You are on page 1of 40

Accepted Manuscript

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


ics, influence of co-existing chemicals and degradation pathway

Jyoti Sharma, I.M. Mishra, Dionysios D. Dionysiou, Vineet Kumar

PII: S1385-8947(15)00503-3
DOI: http://dx.doi.org/10.1016/j.cej.2015.04.021
Reference: CEJ 13510

To appear in: Chemical Engineering Journal

Received Date: 25 December 2014


Revised Date: 1 April 2015
Accepted Date: 2 April 2015

Please cite this article as: J. Sharma, I.M. Mishra, D.D. Dionysiou, V. Kumar, Oxidative removal of Bisphenol A
by UV-C/peroxymonosulfate (PMS): Kinetics, influence of co-existing chemicals and degradation pathway,
Chemical Engineering Journal (2015), doi: http://dx.doi.org/10.1016/j.cej.2015.04.021

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
1 Oxidative removal of Bisphenol A by UV-C/peroxymonosulfate (PMS): Kinetics,
2 influence of co-existing chemicals and degradation pathway
3 Jyoti Sharmaa, I. M. Mishra*a, Dionysios D. Dionysiou*b, Vineet Kumara
4 a: Department of Chemical Engineering, Indian Institute of Technology, Roorkee, Roorkee-
5 247667, India.
6 b: Environmental Engineering and Science Program
7 Department of Biomedical, Chemical and Environmental Engineering (DBCEE)
8 705 Engineering Research Center
9 University of Cincinnati, Cincinnati, OH 45221-0012
10 Abstract

11 In the present study, a sulfate radical-based advanced oxidation process was applied for the
12 degradation of an industrial chemical and suspected endocrine disruptor, Bisphenol A (BPA).
13 UV-C (λ= 254 nm; 40 W power; Io = 1.26 µEs-1) activated peroxymonosulfate (PMS) was
14 used as an oxidant. The effect of operating parameters (initial concentration of BPA, dose of
15 PMS, initial solution pH (pHo), and water matrix components such as chloride (Cl−),
16 bicarbonate (HCO3−) ions and humic acid (HA) was evaluated. At the initial pH of reaction
17 mixture (5.15) and room temperature (29 ± 3˚C),the optimum dosage of PMS was found to
18 be 0.66 mM, giving a BPA removal of 96.7± 0.05 % and a total organic carbon (TOC)
19 removal of 72.5 ± 0.05 % after 360 min of irradiation. With an increase in initial BPA
20 concentration and PMS dosage greater than 0.66 mM, the BPA and TOC removal decreased.
21 The extent of BPA removal increased with an increase in pHo (3 ≤ pHo≤ 12) of the reaction
22 mixture. The degradation of BPA followed pseudo-first-order kinetics and the apparent first
23 order rate constant for BPA was found to be 0.025 min-1 at the optimum oxidation conditions
24 (CBPA=0.22 mM, CPMS = 0.66 mM, pH=5.15, temperature = 29 ± 3˚C).The Cl− ions have
25 negligible inhibition effect on the BPA removal. However, the HCO3− and HA inhibited the
26 BPA oxidation under UV-C irradiation. The identification of intermediates and final products
27 was carried out with HPLC, GC/MS and FTIR, and a degradation pathway was proposed.
28 The present study reveals that the UV-C/PMS oxidation process is effective for BPA removal
29 under real water/wastewater conditions.
30 Keywords: advanced oxidation process (AOP), humic acid, pathway, peroxides,
31 peroxymonosulfate (PMS), persulfate.
32 *Corresponding author E-mail:immishra49@gmail.com, imishfch@iitr.ac.in (I. M. Mishra).
33 * Corresponding author E-mail address: dionysios.d.dionysiou@uc.edu

