.Technol. 43 (2), 509e515.

Huang, H., Sommerfeld, D., Dunn, B.C., Eyring, E.M., Lloyd, C.R.,2001. Ferrate(VI) oxidation of aqueous phenol: kinetics andmechanism. J. Phys. Chem. A 105 (14), 3536e3541.

Huang, G.-Y., Ying, G.-G., Liang, Y.-Q., Zhao, J.-L., Yang, B., Liu, S.,Liu, Y.-S., 2013. Hormonal effects of tetrabromobisphenol Ausing a combination of in vitro and in vivo assays. Comp.Biochem. Physiol Part C 157 (4), 344e351.

Jiang, J.Q., Lloyd, B., 2002. Progress in the development and use offerrate(VI) salt as an oxidant and coagulant for water andwastewater treatment. Water Res. 36 (6), 1397e1408.

Kitamura, S., Suzuki, T., Sanoh, S., Kohta, R., Jinno, N.,Sugihara, K., Yoshihara, S., Fujimoto, N., Watanabe, H.,Ohta, S., 2005. Comparative study of the endocrine-disruptingactivity of bisphenol A and 19 related compounds. Toxicol. Sci.84 (2), 249e259.

Kuramochi, H., Kawamoto, K., Miyazaki, K., Nagahama, K.,Maeda, K., Li, X.-W., Shibata, E., Nakamura, T., Sakai, S.-I.,2008. Determination of physicochemical properties of

Page 9: Ferrate(VI) oxidation of tetrabromobisphenol A in comparison with bisphenol A

wat e r r e s e a r c h 6 2 ( 2 0 1 4 ) 2 1 1e2 1 9 219

tetrabromobisphenol A. Environ. Toxicol. Chem. 27 (12),2413e2418.

Lee, Y., von Gunten, U., 2010. Oxidative transformation ofmicropollutants during municipal wastewater treatment:comparison of kinetic aspects of selective (chlorine, chlorinedioxide, ferrateVI, and ozone) and non-selective oxidants(hydroxyl radical). Water Res. 44 (2), 555e566.

Lee, Y., Cho, M., Kim, J.Y., Yoon, J., 2004. Chemistry of ferrate(Fe(VI)) in aqueous solution and its applications as a greenchemical. J. Ind. Eng. Chem. 10 (1), 161e171.

Lee, Y., Yoon, J., Von Gunten, U., 2005a. Kinetics of the oxidationof phenols and phenolic endocrine disruptors during watertreatment with ferrate (Fe(VI)). Environ. Sci. Technol. 39 (22),8978e8984.

Lee, Y., Yoon, J., von Gunten, U., 2005b. Spectrophotometricdetermination of ferrate (Fe(VI)) in water by ABTS. Water Res.39 (10), 1946e1953.

Lee, Y., Zimmermann, S.G., Kieu, A.T., von Gunten, U., 2009.Ferrate (Fe(VI)) application for municipal wastewatertreatment: a novel process for simultaneous micropollutantoxidation and phosphate removal. Environ. Sci. Technol. 43(10), 3831e3838.

Li, C., Li, X.Z., Graham, N., Gao, N.Y., 2008. The aqueousdegradation of bisphenol A and steroid estrogens by ferrate.Water Res. 42 (1e2), 109e120.

Lin, K., Liu, W., Gan, J., 2009. Reaction of tetrabromobisphenol A(TBBPA) with manganese dioxide: kinetics, products, andpathways. Environ. Sci. Technol. 43 (12), 4480e4486.

Pang, S.-Y., Jiang, J., Gao, Y., Zhou, Y., Huangfu, X., Liu, Y., Ma, J.,2014. Oxidation of flame retardant tetrabromobisphenol a byaqueous permanganate: reaction kinetics, brominatedproducts, and pathways. Environ. Sci. Technol. 48 (1),615e623.

Peter, A., Koester, O., Schildknecht, A., von Gunten, U., 2009.Occurrence of dissolved and particle-bound taste and odorcompounds in Swiss lake waters. Water Res. 43 (8), 2191e2200.

Rush, J.D., Cyr, J.E., Zhao, Z.W., Bielski, B.H.J., 1995. The oxidationof phenol by ferrate(VI) and ferrate(V) - a pulse-radiolysis andstopped-flow study. Free Radic. Res. 22 (4), 349e360.

Sharma, V.K., 2002. Potassium ferrate(VI): an environmentallyfriendly oxidant. Adv. Environ. Res. 6 (2), 143e156.

Sharma, V.K., 2010. Oxidation of nitrogen-containing pollutantsby novel ferrate(VI) technology: a review. J. Environ. Sci.

Health Part a-Toxic/Hazard. Subst. Environ. Eng. 45 (6),645e667.

Sharma, V.K., 2011. Oxidation of inorganic contaminants byferrates (VI, V, and IV)-kinetics and mechanisms: a review. J.Environ. Manag. 92 (4), 1051e1073.

Sharma, V.K., 2013. Ferrate(VI) and ferrate(V) oxidation of organiccompounds: kinetics and mechanism. Coord. Chem. Rev. 257(2), 495e510.

Sharma, V.K., Luther Iii, G.W., Millero, F.J., 2011. Mechanisms ofoxidation of organosulfur compounds by ferrate(VI).Chemosphere 82 (8), 1083e1089.

Templeton, M.R., Andrews, R.C., Hofmann, R., 2008. Particle-associated viruses in water: Impacts on disinfectionprocesses. Crit. Rev. Environ. Sci. Technol. 38 (3), 137e164.

Yang, B., Ying, G.-G., 2013. Oxidation of benzophenone-3 duringwater treatment with ferrate(VI). Water Res. 47 (7), 2458e2466.

Yang, B., Ying, G.-G., Zhang, L.-J., Zhou, L.-J., Liu, S., Fang, Y.-X.,2011a. Kinetics modeling and reaction mechanism offerrate(VI) oxidation of benzotriazoles. Water Res. 45 (6),2261e2269.

Yang, B., Ying, G.-G., Zhao, J.-L., Zhang, L.-J., Fang, Y.-X.,Nghiem, L.D., 2011b. Oxidation of triclosan by ferrate: reactionkinetics, products identification and toxicity evaluation. J.Hazard. Mater. 186 (1), 227e235.

Yang, B., Ying, G.-G., Zhao, J.-L., Liu, S., Zhou, L.-J., Chen, F., 2012.Removal of selected endocrine disrupting chemicals (EDCs)and pharmaceuticals and personal care products (PPCPs)during ferrate(VI) treatment of secondary wastewatereffluents. Water Res. 46 (7), 2194e2204.

Zhang, P., Zhang, G., Dong, J., Fan, M., Zeng, G., 2012. Bisphenol Aoxidative removal by ferrate (Fe(VI)) under a weak acidiccondition. Sep. Purif. Technol. 84, 46e51.

Zhao, J.-L., Ying, G.-G., Yang, B., Liu, S., Zhou, L.-J., Chen, Z.-F.,Lai, H.-J., 2011. Screening of multiple hormonal activities insurface water and sediment from the Pearl River system,South China, using effect-directed in vitro bioassays. Environ.Toxicol. Chem. 30 (10), 2208e2215.

Zhong, Y., Liang, X., Zhong, Y., Zhu, J., Zhu, S., Yuan, P., He, H.,Zhang, J., 2012. Heterogeneous UV/Fenton degradation ofTBBPA catalyzed by titanomagnetite: catalystcharacterization, performance and degradation products.Water Res. 46 (15), 4633e4644.