2015 Sharme J Et Al (Chemical Eng J

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Chemical Engineering Journal 276 (2015) 193–204

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Oxidative removal of Bisphenol A by UV-C/peroxymonosulfate (PMS):


Kinetics, influence of co-existing chemicals and degradation pathway
Jyoti Sharma a, I.M. Mishra a,⇑, Dionysios D. Dionysiou b,⇑, Vineet Kumar a,c
a
Department of Chemical Engineering, Indian Institute of Technology, Roorkee, Roorkee 247667, India
b
Environmental Engineering and Science Program, Department of Biomedical, Chemical and Environmental Engineering (DBCEE), 705 Engineering Research Center,
University of Cincinnati, Cincinnati, OH 45221-0012, United States
c
Present Address: Department of Chemical Engineering, Indian School of Mines, Dhanbad, Jharkhand, India

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 UV activated PMS oxidation is


effective for BPA oxidation and
mineralization.
 BPA degradation rate is not affected
by Cl (5 mM), but it is adversely
affected by HCO3 and HA.
 The possible radical adduct formation
and degradation pathway are also
proposed.
 The proposed pathway supports
sulfate radical attack as the main
route.

a r t i c l e i n f o a b s t r a c t

Article history: In the present study, a sulfate radical-based advanced oxidation process was applied for the degradation of
Received 25 December 2014 an industrial chemical and suspected endocrine disruptor, Bisphenol A (BPA). UV-C (k = 254 nm; 40 W
Received in revised form 1 April 2015 power; Io = 1.26 lE s1) activated peroxymonosulfate (PMS) was used as an oxidant. The effect of operating
Accepted 2 April 2015
parameters (initial concentration of BPA, dose of PMS, initial solution pH (pHo), and water matrix compo-
Available online 9 April 2015
nents such as chloride (Cl), bicarbonate (HCO 3 ) ions and humic acid (HA) was evaluated. At the initial pH
of reaction mixture (5.15) and room temperature (29 ± 3 °C), the optimum dosage of PMS was found to be
Keywords:
0.66 mM, giving a BPA removal of 96.7 ± 0.05% and a total organic carbon (TOC) removal of 72.5 ± 0.05%
Advanced oxidation process (AOP)
Humic acid
after 360 min of irradiation. With an increase in initial BPA concentration and PMS dosage greater than
Pathway 0.66 mM, the BPA and TOC removal decreased. The extent of BPA removal increased with an increase in
Peroxides pHo (3 6 pHo 6 12) of the reaction mixture. The degradation of BPA followed pseudo-first-order kinetics
Peroxymonosulfate (PMS) and the apparent first order rate constant for BPA was found to be 0.025 min1 at the optimum oxidation
Persulfate conditions (CBPA = 0.22 mM, CPMS = 0.66 mM, pH = 5.15, temperature = 29 ± 3 °C). The Cl ions have negli-
gible inhibition effect on the BPA removal. However, the HCO 3 and HA inhibited the BPA oxidation under
UV-C irradiation. The identification of intermediates and final products was carried out with HPLC, GC/MS
and FTIR, and a degradation pathway was proposed. The present study reveals that the UV-C/PMS oxidation
process is effective for BPA removal under real water/wastewater conditions.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction

⇑ Corresponding authors. A growing number of contaminants are entering water bodies


E-mail addresses: immishra49@gmail.com, imishfch@iitr.ac.in (I.M. Mishra),
ranging from traditional heavy metals to emerging micropollutants
dionysios.d.dionysiou@uc.edu (D.D. Dionysiou). such as natural hormones, pesticides, pharmaceuticals and

http://dx.doi.org/10.1016/j.cej.2015.04.021
1385-8947/Ó 2015 Elsevier B.V. All rights reserved.
194 J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204

industrial chemicals. Bisphenol A (BPA) is an industrial chemical of degradation than the Cobalt/H2O2 system [35]. Some dissolved
which is largely used in the production of various items such as metal ions used for activation of oxidants in water/wastewater
epoxy resins, polycarbonate, etc. It is identified as an endocrine treatment may lead to several health problems such as asthma,
disrupting chemical (EDC), and it finds its way in the water bodies pneumonia, and other lung ailments [32,36–38]. However, the
either directly or indirectly via degradation of various BPA contain- use of UV irradiation is considered to be environment-friendly.
ing materials such as plastic bottles, containers, toys, food/soft PMS is slightly more powerful than hydrogen peroxide as an oxi-
drink cans, packaged food, etc. [1–3]. The presence of EDCs in the dant (Eo(HSO 
5 /HSO4 ) = +1.82 V; Eo(H2O2/H2O) = 1.76 V) [39].
water bodies affects the aquatic lives and other life-forms through When irradiated by a low pressure mercury lamp (k = 254 nm),
consumption of contaminated water [4]. The concentration of BPA PMS generates SO 
4 and HO radicals through the cleavage of
in water/wastewaters can be as low as 0.01 lg/L to as high as peroxide bond [40,41], as explained by Eq. (1), and subsequently
17 mg/L in some industrial effluents and landfill leachates produces SO2 + 
4 , H and HO species through reaction with water
[2,3,5,6]. It has affinity to estrogen acceptor and thereby induces (Eq. 3):
feminization in a number of aquatic organisms [7,8]. The United þ 2
States Environmental Protection Agency (USEPA) and the SO
4 þ H2 O ! H þ SO4 þ HO

ð3Þ
European Food Safety Authority (EFSA) set a reference dosage of Several studies reported on the effect of water matrix compo-
50 lg/kg body weight/day of BPA for human consumption [9]. nents such as natural organic matter (NOM) and alkalinity (carbon-
Conventional physical–chemical water treatment processes ate/bicarbonate) on the degradation of target compounds by AOPs.
such as adsorption, filtration, chlorination, and coagulation/floccu- NOM represents a complex mixture of naturally occurring
lation as well as biological processes are not able to remove BPA organic compounds in the surface and ground water. Humic acid
from water satisfactorily, they rather transfer it from one phase (HA) is generally used as a representative NOM [13,16,42]. The
to another (adsorption, filtration, coagulation/flocculation) [10]. effect of water matrix components on the oxidation process is
The limitations of conventional processes to remove BPA have poorly understood. It is assumed traditionally that they have neg-
led to efforts to explore alternative methods such as advanced oxi- ative impact on the reaction system [30]. However, a few studies
dation processes (AOPs). A number of AOP studies described showed that the presence of water matrix components enhances
Fenton/photo-Fenton degradation [11,12], sonochemical degrada- the efficiency of the sulfate radical-based AOPs and reduces the
tion [13], photochemical/photocatalytic oxidation [14–16], ozona- efficiency of HO radical-based AOPs under real water treatment
tion [17], and hybrid processes [18,19] for BPA removal from conditions [43,44]. The composition of HA is source-dependent
water. and varies greatly. Therefore, its impact on the oxidation of the tar-
The AOPs are based on highly reactive species such as hydroxyl get compound may sometimes be confounding, depending on its
(HO) and/or sulfate (SO 4 ) radicals. These radicals are capable of source, the method of oxidation used for the target compound,
degrading the organic molecules and yielding intermediates and and the presence and composition of various other water matrix
CO2. Radical species can be produced mainly from the thermal, components [45,46].
UV and transition metal activation of oxidants such as hydrogen The reaction intermediates and degradation products play a
peroxide, persulfate (PS) and peroxymonosulfate (PMS). crucial role in the toxicity of treated water. Some studies have
PS and PMS are comparatively newer oxidants in surface water reported on the byproducts of AOPs during/at the end of the pro-
treatment applications, but they have long been used as in-situ cess [11,12,47,48]. The mechanism of degradation and the forma-
chemical oxidation agents in ground water applications. PS and tion of intermediates and byproducts have been studied mostly
PMS were used by researchers for the removal of pharmaceuticals for HO radical-based processes. Not much has been reported on
[20], herbicides [21], phenols [22], perfluorinated compounds [23], these aspects for the UV-C activated PMS-based AOP for the treat-
disinfection of Escherichia coli [24] and chlorinated compounds ment of BPA contaminated water.
[25,26] in water treatment and groundwater remediation. The present study, reports on the degradation of 0.22 mM
PS and PMS are favored for use in environmental remediation (50 mg/L) BPA solution by UV-C/PMS AOP under natural reaction
applications because of their easier activation and higher stability mixture pH and room temperature conditions. The optimum
than that of hydrogen peroxide under typical water/ground water dosage of PMS was used for the mineralization and degradation
treatment conditions [27]. The high standard reduction potential of BPA. The effects of pHo and water matrix components
(2.5–3.1 V) of SO4 radicals and wider operating pH range [28,29] (Cl, HCO3 and HA) on the removal of BPA were also studied.
makes SO 4 radicals comparatively more effective and selective The intermediates and the end-products were analyzed and identi-
for oxidation of target compound in water/wastewater treatment fied by GC–MS and FTIR spectroscopy, and a reasonable degrada-
[30–33]. tion pathway is also proposed.
PMS is a triple salt of potassium monopersulfate
(2KHSO5KHSO4K2SO4), a product manufactured by DuPont, and 2. Experimental
available commercially as OXONEÒ. PMS is considered to be a
green oxidant because it is benign, as are most of the byproducts 2.1. Reagents and chemicals
of its reactions [34]. The decomposition rate of organic compounds
with PMS at room temperature is low, but it can be enhanced Bisphenol-A (BPA, CAS 80-05-7), molecular weight = 228 with a
through activation by photolysis or thermolysis (Eq. 1) or by using purity >99% was purchased from Sigma Aldrich and used as
transition metals such as cobalt (Eq. 2), thereby producing highly received. PMS, anhydrous sodium sulfate, HA (sodium salt),
reactive SO 
4 and HO radicals dichloromethane and NaHCO3 were obtained from Hi Media,
Mumbai; acetonitrile (>99% purity) from Merck India and NaCl
hv=D
HSO5 ƒƒƒƒƒ! SO
4 þ HO