1
1 1. Introduction
2 A growing number of contaminants are entering water bodies ranging from traditional heavy
3 metals to emerging micropollutants such as natural hormones, pesticides, pharmaceuticals
4 and industrial chemicals. Bisphenol A (BPA) is an industrial chemical which is largely used
5 in the production of various items such as epoxy resins, polycarbonate, etc. It is identified as
6 an endocrine disrupting chemical (EDC), and it finds its way in the water bodies either
7 directly or indirectly via degradation of various BPA containing materials such as plastics,
8 bottles, containers, toys, food/ soft drink cans, packaged food, etc.[1-3]. The presence of
9 EDCs in the water bodies affects the aquatic lives and other life-forms through consumption
10 of contaminated water [4]. The concentration of BPA in water/wastewaters can be as low as
11 0.01 µg/L to as high as 17 mg/L in some industrial effluents and landfill leachates [2, 3, 5, 6].
12 It has affinity to estrogen acceptor and thereby induces feminization in a number of aquatic
13 organisms [7, 8]. The United States Environmental Protection Agency (USEPA) and the
14 European Food Safety Authority (EFSA) set a reference dosage of 50 µg/kg body weight/day
15 of BPA for human consumption [9].
16 Conventional physical-chemical water treatment processes such as adsorption, filtration,
17 chlorination, and coagulation/flocculation as well as biological processes are not able to
18 remove BPA from water satisfactorily, they rather transfer it from one phase to another
19 (adsorption, filtration, coagulation/flocculation) [10]. The limitations of conventional
20 processes to remove BPA have led to efforts to explore alternative methods such as advanced
21 oxidation processes (AOPs). A number of AOP studies described Fenton/photo-Fenton
22 degradation [11, 12], sonochemical degradation [13], photochemical/photocatalytic oxidation
23 [14-16], ozonation [17], and hybrid processes [18, 19] for BPA removal from water.
24 The AOPs are based on highly reactive species such as hydroxyl (HO●) and/or sulfate
25 (SO4●−) radicals. These radicals are capable of degrading the organic molecules and yielding
26 intermediates and CO2. Radical species can be produced mainly from the thermal, UV and
27 transition metal activation of oxidants such as hydrogen peroxide, persulfate (PS) and
28 peroxymonosulfate (PMS).

29 PS and PMS are comparatively newer oxidants in surface water treatment applications, but
30 they have long been used as in-situ chemical oxidation agents in ground water applications.
31 PS and PMS were used by researchers for the removal of pharmaceuticals [20], herbicides
32 [21], phenols [22], perfluorinated compounds [23], disinfection of Escherichia coli [24] and
33 chlorinated compounds [25, 26] in water treatment and groundwater remediation.

2
1 PS and PMS are favored for use in environmental remediation applications because of their
2 easier activation and higher stability than that of hydrogen peroxide under typical
3 water/ground water treatment conditions [27]. The high standard reduction potential (2.5−3.1
4 V) of SO4●‾ radicals and wider operating pH range [28, 29] makes SO4●‾ radicals
5 comparatively more effective and selective for oxidation of target compound in
6 water/wastewater treatment [30-33].

7 PMS is a triple salt of potassium monopersulfate (2KHSO5.KHSO4.K2SO4), a product


8 manufactured by DuPont, and available commercially as OXONE. PMS is considered to be
9 a green oxidant because it is benign, as are most of the byproducts of its reactions [34]. The
10 decomposition rate of organic compounds with PMS at room temperature is low, but it can be
11 enhanced through activation by photolysis or thermolysis (Eq.1) or by using transition metals
12 such as cobalt (Eq. 2), thereby producing highly reactive SO4●− and HO● radicals.
 ● ●
13 HSO5‾ hv/
  SO4 ‾ + HO (1)
14 HSO5‾ + Metal ion (Mn) → Mn+1 + SO4●‾ + OH‾ (2)
15 Cobalt/PMS is reported to have higher rate of degradation than Cobalt/H2O2 system. Cobalt
16 can effectively activate PMS and generated SO4●‾ radicals at enhanced rate, which in turn
17 gives higher rate of degradation than the Cobalt/H2O2 system [35]. Some dissolved metal ions
18 used for activation of oxidants in water/wastewater treatment may lead to several health
19 problems such as asthma, pneumonia, and other lung ailments [32, 36-38]. However, the use
20 of UV irradiation is considered to be environment-friendly. PMS is slightly more powerful
21 than hydrogen peroxide as an oxidant (Eo (HSO5‾/HSO4●‾) = +1.82 V; Eo (H2O2/H2O) = 1.76
22 V) [39]. When irradiated by a low pressure mercury lamp (λ = 254 nm), PMS generates
23 SO4●‾ and HO● radicals through the cleavage of peroxide bond [40, 41], as explained by
24 Eq.(1), and subsequently produces SO42‾, H+ and HO● species through reaction with water
25 (Eq. 3):

26 SO4●‾ + H2O → H+ + SO42‾ + HO● (3)


27 Several studies reported on the effect of water matrix components such as natural organic
28 matter (NOM) and alkalinity (carbonate/bicarbonate) on the degradation of target compounds
29 by AOPs.
30 NOM represents a complex mixture of naturally occurring organic compounds in the surface
31 and ground water. Humic acid (HA) is generally used as a representative NOM [13, 16, 42].
32 The effect of water matrix components on the oxidation process is poorly understood. It is