ð1Þ and NaOH were obtained from RFCL, New Delhi. Sulfuric acid
(H2SO4) was purchased from Fine Chemicals, New Delhi. Ethanol
and tert-butyl alcohol (TBA) were obtained from Merck (India)
HSO5 þ Metal ion ðMn Þ ! Mnþ1 þ SO
4 þ OH

ð2Þ
and Loba Chemie (Mumbai), respectively. A quenching agent,
Cobalt/PMS is reported to have higher rate of degradation than sodium thiosulfate (Na2S2O35H2O, 99.5%) (New India Chemical
Cobalt/H2O2 system. Cobalt can effectively activate PMS and gener- Enterprises, Kochi) was used to stop the reaction in the sample.
ate SO
4 radicals at enhanced rate, which in turn gives higher rate Millipore water was used from laboratory (Milli-Q Biocel) system.
J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204 195

2.2. Batch experiments of the oxidation process (the details of the extraction method are
reported in Supporting Information Material) [51,52].
The present study was conducted using a borosilicate glass
photo-reactor (1 L). 450 mL of BPA (0.22 mM or 50 mg/L) aqueous 2.6. Identification of dominant radicals
solution was irradiated by using 5  8 W low pressure mercury
UV-C light (k = 254 nm). The radiation intensity as measured by In order to understand the principal oxidation radicals, HO or
ferrioxalate actinometry was 1.26 lE s1. A magnetic stirrer and SO
4 or both effecting the degradation of BPA, ethanol and TBA
a cooling fan system were used for proper mixing and cooling dur- were used as scavenging agents for HO and SO 4 radicals
ing the degradation experiments. Initial pH of the BPA solution and [31,15,53]. The scavenger, oxidant and BPA were used in the molar
the room temperature at the time of experimentation were 6.2 and ratio of 200:1:1. The radical identification studies were conducted
29 ± 3 °C, respectively. The degradation experiments were con- at optimum conditions (CBPA = 0.22 mM, CPMS = 0.66 mM,
ducted for 360 min under these conditions. Based on preliminary t = 360 min), but with excess of the scavenging agent [15].
tests, 1 mL sodium thiosulfate solution (1 M) per mL of sample
was used to quench the reaction. During the reaction, a solid pro- 3. Results and discussion
duct was also detected. The quenched sample was filtered through
a 0.20 lm syringe filter (Axiva, India). The filtrate was used for 3.1. Effect of PMS dosage
HPLC analysis and the filtered solid matter was subjected to FTIR
analysis. The experiments were carried out at the natural pH of From Fig. 1, it is seen that 96.7% BPA was removed in 360 min
the reaction mixture except during the studies on the effect of ini- reaction time when no scavenger was used. However, in the pres-
tial pH on degradation. In such experiments, the initial pH of the ence of ethanol, the BPA removal decreased to 35% only. In the
solution was adjusted using 0.1 N NaOH or 0.1 N H2SO4, as presence of TBA, the BPA removal was 85%. The scavenging rate
required. Control experiments were carried out in the dark as well constants for ethanol vary from 1.2 to 2.8  109 M1 s1 for HO
as in the absence of PMS. A loss of 9.78% of BPA was detected under and from 1.6 to 7.7  107 M1 s1 for SO 4 [31,53]. For TBA, the
UV-C irradiation in the absence of PMS, while up to 42% BPA was rate constant is approximately 1000 fold greater for HO (3.8–
removed by PMS alone at the room temperature (29 ± 3 °C) and 7.6  108 M1 s1) than that for SO (4–9.1  105 M1 s1)
4
natural reaction mixture pH of the system without UV-C, as the [31,53].
PMS gets activated by temperature also [49]. The ratio of scavenging rates of oxidant radicals by ethanol and
TBA varies between 1.3 and 3.5 for HO radicals and between 18
2.3. HPLC measurements and 200 for SO4 radicals. The scavenging by ethanol is about one
order of magnitude faster for HO radicals and about two orders
The BPA concentration in the filtrate was determined by HPLC of magnitude faster for SO4 radicals than the scavenging rates of
(Waters India), equipped with a C18 column (3.9 mm  150 mm) the respective radicals by TBA. Therefore, the BPA degradation is
and a UV detector set at 276 nm. The mobile phase was the mixed much lower in the presence of ethanol than that for TBA. In the
solvent of acetonitrile (HPLC grade) and water (50/50; v/v) at a case of TBA, SO4 radicals are scavenged at much lower rate than
flow rate of 1 mL/min. HO radicals, and thus, the SO 4 radicals present in the reaction
mixture are responsible for BPA degradation.
2.4. Other measurements This behavior can also be explained by calculating the utiliza-
tion efficiency of PMS after 360 min of reaction. The inset of
A TOC analyzer (TOC-VCPH, Shimadzu) and a pH meter Fig. 1 shows an insignificant utilization efficiency of PMS (9%)
(Cyberscan 510 pH) were used to determine the total organic car- due to poor degradation of BPA (35%), and a high consumption
bon (TOC) and pH of the treated samples (filtrate), respectively. of PMS because of high reaction rates of ethanol with SO 4 and
The samples were quenched by using 1 M sodium thiosulfate aque-
ous solution before TOC analysis. FTIR spectral measurements of
the solid product were carried out using a Thermo Nicolet (model
Magna 760) instrument (the details of the method are reported in
the Supporting Information Material) for the identification of func-
tional groups [50]. All the experiments were carried out in tripli-
cate and the average values are reported with the error bars
representing 95% confidence of interval.

2.5. GC/MS analysis

The analysis of the products was carried out by a GC/MS (Clarus


-680 model GC with Elite 5-MS, 30 m  0.25 mm capillary column
coupled with Clarus SQ8C model MS; Perkin Elmer, USA). Helium
was used as the carrier gas at a flow rate of 1.5 mL/ min. One
microliter aliquot samples were injected at 220 °C in splitless
mode. The temperature program for GC followed the sequence:
start at 40 °C, hold for 5 min; 40–200 °C (10 °C/min), hold for
5 min; 200–230 °C (5 °C/min), hold for 5 min. The transfer line
and ion source temperature were maintained at 200 °C and
170 °C, respectively. Qualitative analysis was performed in the
electron impact (EI) mode at 70 eV using the full scan mode in
the m/z range of 40–500 amu. The unknown peaks were identified Fig. 1. Effect of scavengers on the degradation of BPA at 29 ± 3 °C: —d— [PMS/BPA],
using the library of Wiley and NIST. Dichloromethane (DCM) was —s— [TBA/PMS/BPA], —.— [ethanol/PMS/BPA], (Inset: PMS utilization efficiency
used as the solvent to extract the BPA and the organic products for different systems at 360 min of reaction time).
196 J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204

HO radicals. In the presence of TBA, the PMS utilization efficiency 3.2. Effect of BPA concentration
increases to 55%, while the PMS utilization efficiency increases to
92% in the absence of alcohol (scavenging agent). Fig. 3 shows the variation of residual fractional BPA concentra-
The higher removal of BPA in the presence of TBA in comparison tion with the time of oxidation for four initial BPA concentration
to ethanol explains the selectivity of SO
4 radicals for BPA degrada- values of 0.04 mM (10 mg/L), 0.13 mM (30 mg/L), 0.22 mM
tion. The maximum efficiency in alcohol-free system explains the (50 mg/L) and 0.31 mM (70 mg/L). As the initial concentration
dominance and selectivity of SO 4 radicals [54]. Wang et al. [55] increased, the residual BPA concentration also increased: 0.1, 0.6,
also reported >90% utilization efficiency of PMS for As(III) removal 1.65 and 21 mg/L for an initial BPA concentration of 10, 30, 50
from groundwater at As(III)/PMS molar ratio of 1:2. and 70 mg/L, respectively. The retardation in BPA removal may
Fig. 2 shows the effect of PMS concentration (0.16 < CPMS be attributed to the requirement of a larger number of SO 4 radicals
< 1.62 mM) on the degradation of BPA. The experiments at room by the BPA molecules and the intermediates present in the solution
temperature (29 ± 3 °C) showed 42% removal of BPA without and the competition between BPA molecules and the intermedi-
UV-C irradiation. This means that the PMS was thermally ates for free radicals. These results are consistent with those
activated [49], and therefore, it also partly contributed to the reported in literature [15]. The initial degradation rates of the pho-
removal of BPA. The removal of BPA increased from 64.2% to tooxidation process were estimated from the slope of the plot of
96.7% when the CPMS was increased from 0.16 mM to 0.66 mM. BPA concentration versus time at t = 0 for the initial 60 min of reac-
At CPMS > 0.66 mM, the incremental removal of BPA was not tion time (Inset of Fig. 3). It can be seen that the initial degradation
significant. This is because of the scavenging of SO 4 and HO

rate of BPA increased from 2.2  104 mmol L1 min1 to
radicals by HSO 
5 , and the formation of less reactive SO5 radicals 10.0  104 mmol L1 min1 with an increase in the initial BPA
[56]: concentration from 0.04 mM to 0.31 mM. This is because of the
availability of the larger number of BPA molecules to oxidizing free
HSO5 þ HO ! SO
5 þ OH

ð4Þ radicals and thereby enhancing the rate of reaction [58]. In accor-
HSO5 þ SO
4 ! SO2
4
þ
þH ð5Þ dance with these observations, the degradation time for >90% BPA
removal is also found to have decreased for the initial BPA concen-
Thus, the optimum CPMS was found to be 0.66 mM. Wang and tration <0.13 mM.
Chu [41] also observed a similar trend during the degradation of
2,4,5-trichlorophenoxy acetic acid (0.1 mM) by UV activated PMS
in the presence of iron. Ji et al. [57] also found that the CuO acti- 3.3. Effect of initial solution pH on BPA degradation
vated PMS is effective at 1.5 mM PMS with 100% phenol (50 mg/
L) removal and 86% TOC removal in 60 min. The phenol removal The pH of the reaction mixture has profound effect on the total
efficiency decreased when PMS dosage was 1.25 mM. radical concentration and their formation in SO 4 based AOPs [40].

The degradation of BPA by UV-C/PMS system can be explained In order to investigate the effect of initial pH of reaction mixture on
by pseudo-first order reaction kinetics (ln(C/C0) = kt) [19,44]. BPA degradation by PMS under UV-C irradiation, experiments were
The nonlinear regression analysis shows a good fit to the experi- performed at five initial pH values of BPA reaction mixture (pHo:
mental data during the initial 60 min of reaction time (inset of 3.0, 5.0, 7.0, 10.0 and 12.0).
Fig. 2). The apparent first order rate constant for the removal of It is seen that the BPA removal increased with an increase in the
BPA at its initial concentration of 0.22 mM, varied from initial pH of the reaction mixture (Fig. 4): 74.3% (pHo = 3), 88.3%
0.003 min1 to 0.025 min1 with an increase in the PMS dosage (pHo = 5), 97.1% (pHo = 7), 97.9% (pHo = 10), and 98.8% (pHo = 12)
from an initial 0.16 mM to an optimum dosage of 0.66 mM, but as against 96.7% BPA removal at the natural reaction mixture pH.
decreased to 0.011 min1 at 1.62 mM PMS dosage. Based on the The time course of BPA removal at different pH values follows
above results, 0.66 mM PMS concentration was used in further the order: pHo3 < pHo5 < pHN  pHo7 < pHo10 < pHo12.
experimental study. The variation in removal with pHo can be explained on the basis
of three factors (a) absorption coefficient, (b) conversion of SO 4
radicals to HO radicals with pH, and (c) the rate of reaction of
radicals with deprotonated BPA.

Fig. 2. Effect of PMS concentration on the degradation efficiency of BPA. Fig. 3. Time course of BPA degradation under UV/PMS system [PMS]o = 0.66 mM;
[BPA]o = 0.22 mM. [PMS]o = —d— 0.16 mM, —s— 0.32 mM, —.— 0.66 mM, —D— [BPA]o = —d— 0.04 mM; —s— 0.13 mM; —.— 0.22 mM; —D— 0.31 mM. (Inset:
0.98 mM, —j— 1.30 mM, —h— 1.62 mM (Inset: Pseudo-first–order kinetics). Initial rate vs BPA concentration).
J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204 197

Fig. S2(a)). It is thermodynamically feasible for PMS (1.85 V) to oxi-


dize chloride ions (Cl) into less reactive chlorine species viz. Cl2/
2Cl (1.36 V) and Cl/HOCl (1.48 V) [24,62]. Therefore, the observed
higher degree of inhibition to BPA degradation at 10–20 mM Cl
concentration than that at 5 mM is due to the consumption of sul-
fate radicals by Cl ions [29], and the formation of less reactive
chlorine species Cl2, HOCl, Cl and Cl2 [24,62,63]. The chemical
reactions involved in (but not limited to) UV-C/PMS/BPA system
in the presence of Cl ions can be given as follows [24,64]:

HSO5 þ Cl ! SO2
4 þ HOCl ð6Þ

HSO5 þ 2Cl þ Hþ ! SO2
4 þ Cl2 þ H2 O ð7Þ
 2 
SO4 þ Cl ! SO4 þ Cl ð8Þ
  
Cl þ Cl $ Cl2 ð9Þ
  
Cl2 þ Cl2 ! Cl2 þ 2Cl ð10Þ

Cl2 þ H2 O ! HOCl þ Hþ þ Cl ð11Þ

Fig. 4. Effect of pH on the degradation efficiency of BPA. [BPA]o = 0.22 mM, HOCl ! Hþ þ OCl ðpKa ¼ 7:5Þ ð12Þ
[PMS]o = 0.66 mM, —D— pHN = 5.15, —s— pH = 3, —.— pH = 5, —j— pH = 7,
—h— pH = 10, —d— pH = 12. Similar trends have also been observed by Yuan et al. [64] for
sulfate radical-based (Co/PMS) AOP of Acid Orange 7.
As the pH of the reaction system increases, the absorption coef-
HA is shown to act as a photosensitizer, light filter and quencher
ficient of PMS ( PMS) increases ( PMS = 13.8–149.5 M1 cm1; pH
” ”
of free radicals. The overall effect of HA on the photodegradation of
6–12), with the enhanced photolysis of PMS, which, in turn,
BPA will, therefore, depend on the competition between the BPA
resulted in the increased concentration of SO 
4 and HO radicals
and the HA molecules [45]. It has also been reported that the dif-
[40]. The BPA molecule has two dissociation constants: pKa1 = 9.6
ferences in the quality and the type of the HA (or NOM) would give
and pKa2 = 10.2 [2]. BPA can dissociate as Bis(OH)(O) or Bis(O)2
different results on the degradation of organics [65,66]. In the pre-
in the reaction mixture. As indicated by the BPA speciation dia-
sent study, the HA is found to adversely affect the oxidation of BPA
gram (Supporting Information Material Fig. S1), the first deproto-
(Supporting Information Material Fig. S2(b)). The degradation of
nation of BPA starts at about pH 7 and the second one at about
BPA in the presence of 10 mg/L HA was 54.0% as compared to
pH 9. At pH < pKa1, undissociated BPA (i.e., Bis(OH)2) is predomi-
96.7% without HA. An increase in the HA concentration further
nantly present in the solution. In the pH range, pKa1 < pH < pKa2,
increases the degree of inhibition. Similar trends were also
the predominant species is Bis(OH)(O), and at pH > pKa2, the
observed by Zhang et al. [13]. At high concentration, the HA may
majority species is the Bis(O)2. The phenolate ions are reported
be in competition with BPA for SO 
4 and HO radicals, thereby
to have higher reaction rate constant in comparison to neutral
reducing the availability of oxidizing radicals for BPA degradation.
molecules of BPA [59]. It may, therefore, be suggested that all the
This slows down the BPA degradation rate. The increased concen-
above three factors act in a synergistic manner to effect the
tration of HA also lowers the solution transparency, thereby
removal of BPA from the solution.
inhibiting the UV-C irradiation to pass through the reaction
It has also been reported, that the degradation of butylated
mixture.
hydroxyl anisole increases as the pH increases from 3 to 11 in a
The bicarbonate ion (HCO 3 ) is an efficient hydroxyl and sulfate
PS/UV system [60], and the degradation of nitrobenzene improves
radical scavenger [53,67]. Therefore, the presence of bicarbonate
as the pH increases from 7 to 12 in a thermally activated PS system
ions would have an adverse effect on BPA degradation. The results
[61] due to conversion of SO 
4 to HO at alkaline pH [40]. During the
of the present study also support the above inference (Supporting
course of oxidation, the pH of the reaction mixture decreases from
Information Material Fig. S2(c)). The redox potential of bicarbonate
its initial value with its value dropping down to 3 for 5 < pHo < 10
ion is 1.78 V and its capability for scavenging SO 
4 and HO radicals,
and to 9 for pH 12. This is because of the acidic byproducts
are given by Eqs. (13) and (14) [53]:
obtained at the end of the oxidation (as explained in Section 4).
It can be seen that the BPA removal increases as pHo increases SO  2  þ
ðk13 ¼ 1:6  106 M1 s1 Þ
4 þ HCO3 ! SO4 þ CO3 þ H ð13Þ
from the acidic region to the alkaline region, although the differ- 
HO þ HCO3 ! CO
3 þ H2 O ðk14 ¼ 8:5  106 M1 s1 Þ ð14Þ
ence is not appreciable. On the basis of above results, further study
was carried out at the natural solution pH (pHN), because of the So the scavenging or reactive radical species (SO 
4 or HO ) is
marginal increment in BPA removal with an increase in pHo from responsible for the BPA removal inhibition. A comparative chart
the natural pH of the solution to the alkaline region. for the effect of water matrix components on the BPA removal by
UV-C/PMS system after 360 min of irradiation is illustrated in
3.4. Effect of co-existing water matrix chemicals Fig. 5. In comparison to 96.7% removal of BPA by UV-C/PMS alone,
the removal of BPA increases to 99.0% in the presence of 5 mM Cl
The components of the water matrix in ionized form also affect ion, but decreases to 54% and 70% in the presence of 5 mM HCO 3

the fate of the target compound. The effect of water matrix compo- ion and 5 mg/L HA (inset of Fig. 5), respectively. This means that
nents such as Cl and HCO the BPA degradation is not affected much in the chloride bearing
3 ions and HA on BPA removal by UV-C/
PMS oxidation was studied. The experiments were conducted in water, whereas the presence of HCO 3 and HA have measurable

the concentration range of 5–20 mM for Cl and HCO adverse effect on BPA degradation by PMS under UV-C irradiation.
3 ions, and
in the concentration range of 5-20 mg/L for HA. The results indi-
cated that the BPA degradation was not significantly influenced 3.5. Degree of mineralization and PMS consumption
at 5 mM Cl ion concentration. However, as the concentration of
Cl ions was increased to 10 and 20 mM, an inhibition in degrada- TOC is used as a measure of mineralization of target organic
tion of BPA was observed (Supporting Information Material compounds. The requirement of Oxone for the oxidation of
198 J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204

process, the solution pH dropped rapidly within the initial 30 min


of reaction, and thereafter, followed a gradual decrease
(Supporting Information Material Fig. S3).The decrease in pH is
because of the formation of H+, HSO 5 species and the oxidation
intermediates which are acidic in nature. The intermediates which
are formed during the BPA oxidation also contributed to the resid-
ual TOC in the reaction mixture [70,53].
After 60 min of reaction time, 81.9% BPA was removed as
against 24.4% TOC removal. After 180 min, the BPA degradation
was found to be 93.0% with 51.6% TOC removal. After 360 min of
irradiation, the BPA removal was 96.7% with a TOC removal of
72.5%. This showed BPA decrement of only 14.8% with a TOC decre-
ment of 48% in the later 300 min of oxidation. The ratio of TOC
removal to that of BPA improves from about 0.29 after 60 min to
about 0.75 after 360 min of reaction. This means that the interme-
diates get mineralized, albeit slowly.

Fig. 5. Effect of degradation of BPA using Cl, HA and HCO


3 under UV/PMS system.
[BPA]o = 0.22 mM, [PMS]o = 0.66 mM; UV/PMS; Cl or HCO3: 5 mM; Cl or
  
HCO3 : 10 mM; Cl or HCO3 : 20 mM. (Inset: UV/PMS; HA: 5 mg/L; 4. Transformation products
10 mg/L; 20 mg/L).

To the best of the author’s knowledge, the mechanism of BPA


degradation by UV-C activated PMS has not yet been reported.
50 mg/L of BPA was calculated to be 4.85 g/L from the following
The extract obtained from the reaction mixture using DCM as a sol-
equation (Eq. 15) as [68]
vent was subjected to GC/MS analysis. Methanol [70], acetonitrile
[71], ethyl acetate [11,72], diethyl ether and DCM [12,73,74] have
C15 H16 O2 þ 72 HSO5 ! 72 SO4 þ 44 H2 O þ 15 CO2 ð15Þ been used as solvents to extract the oxidation products from AOP
The effect of PMS dosage on the degree of mineralization of BPA reaction mixture. Each solvent has its limitation in extracting some
(50 mg/L) was studied at the PMS dosage of 0.16, 0.66 and 1.62 mM chemicals and intermediates and leaving out some others.
(Inset of Fig. 6). It is seen that at 0.16 mM dosage of PMS, 64% Therefore, it is possible that some chemicals and intermediates
removal of BPA and 58% TOC removal were obtained with a con- may not have been extracted by DCM. The reaction products at
sumption of 83% of PMS in the oxidation process. very low concentration may also have escaped detection by the
With an increase in the PMS dosage from 0.66 mM to 1.62 mM, GC/MS. The GC/MS mass spectra of these intermediates were eval-
the BPA degradation decreased from 96.7% to 90% and the TOC uated on the basis of literature information using the NIST and
removal was also observed to decrease from 72.5% to 67%, respec- Wiley libraries. The FTIR analysis confirmed the formation of inter-
tively, with the PMS consumption being 35% and 37%, respectively. mediates through identification of bonds and functional groups.
Yang et al. [69] also reported 40% PMS consumption for the GC chromatogram at 180 min reaction time revealed seven
degradation of the azo dye (AO7) under 254 nm UV irradiation, stable intermediates (Supporting Information Material Fig. S4).
when applied in the PMS: AO7 molar ratio of 10:1. Fig. 6 shows The number of peaks representing intermediates decreased to five
the TOC removal corresponding to BPA degradation during the as also their magnitudes at 360 min reaction time. Relative abun-
course of oxidation over a period of 360 min. During the oxidation dance of the degradation intermediates was calculated by normal-
izing the peak areas at the defined reaction time to the highest
peak area (Supporting Information Material Fig. S5). The stable
intermediates along with some unstable compounds which were
identified from the rigorous analysis of the data are reported in
Table 1.
Among the identified products, one product having a molecular
weight of 206 is detected for the first time. Some of the identified
compounds have lower molecular weights (acids, 3,5-dimethy-
lethyl phenol, phenol) than BPA, while some have higher molecular
weights (1-isopropyl-1-phenoxy-2-isopropyl-4-phenol, and qui-
none of dihydroxylated BPA) than BPA.
The formation of compounds having similar molecular weight
characteristics are also reported in the literature [11,71].
The UV-spectra of the freshly prepared reaction mixture com-
prising 0.22 mM BPA and 0.66 mM PMS initially were obtained at
different reaction times over a time period of 0–360 min. For the
freshly prepared solution, the absorbance spectrum shows a band
with a maximum wave length (kmax) around 276 nm. This spectral
band is the manifestation of BPA. There is intense absorption
because of PMS. The spectra of the reaction mixture after different
reaction times show the absence of BPA and an increase in the
absorbance over the entire range of the wavelength span from
Fig. 6. Time course of fractional removal of BPA and TOC. [BPA]o = 0.22 mM;
190 nm to 500 nm. This is the characteristic of absorbance and dis-
[PMS]o = 0.66 mM. BPA removal; TOC removal. [Inset: BPA removal, h TOC persion of light due to presence of almost invisible tiny particles
removal, PMS consumption at specific PMS dosage]. (Supporting Information Material Fig. S6) [75]. This illustrates that
J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204 199

Table 1
Possible by-products formed during oxidation of BPA under optimum conditions.