3
1 assumed traditionally that they have negative impact on the reaction system [30]. However, a
2 few studies showed that the presence of water matrix components enhanced the efficiency of
3 the sulfate radical-based AOPs and reduced the efficiency of HO● radical-based AOPs under
4 real water treatment conditions [43, 44]. The composition of HA is source-dependent and
5 varies greatly. Therefore, its impact on the oxidation of the target compound may sometimes
6 be confounding, depending on its source, the method of oxidation used for the target
7 compound, and the presence and composition of various other water matrix components [45,
8 46].
9 The reaction intermediates and degradation products play a crucial role in the toxicity of
10 treated water. Some studies have reported on the byproducts of AOPs during/at the end of the
11 process [11, 12, 47, 48]. The mechanism of degradation and the formation of intermediates
12 and byproducts have been studied mostly for HO● radical-based processes. Not much has
13 been reported on these aspects for the UV-C activated PMS-based AOP for the treatment of
14 BPA contaminated water.
15 The present study, reports on the degradation of 0.22 mM (50 mg/L) BPA solution by UV-
16 C/PMS AOP under natural reaction mixture pH and room temperature conditions. The
17 optimum dosage of PMS was used for the mineralization and degradation of BPA. The
18 effects of pHo and water matrix components (Cl−, HCO3‾ and HA) on the removal of BPA
19 were also studied. The intermediates and the end-products were analyzed and identified by
20 GC-MS and FTIR spectroscopy, and a reasonable degradation pathway is also proposed.
21 2 Experimental
22 2.1 Reagents and chemicals
23 Bisphenol-A (BPA, CAS 80-05-7), molecular weight = 228 with a purity > 99 % was
24 purchased from Sigma Aldrich and used as received. PMS, anhydrous sodium sulfate, HA
25 (sodium salt), Dichloromethane and NaHCO3 were obtained from Hi Media, Mumbai;
26 acetonitrile (> 99 % purity) from Merck India and NaCl and NaOH were obtained from
27 RFCL, New Delhi. Sulfuric acid (H2SO4) was purchased from Fine Chemicals, New Delhi.
28 Ethanol and tert-butyl alcohol (TBA) were obtained from Merck (India) and Loba Chemie
29 (Mumbai), respectively. A quenching agent, sodium thiosulfate (Na2S2O3.5H2O, 99.5 %)
30 (New India Chemical Enterprises, Kochi) was used to stop the reaction in the sample.
31 Millipore water was used from laboratory (Milli-Q Biocel) system.
32
33
34

4
1 2.2 Batch experiments
2 The present study was conducted using a borosilicate glass photo-reactor (1 L). 450 mL of
3 BPA (0.22 mM or 50 mg/L) aqueous solution was irradiated by using 5 х 8 W low pressure
4 mercury UV-C light (λ = 254 nm). The radiation intensity as measured by ferrioxalate
5 actinometry was 1.26µEs-1. A magnetic stirrer and a cooling fan system were used for proper
6 mixing and cooling during the degradation experiments. Initial pH of the BPA solution and
7 the room temperature at the time of experimentation were 6.2 and 29 ± 3 oC, respectively.
8 The degradation experiments were conducted for 360 min under these conditions. Based on
9 preliminary tests, 1 mL sodium thiosulfate solution (1 M) per mL of sample was used to
10 quench the reaction. During the reaction a solid product was also detected. The quenched
11 sample was filtered through a 0.20 µm syringe filter (Axiva, India). The filtrate was used for
12 HPLC analysis and the filtered solid matter was subjected to FTIR analysis. The experiments
13 were carried out at the natural pH of the reaction mixture except during the studies on the
14 effect of initial pH on degradation. In such experiments, the initial pH of the solution was
15 adjusted using 0.1 N NaOH or 0.1 N H2SO4, as required. Control experiments were carried
16 out in the dark as well as in the absence of PMS. A loss of 9.78 % of BPA was detected
17 under UV-C irradiation in the absence of PMS, while up to 42 % BPA was removed by PMS
18 alone at the room temperature (29 ± 3 C) and natural reaction mixture pH of the system
19 without UV-C, as the PMS gets activated by temperature also [49].
20 2.3 HPLC measurements
21 The BPA concentration in the filtrate was determined by HPLC (Waters India), equipped
22 with a C18 column (3.9 mm x 150 mm) and a UV detector set at 276 nm. The mobile phase
23 was the mixed solvent of acetonitrile (HPLC grade) and water (50/50; v/v) at a flow rate of 1
24 mL/min.
25 2.4 Other measurements
26 A TOC analyzer (TOC-VCPH, Shimadzu) and a pH meter (Cyberscan 510 pH) were used to
27 determine the total organic carbon (TOC) and pH of the treated samples (filtrate),
28 respectively. The samples were quenched by using 1 M sodium thiosulfate aqueous solution
29 before TOC analysis. FTIR spectral measurements of the solid product were carried out using
30 a Thermo Nicolet (model Magna 760) instrument (the details of the method are reported in
31 the Supporting Information Material) for the identification of functional groups [50]. All the
32 experiments were carried out in triplicate and the average values are reported with the error
33 bars representing 95 % confidence of interval.
34