Compound/formula Tentative structure MW/ Detected Remaining


(m/z) intermediates
reported by/in other
AOPs
p
Bisphenol-A 228

p
Formic acid (16.75 min) 116

p
p-Phenol-2-hydroxy carboxylic acid (19.2 min) 168

p
Biphenyldiol (24.5 min) 186

p
Quinone of MHBPA (21.11 min) 242

Benzaldehyde 106 [70]

p-Hydroxyacetophenone 136 [16,17]

Benzophenone 182 [33]

p
Monohydroxylated BPA (MHBPA) 24.1 min 244

p
Phenol,2,4-bis(1,1-dimethyl ethyl) (18.2 min) 206

p
Phenol,3,5-bis(1,1-dimethyl(ethyl) 206

p
2-Hydroxy-4-(2-(3-hydroxy-4-oxocyclohexa-2,5-dienyl)propan-2- 257
yl)cyclohexa-2,5-dienone [or quinone of dihydroxylated BPA]
(26.11 min)

(continued on next page)


200 J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204

Table 1 (continued)

Compound/formula Tentative structure MW/ Detected Remaining


(m/z) intermediates
reported by/in other
AOPs
p
Quinone of dihydroxylated BPA 257

p
4-isopropyl-1-phenoxy-2-isopropyl-4-phenol (29.12 min) 270

Hydroxylated (4-isopropyl-1-phenoxy-2-isopropyl-4-phenol) 286 [12]

Coupling product 320 [12]

the depletion of BPA with reaction time is accompanied by the H-abstraction cannot be ruled out, and the nature, ring position,
increasing amount of a brownish solid product. and H-bond formation capacity of the substituted phenols will
The FTIR analysis of the filtrate confirmed the formation of dictate the H-abstraction efficiency by different radicals [75,83].
various products as the spectra show the presence of benzene The sulfate radicals react with aromatic ring and olefinic double
and aromatic ring stretch (<1000 cm1), along with acidic (2800– bonds, thereby forming sulfate radical adducts having very short
3200 cm1), carbonyl (1660–1651 cm1), and hydroxyl functional life (<200 ns) [82,49]. The sulfate adducts can also be formed due
groups (3300–3400 cm1). The FTIR spectrum of the filtered to electron transfer from aromatic ring to SO 4 by releasing a sul-
brownish solid product (Supporting Information Material inset of fate group (SO24 ) [84,85]. The sulfate radical adducts, thus formed,
Fig. S6) shows peaks at 3408 cm1 (–OH stretching, vibrations), produce hydroxycyclohexadienyl (HCHD) radicals by nucleophilic
2969 cm1 (–OH stretching of acidic groups as the 2400– attack of water [22,86,87]. The elimination of water molecule from
3000 cm1 medium intensity band is described for acidic com- HCHD radical (under acidic conditions) is responsible for the pro-
pounds [76], 1612 cm1 (benzene ring, aromatic groups), duction of highly oxidizing phenoxyl BPA (PhxBPA) radical
1505 cm1 (phenyl ring), 1221 cm1 (C–O–C stretching), 832 (band (Fig. 7) [22,88]. Alternative path of PhxBPA radical formation can
of para substituted aromatic compounds), and 567 cm1 attributed be via H-abstraction from BPA molecule as stated earlier.
to C–H stretching out of the benzene plane. These characteristics Addition of oxygen (from air and/or decomposition of PMS) to
indicate the formation of solid polyphenols. Similar characteristics HCHD radical will form dioxygen radical adducts which are further
were also observed during the degradation of phenol by thermally responsible for the formation of hydroxylated BPA byproducts such
activated PMS [75], and for polyphenol polymers containing phe- as 1,2-benzenediol (catechol) and hydroquinone [22,87] by elimi-
nylene and oxyphenylene units [77]. nation of HO2.
A number of studies reported on the toxicity of the degradation Also, because of the b-scission (C–C) in the PhxBPA radical, the
byproducts [78–80]. It has been concluded that the phenolic moi- phenoxy (PhO) and isopropenylphenol (IPP) radicals are formed
ety is responsible for the estrogenic activity in comparison to qui- (Fig. 7) which is responsible for further degradation of BPA to phe-
none like and aromatic ring opened products [16,81]. Fang et al. nol, hydroquinone, benzoquinone, and simpler organic acids
[81] showed that the substituted group (–OH) on the aromatic [16,89]. The addition/coupling of these radicals to each other or
structure imparted binding affinity to receptors through H-bond with themselves or with neutral BPA molecules generate oligo-
interaction. In the present study, only one hydroxylated BPA inter- meric/high molecular weight compounds [12] (as is also evidenced
mediate (stable) was identified while the others were the quinones by the formation of polyphenols observed through FTIR spec-
and the ring opened structures. These compounds are stated to be troscopy). The detected species with m/z = 320 is due to the
less estrogenic as compared to BPA [16,81,82]. bimolecular combination of PhxBPA followed by the elimination
of IPP radicals. The combination of IPP radicals followed by the
removal of water molecules results in the formation of a byproduct
5. Degradation mechanism and pathway identified at m/z = 270 in this study [12].
The monohydroxylated BPA (m/z = 244) is formed by the reac-
The radical identification study performed in this work demon- tion of isopropenylphenol with hydroquinone or as a result of
strates that the UV-C activation of PMS produces SO
4 radicals pre- HO addition on the aromatic ring of BPA [71]. Monohydroxylated
dominantly. The sulfate radicals react with the substrate via direct BPA serves as a precursor for the formation of its quinone
electron transfer and/or by addition/elimination reaction, and (m/z = 242) and the quinone of dihydroxylated BPA (m/z = 257).
form HO adducts via hydrolysis [42,82]. But the possibility of This may be due to the SO 4 radicals which are predominantly
J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204 201

Fig. 7. Possible formation of radical adducts during BPA oxidation by UV/PMS system.

present in the reaction media and have high reactivity toward The degradation pathway and mechanism are illustrated in
hydroxylated intermediates. The byproduct identified at m/z = 206 Fig. 8. Several researchers [22,87,90] have reported the hydroxycy-
is observed for the first time in the reaction mixture after AOP. clohexadienyl and phenoxy radical pathways in SO 4 based AOPs
This might be formed due to resonance stabilization of PhO radical for phenol. The formation of quinone of dihydroxylated BPA, iso-
and the alkyl chain of the IPP radical. propyl phenol, and phenol proves the involvement of hydroxycy-
The detection of low molecular weight (m/z = 168, m/z = 116) clohexadienyl and phenoxy radicals. Further attack of sulfate and
acidic compounds, such as oxalic, acetic and formic acids, may be hydroxyl radicals may be responsible for the formation of lower
the result of further oxidation and ring opening of the aromatic molecular weight phenols and acids.
compounds.
Similar pathways were also reported for the phenol degradation
by UV/H2O2, UV/PS and UV/PMS AOP [22] for TCA degradation 6. Conclusions
under UV activated persulfate [42] and Fenton like treatment of
BPA [12]. The results obtained in this study indicate that the AOP using
It can, therefore, be concluded that when SO PMS under UV-C irradiation is a feasible treatment method for
4 radical oxidizes
the BPA molecule, one electron is transferred from the BPA mole- the elimination of BPA from aqueous phase. It is found that the per-
cule to produce C-centered BPA radical cation (BPA+). This carbon cent of BPA removal decreases with an increase in the initial BPA
centered radical is also known as HCHD radical. Hydrolysis of the concentration and at the higher PMS concentration than its opti-
radical cation forms OH adducts, (OH) (BPA+). H abstraction gener- mum value. The optimum dosage of PMS was observed to be
ates a hydroxyl BPA radical, ((OH)BPA) which gets degraded to 0.66 mM for 0.22 mM BPA concentration. At the optimum treat-
smaller molecules. The reaction process can proceed as follows: ment conditions, the removal of BPA and TOC was found to be
96.7% and 72.5%, respectively. The degradation of BPA under UV-
þ 2
C/PMS system followed the pseudo-first order reaction kinetics,
SO
4 þ BPA ! BPA þ SO4 ð16Þ under optimum conditions, with the rate constant being
BPAþ þ H2 O ! ðOHÞBPA þ Hþ ð17Þ 0.025 min1 in 60 min of UV-C irradiation.
HO þ BPA ! ðOHÞBPA ð18Þ The BPA degradation increases with an increase in the initial pH
þ of the solution which is ascribed to the absorption coefficient, con-
ðOHÞBPA ! intermediates ! CO2 þ H2 O ð19Þ version of SO 
4 radicals to HO and the speciation of BPA with
202 J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204

CH3
References
OH OH

CH3
[1] M. Furhacker, S. Scharf, H. Weber, Bisphenol A: emission from point sources,
BPA ( m/ z=228) Chemosphere 41 (2000) 751–756.
[2] C.A. Staples, P.B. Dorn, G.M. Klecka, S.T. Oblock, L.R. Harris, A review of the
SO4 - , HO
environmental fate, effects and exposures of Bisphenol A, Chemosphere 36
HCHD (1998) 2149–2173.
-H 2O
[3] T. Yamamoto, A. Yasuhara, H. Shiraishi, O. Nakasugi, Bisphenol A in hazardous
PhenoxyBPA waste landfill leachates, Chemosphere 42 (2001) 415–418.
[4] S.A. Snyder, P. Westerhoff, Y. Yoon, D.L. Sedlak, Pharmaceuticals, personal care
products, and endocrine disruptors in water: implications for the water
IPP and Phenoxy radicals
industry, Environ. Eng. Sci. 20 (2003) 449–469.
[5] H. Fromme, T. Kuchler, T. Otto, K. Pilz, J. Muller, A. Wenzel, Occurrence of
phthalates and Bisphenol-A and F in the environment, Water Res. 36 (2002)
1429–1438.
OH
[6] J.A. Rogers, L. Metz, V.W. Yong, Review: endocrine disrupting chemicals and
OH C
H
COOH immune responses: a focus on Bisphenol-A and its potential mechanisms, Mol.
OH OH Immunol. 53 (2013) 421–430.
( m/ z=168) [7] J.G. Teeguarden, S. Hanson-Drury, A systematic review of Bisphenol A ‘‘low
OH OH
( m/ z=186) dose’’ studies in the context of human exposure: a case for establishing
OR standards for reporting ‘‘low-dose’’ effects of chemicals, Food Chem. Toxicol.
CH 3 62 (2013) 935–948.
OH OH [8] C.D. Metcalfe, T.L. Metcalfe, Y. Kiparissis, B. Koenig, C. Khan, R.J. Hughes, T.R.
( m/ z=206) CH 3 Croley, R.E. March, T. Potter, Estrogenic potency of chemicals detected in
OH
CH 3 ( m/ z=244) sewage treatment plant effluents as determined by in vivo assays with
CH3
Japanese medaka (Oryziaslatipes), Environ. Toxicol. Chem. 20 (2001) 297–
OH C O CH
CH3
CH3 308.
CH3 O O [9] K. Krishnan, M. Gagne, A. Nong, L.L. Aylward, S.M. Hays, Biomonitoring
( m/ z=270) CH3 equivalents for bisphenol A (BPA), Regul. Toxicol. Pharmcol. 58 (2010) 18–24.
OH
( m/ z=242) [10] N.H. Ince, I.G. Apikayan, Synthetic endocrine disruptors in the environment
OH
HO O OH CH3 CH3 and water remediation by advanced oxidation processes, J. Environ. Manage.
O O O O 85 (2007) 816–832.
( m/ z=320) CH3
CH3 [11] H. Katsumata, S. Kawabe, S. Kaneco, T. Suzuki, K. Ohta, Degradation of
OH ( m/ z=257) OH OH
bisphenol A in water by the photo-Fenton reaction, J. Photochem. Photobiol. A:
Chem. 162 (2004) 297–305.
Ring opening [12] J. Poerschmann, U. Trommler, T. Gorecki, Aromatic intermediate formation
during oxidative degradation of Bisphenol A by homogeneous sub-
stoichiometric Fenton reaction, Chemosphere 79 (2010) 975–986.
Oxalic,Formic,Acetic, Hexadionic, Propanoic acid esters
[13] K. Zhang, N. Gao, Y. Deng, T.F. Lin, Y. Ma, M. Sui, Degradation of Bisphenol-A
using ultrasonic irradiation assisted by low concentration hydrogen peroxide,
CO2+ H 2O J. Environ. Sci. 23 (2011) 31–36.
[14] M. Bistan, T. Tisler, A. Pintar, Catalytic and photocatalytic oxidation of aqueous
bisphenol A solutions: removal, toxicity, and estrogenicity, Ind. Eng. Chem.
Fig. 8. Proposed pathway of BPA degradation by UV/PMS system.
Res. 51 (2012) 8826–8834.
[15] S.H. Yoon, S. Jeong, S. Lee, Oxidation of bisphenol A by UV/S2O2 8 : comparison
increasing pH. The studies on the effect of water matrix compo- with UV/H2O2, J. Environ. Technol. 33 (2012) 123–128.
nents indicated no inhibitory effect of Cl up to 5 mM Cl concen- [16] J. Zhang, B. Sun, X. Guan, Oxidative removal of bisphenol A by permanganate:
tration on BPA removal. The addition of 5 mM HCO 3 and 5 mg/L HA
kinetics, pathways and influences of co-existing chemicals, Sep. Purif. Technol.
107 (2013) 48–53.
resulted in a decrease of BPA degradation from 96.7% to 54%, and to [17] J. Lee, H. Park, J. Yoon, Ozonation characteristics of bisphenol A in water,
70%, respectively. Thus, it can be concluded that the presence of Environ. Technol. 24 (2003) 241–248.
chlorides in water (at low concentrations) does not affect the [18] R.A. Torres, C. Petrier, E. Combet, F. Moulet, C. Pulgarin, Bisphenol A
mineralization by integrated ultrasound-UV-iron (II) treatment, Environ. Sci.
BPA degradation. However, HA and bicarbonates, even at low con-
Technol. 41 (2006) 297–302.
centrations, adversely affect the removal of BPA. [19] Y.-F. Huang, Y.-H. Huang, Identification of produced powerful radicals involved
The identification of intermediates shows the presence of in the mineralization of bisphenol A using a novel UV-Na2S2O8/H2O2-Fe(II, III)
hydroxylated products along with quinone(s). The BPA degradation two-stage oxidation process, J. Hazard. Mater. 162 (2009) 1211–1216.
[20] X. Liu, T. Zhang, Y. Zhao, L. Fang, Y. Shao, Degradation of atenolol by UV/
with sulfate radicals is found to proceed via one electron transfer peroxymonosulfate: kinetics, effects of operational parameters and
reaction mechanism. The degradation pathway reveals that mechanism, Chemosphere 93 (2013) 2717–2724.
HCHD and phenoxyl type radicals are probably involved in BPA [21] T.W. Chan, N.J.D. Graham, W. Chu, Degradation of iopromide by combined UV
irradiation and peroxydisulfate, J. Hazard. Mater. 181 (2010) 508–513.
degradation, followed by the formation of sulfate radical adducts. [22] T. Olmez-Hanci, I. Arslan-Alaton, Comparison of sulfate and hydroxyl radical
The possible BPA degradation mechanism involved hydroxylation, based advanced oxidation of phenol, Chem. Eng. J. 24 (2013) 10–16.
oxidative skeletal rearrangement, demethylation, dehydration and [23] H. Hori, A. Yamamoto, E. Hayakawa, S. Taniyasu, N. Yamashita, S. Kutsuna, H.
Kiatagawa, R. Arakawa, Efficient decomposition of environmentally persistent
ring opening. perfluorocarboxylic acids by use of persulfate as a photochemical oxidant,
Environ. Sci. Technol. 39 (2005) 2383–2388.
Acknowledgements [24] G.P. Anipsitakis, T.P. Tufano, D.D. Dionysiou, Chemical and microbial
decontamination of pool water using activated potassium
peroxymonosulfate, Water Res. 42 (2008) 2899–2910.
One of the authors (Sharma J.) gratefully acknowledges the [25] R.H. Waldemer, P.G. Tratnyek, R.L. Johnson, J.T. Nurmi, Oxidation of
financial support provided by the Ministry of Human Resource chlorinated ethenes by heat-activated persulfate: kinetics and products,
Development (MHRD), Govt. of India, for carrying out this work Environ. Sci. Technol. 41 (2007) 1010–1015.
[26] J. Cong, G. Wen, T. Huang, L. Deng, J. Ma, Study on enhanced ozonation
and also thanks Mr. P. Kundu, for his help in the improvement of degradation of para-chlorobenzoic acid by peroxymonosulfate in aqueous
the graphical representation. D. D. Dionysiou also acknowledges solution, Chem. Eng. J. 264 (2015) 399–403.
support from the University of Cincinnati through a UNESCO co- [27] J.A. Khan, X. He, N.S. Shah, H.M. Khan, E. Hapeshi, D. Fatta-Kassinos, D.D.
Dionysiou, Kinetic and mechanism investigation on the photochemical
Chair Professor position on ‘‘Water Access and Sustainability". degradation of atrazine with activated H2O2, S2O2 
8 and HSO5 , Chem. Eng. J.
252 (2014) 393–403.
Appendix A. Supplementary data [28] W.M. Latimer, Oxidation Potentials, Prentice-Hall, Inc., Englewood Cliffs, NJ,
1952.
[29] C. Liang, Z.-S. Wang, N. Mohanty, Influences of carbonate and chloride ions on
Supplementary data associated with this article can be found, in persulfate oxidation of trichloroethylene at 20 °C, Sci. Total Environ. 370
the online version, at http://dx.doi.org/10.1016/j.cej.2015.04.021. (2006) 271–277.
J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204 203

[30] K.-C. Huang, R.A. Couttenye, G.E. Hoag, Kinetics of heat-assisted persulfate [59] Y. Lee, U. von Gunten, Quantitative structure reactivity relationships (QSARs)
oxidation of methyl tert-butyl ether (MTBE), Chemosphere 49 (2002) 413– for the transformation of organic micropollutants during oxidative water
420. treatment, Water Res. 46 (2012) 6177–6195.
[31] G.P. Anipsitakis, D.D. Dionysiou, Radical generation by the interaction of [60] T.K. Lau, W. Chu, N.J.W. Graham, The aqueous degradation of butylated
transition metals with common oxidants, Environ. Sci. Technol. 38 (2004) hydroxyanisole by UV/S2O2 8 : study of reaction mechanisms via dimerization
3705–3712. and mineralization, Environ. Sci. Technol. 41 (2007) 613–619.
[32] G.P. Anipsitakis, E. Stathatos, D.D. Dionysiou, Heterogeneous activation of [61] C. Liang, H.-W. Su, Identification of sulfate and hydroxyl radicals in thermally
oxone using Co3O4, J. Phys. Chem. B. 109 (2005) 13052–13055. activated persulfate, Ind. Eng. Chem. Res. 48 (2009) 5558–5562.
[33] Yi-F. Huang, Yao-H. Huang, Behavioral evidence of the dominant radicals and [62] Z. Wang, R. Yuan, Y. Guo, L. Xu, J. Liu, Effects of chloride ions on bleaching of
intermediates involved in Bisphenol A degradation using an efficient Co2+/PMS azo dyes by Co2+/oxone regent: kinetic analysis, J. Hazard. Mater. 190 (2011)
oxidation process, J. Hazard. Mater. 167 (2009) 418–426. 1083–1087.
[34] X. Chen, X. Qiao, D. Wang, J. Lin, J. Chen, Kinetics of oxidative decolorization [63] G.-D. Fang, D.D. Dionysiou, Y. Wang, S.R. Al-Abed, D.-M. Zhou, Sulfate radical-
and mineralization of Acid Orange 7 by dark and photoassisted Co2+-catalyzed based degradation of polychlorinated biphenyls: effects of chloride ion and
peroxymonosulfate system, Chemosphere 67 (2007) 802–808. reaction kinetics, J. Hazard. Mater. 227–228 (2012) 394–401.
[35] S.K. Ling, S. Wang, Y. Peng, Oxidative degradation of dyes in water using Co2+/ [64] R. Yuan, S.N. Ramjaun, Z. Wang, J. Liu, Effects of chloride ion on degradation of
H2O2 and Co2+/peroxymonosulfate, J. Hazard. Mater. 178 (2010) 385–389. Acid Orange 7 by sulfate radical-based advanced oxidation process:
[36] Q.J. Yang, H. Choi, D.D. Dionysiou, Nanocrystalline cobalt oxide immobilized on implications for formation of chlorinated aromatic compounds, J. Hazard.
titanium dioxide nanoparticles for the heterogeneous activation of Mater. 196 (2011) 173–179.
peroxymonosulfate, Appl. Catal. B: Environ. 74 (2007) 170–178. [65] Y. Chen, K. Zhang, Y.G. Zuo, Direct and indirect photodegradation of estriol in
[37] P.R. Shukla, S. Wang, H. Sun, H.M. Ang, M. Tade, Activated carbon supported the presence of humic acid, nitrate and iron complexes in water solutions, Sci.
cobalt catalysts for advanced oxidation of organic contaminants in aqueous Total Environ. 463 (2013) 802–809.
solution, Appl. Catal. B: Environ. 100 (2010) 529–534. [66] S.K. Atkinson, V.L. Marlatt, L.E. Kimpe, D.R.S. Lean, V.L. Trudeau, J.M. Blais,
[38] P. Shukla, H. Sun, S. Wang, H.M. Ang, M.O. Tade, Nanosized Co3O4/SiO2 for Environmental factors affecting ultraviolet photodegradation rates and
heterogeneous oxidation of phenolic contaminants in waste water, Sep. Purif. estrogenicity of estrone and ethinylestradiol in natural waters, Arch.
Technol. 77 (2011) 230–236. Environ. Contam. Toxicol. 60 (2011) 1–7.
[39] E.A. Betterton, M.R. Hoffmann, Kinetics and mechanism of the oxidation of [67] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of rate
aqueous hydrogen sulfide by peroxymonosulfate, Environ. Sci. Technol. 24 constant for reaction hydrated electrons, hydrogen atoms and hydroxyl
(1990) 1819–1824. radicals (HO/O) in aqueous solution, J. Phys. Chem. Ref. Data 17 (1988)
[40] Y.-H. Guan, J. Ma, X.-C. Li, J.-Y. Fang, L.-W. Chen, Influence of pH on the 513–886.
formation of sulfate and hydroxyl radicals in the UV/peroxymonosulfate [68] W.-D. Oh, S.-K. Lua, Z. Dong, T.-T. Lim, High surface area DPA-hematite for
system, Environ. Sci. Technol. 45 (2011) 9308–9314. efficient detoxification of bisphenol A via peroxymonosulfate activation, J.
[41] Y.R. Wang, W. Chu, Photo-assisted degradation of 2,4,5- Mater. Chem. A 2 (2014) 15836–15845.
trichlorophenoxyacetic acid by Fe(II)-catalyzed activation of Oxone process: [69] S. Yang, P. Wang, X. Yang, L. Shan, W. Zhang, X. Shao, R. Niu, Degradation
the role of UV irradiation, reaction mechanism and mineralization, Appl. Catal. efficiencies of azo dye Acid Orange 7 by the interaction of heat, UV and anions
B: Environ. 123–124 (2012) 151–161. with common oxidants: Persulfate, peroxymonosulfate and hydrogen
[42] X. Gu, S. Lu, Z. Qiu, Q. Sui, Z. Miao, K. Lin, Y. Liu, Q. Luo, Comparison of peroxide, J. Hazard. Mater. 179 (2010) 552–558.
photodegradation performance of 1,1,1-trichloroethane in aqueous solution [70] T. Olmez-Hanci, I. Arslan-Alaton, B. Genc, Bisphenol A treatment by the hot
with the addition of H2O2 or S2O2 8 oxidants, Ind. Eng. Chem. Res. 51 (2012) persulfate process: oxidation products and acute toxicity, J. Hazard. Mater. 263
7196–7204. (2013) 283–290.
[43] L.R. Bennedsen, J. Muff, E.G. Sogaard, Influence of chloride and carbonates on [71] M. Molkenthin, T. Olmez-Hanci, M.R. Jekel, I. Arslan-Alaton, Photo-Fenton-like
the reactivity of activated persulfate, Chemosphere 86 (2012) 1092–1097. treatment of BPA: effect of UV light source and water matrix on toxicity and
[44] M. Sanchez-Polo, M.M. Abdel Daiem, R. Ocampo-Perez, J. Rivera-Utrilla, A.J. transformation products, Water Res. 47 (2013) 5052–5064.
Mota, Comparative study of the photodegradation of bisphenol A by HO, SO 4 [72] K.S. Tay, N. Abd Rahman, Mhd.R.B. Abas, Degradation of bisphenol A by

and CO 3 /HCO3 radicals in aqueous phase, Sci. Total Environ. 463–464 (2013) ozonation: rate constants, influence of inorganic anions, and by-products,
423–431. Maejo Int. J. Sci. Technol. 6 (2012) 77–94.
[45] M. Zhan, X. Yang, Q. Xian, L. Kong, Photosensitized degradation of bisphenol A [73] C. Li, X.Z. Li, N. Graham, N.Y. Gao, The aqueous degradation of bisphenol A and
involving reactive oxygen species in the presence of humic substances, steroid estrogens by ferrate, Water Res. 42 (2008) 109–120.
Chemosphere 63 (2006) 378–386. [74] K. Lin, W. Liu, D. Jaygan, Oxidative removal of bisphenol A by manganese
[46] W. Songlin, Z. Ning, W. Si, Z. Qi, Y. Zhi, Modeling the oxidation kinetics of sono- dioxide: efficacy, products, and pathways, Environ. Sci. Technol. 43 (2009)
activated persulfate’s process on the degradation of humic acid, Ultrason. 3860–3864.
Sonochem. 23 (2015) 128–134. [75] V.C. Mora, J.A. Rosso, D.O. Martire, M.C. Gonzalez, Phenol depletion by
[47] M. Deborde, S. Rabouan, P. Mazellier, J.-P. Duguet, B. Legube, Oxidation of thermally activated peroxydisulfate at 70 °C, Chemosphere 84 (2011) 1270–
bisphenol A by ozone in aqueous solution, Water Res. 42 (2008) 4299–4308. 1275.
[48] R.A. Torres, F. Abdelmalek, E. Combet, C. Petrier, C. Pulgarin, A comparative [76] <http://chemistry.oregonstate.edu/courses/ch361-464/ch362/irinterp.htm> as
study of ultrasonic cavitation and Fenton’s reagent for bisphenol A assessed on 11 June 2014.
degradation in deionised and natural waters, J. Hazard. Mater. 146 (2007) [77] M. Akita, D. Tsutsumi, M. Kobayashi, H. Kise, Structural change and catalytic
546–551. activity of horseradish peroxidase in oxidative polymerization of phenol,
[49] M.G. Antoniou, A.A. de la Cruz, D.D. Dionysiou, Degradation of microcystin-LR Biosci. Biotechnol. Biochem. 65 (2001) 1581–1588.
using sulfate radicals generated through photolysis, thermolysis and e [78] K. Chiang, T.M. Lim, L. Tsen, C.C. Lee, Photocatalytic degradation and
transfer mechanisms, Appl. Catal. B: Environ. 96 (2010) 290–298. mineralization of bisphenol A by TiO2 and platinized TiO2, Appl. Catal. A:
[50] D.H. Lataye, I.M. Mishra, I.D. Mall, Adsorption of 2-picoline onto bagasse fly ash General 261 (2004) 225–237.
from aqueous solution, Chem. Eng. J. 138 (2008) 35–46. [79] P.J. Chen, K.G. Linden, D.E. Hinton, S. Kashiwada, E.J. Rosenfeldt, S.W. Kullman,
[51] A. Gonzalez-Casado, N. Navas, M. del Olmo, J.L. Vilchez, Determination of Biological assessment of bisphenol A degradation in water following direct
Bisphenol A in water by micro liquid–liquid extraction followed by silylation photolysis and UV advanced oxidation, Chemosphere 65 (2006) 1094–1102.
and gas chromatography–mass spectrometry analysis, J. Chromatogr. Sci. 36 [80] Y. Ohko, I. Ando, C. Niwa, T. Tatsuma, T. Yamamura, T. Nakashima, Y. Kubota, A.
(1998) 565–570. Fujishima, Degradation of Bisphenol A in water by TiO2 photocatalyst, Environ.
[52] M. de1 Olmo, A. Gonzalez-Casado, N.A. Navas, J.L. Vilchez, Determination of Sci. Technol. 35 (2001) 2365–2368.
bisphenol A (BPA) in water by gas chromatography–mass spectrometry, Anal. [81] H. Fang, W. Tong, L.M. Shi, R. Blair, R. Perkins, W. Branham, B.S. Hass, Q. Xie, S.L.
Chim. Acta 346 (1997) 87–92. Dial, C.L. Moland, D.M. Sheehan, Structure–activity relationships for a large
[53] A. Ghauch, A.M. Tuqan, Oxidation of bisoprolol in heated persulfate/H2O diverse set of natural, synthetic, and environmental estrogens, Chem. Res.
systems: kinetics and products, Chem. Eng. J. 183 (2012) 162–171. Toxicol. 14 (2001) 280–294.
[54] A. Ghauch, A.M. Tuqan, N. Kibbi, Ibuprofen removal by heated persulfate in [82] R.O.C. Norman, P.M. Storey, P.R. West, Electron spin resonance studies. Part
aqueous solution: a kinetics study, Chem. Eng. J. 197 (2012) 483–492. XXV. Reactions of the sulphate radical anion with organic compounds, J. Chem.
[55] Z. Wang, R.T. Bush, L.A. Sullivan, C. Chen, J. Liu, Selective oxidation of arsenite Soc. B: Phys. Org. (1970) 1087–1095.
by peroxymonosulfate with high utilization efficiency of oxidant, Environ. Sci. [83] G. Merga, B.S.M. Rao, H. Mohan, J.P. Mittal, Reactions of OH and SO 4 with some
Technol. 48 (2014) 3978–3985. halobenzenes and halotoluenes: a radiation chemical study, J. Phys. Chem. 98
[56] P. Maruthamuthu, P. Neta, Radiolytic chain decomposition of (1994) 9158–9164.
peroxomonophosphoric and peroxomonosulfuric acids, J. Phys. Chem. 81 [84] M.G. Antoniou, A.A. de la Cruz, D.D. Dionysiou, Intermediates and reaction
(1977) 937–940. pathways from the degradation of Microcystin-LR with sulfate radicals,
[57] F. Ji, C. Li, L. Deng, Performance of CuO/PMS system: heterogeneous catalytic Environ. Sci. Technol. 44 (2010) 7238–7244.
oxidation of phenol at ambient conditions, Chem. Eng. J. 178 (2011) 239– [85] X. Xie, Y. Zhang, W. Huang, S. Huang, Degradation kinetics and mechanism of
243. aniline by heat-assisted persulfate oxidation, J. Environ. Sci. 24 (2012) 821–826.
[58] N.S. Shah, X. He, H.M. Khan, J.A. Khan, K.E. O’Shea, D.L. Boccelli, D.D. Dionysiou, [86] P. Neta, V. Madhavan, H. Zemel, R.W. Fessenden, Rate constants and
Efficient removal of endosulfan from aqueous solution by UV-C/peroxides: a mechanism of reaction of sulfate radical anion with aromatic compounds, J.
comparative study, J. Hazard. Mater. 263 (2013) 584–592. Am. Chem. Soc. 99 (1977) 163–164.
204 J. Sharma et al. / Chemical Engineering Journal 276 (2015) 193–204

[87] G. Anipsitakis, D.D. Dionysiou, M.A. Gonzalez, Cobalt-mediated activation of [89] Y.-H. Cui, X.-Y. Li, G. Chen, Electrochemical degradation of bisphenol A on
peroxymonosulfate and sulfate radical attack on phenolic compounds. different anodes, Water Res. 43 (2009) 1968–1976.
Implications of chloride ions, Environ. Sci. Technol. 40 (2006) 1000–1007. [90] T. Miyazaki, Y. Katsumura, M. Lin, Y. Muroya, H. Kudo, M. Taguchi, M. Asano,
[88] M. Foti, K.U. Ingold, J. Lusztyk, The surprisingly high reactivity of phenoxyl M. Yoshida, Radiolysis of phenol in aqueous solution at elevated temperatures,
radicals, J. Am. Chem. Soc. 116 (1994) 9440–9447. Radiat. Phys. Chem. 75 (2006) 408–415.

You might also like