5
1 2.5 GC/MS analysis
2 The analysis of the products was carried out by a GC/MS (Clarus -680 model GC with Elite
3 5-MS, 30 m x 0.25 mm capillary column coupled with Clarus SQ8C model MS; Perkin
4 Elmer, USA). Helium was used as the carrier gas at a flow rate of 1.5 mL/ min. One µL
5 aliquot samples were injected at 220 ˚C in splitless mode. The temperature program for GC
6 followed the sequence: start at 40 ˚C, hold for 5 min; 40-200 ˚C (10 ˚C/min), hold for 5 min;
7 200-230 ˚C (5 ˚C/min), hold for 5 min. The transfer line and ion source temperature were
8 maintained at 200 ˚C and 170 ˚C, respectively. Qualitative analysis was performed in the
9 electron impact (EI) mode at 70 eV using the full scan mode in the m/z range of 40-500 amu.
10 The unknown peaks were identified using the library of Wiley and NIST. Dichloromethane
11 (DCM) was used as the solvent to extract the BPA and the organic products of the oxidation
12 process (the details of the extraction method are reported in Supporting Information Material)
13 [51, 52].
14 2.6 Identification of dominant radicals
15 In order to understand the principle oxidation radicals, HO● or SO4●‾ or both effecting the
16 degradation of BPA, ethanol and tert-butyl alcohol (TBA) were used as scavenging agents for
17 HO● and SO4●‾ radicals [31, 15, 53]. The scavenger, oxidant and BPA were used in the molar
18 ratio of 200/1/1. The radical identification studies were conducted at optimum conditions
19 (CBPA=0.22 mM, CPMS = 0.66 mM, t = 360 min), but with excess of the scavenging agent
20 [15].

21 3 Results and discussion


22 3.1 Effect of PMS dosage
23 From Fig. 1, it is seen that 96.7 % BPA was removed in 360 min reaction time when no
24 scavenger was used. However, in the presence of ethanol, the BPA removal was decreased to
25 35 % only. In the presence of TBA, the BPA removal was 85 %. The scavenging rate
26 constants for ethanol vary from 1.2 to 2.8 x 109 M-1 s-1 for HO● and from 1.6 to 7.7 x 107 M-
27 s for SO4●‾ [31, 53]. For TBA the rate constant is approximately 1000 fold greater for HO●
1 -1

28 (3.8 to 7.6 x 108 M-1s-1) than that for SO4●‾ (4 to 9.1 x 105 M-1s-1) [31, 53].
29 The ratio of scavenging rates of oxidant radicals by ethanol and TBA varies between 1.3 to ~
30 3.5 for HO● radicals and between ~18 to 200 for SO4●‾ radicals. The scavenging by ethanol is
31 about one order of magnitude faster for HO● radicals and about two orders of magnitude
32 faster for SO4●‾ radicals than the scavenging rates of the respective radicals by TBA.
33 Therefore, the BPA degradation is much lower in the presence of ethanol than that for TBA.

6
1 In the case of TBA, SO4●‾ radicals are scavenged at much lower rate than HO● radicals, and
2 thus, the SO4●‾ radicals present in the reaction mixture are responsible for BPA degradation.
3 This behavior can also be explained by calculating the utilization efficiency of PMS after 360
4 min of reaction. The inset of Fig. 1 shows an insignificant utilization efficiency of PMS (~9
5 %) due to poor degradation of BPA (~35 %), and a high consumption of PMS because of
6 high reaction rates of ethanol with SO4●− and HO● radicals. In the presence of TBA, the PMS
7 utilization efficiency increases to 55 %, while the PMS utilization efficiency increases to ~ 92
8 % in the absence of alcohol (scavenging agent).
9 (Fig. 1)
10 The higher removal of BPA in the presence of TBA in comparison to ethanol explains the
11 selectivity of SO4●− radicals for BPA degradation. The maximum efficiency in alcohol-free
12 system explains the dominance and selectivity of SO4●− radicals [54]. Wang et al. [55] also
13 reported > 90 % utilization efficiency of PMS for As(III) removal from groundwater at
14 As(III)/PMS molar ratio of 1:2.
15 Fig. 2 shows the effect of PMS concentration (0.16 < CPMS < 1.62 mM) on the degradation of
16 BPA. The experiments at room temperature (29±3 C) showed 42 % removal of BPA without
17 UV-C irradiation. This means that the PMS was thermally activated [49], and therefore, it
18 also partly contributed to the removal of BPA. The removal of BPA increased from ~ 64.2 %
19 to 96.7 % when the CPMS was increased from 0.16 mM to 0.66 mM. At CPMS > 0.66 mM, the
20 incremental removal of BPA was not significant. This is because of the scavenging of SO4●‾
21 and HO● radicals by HSO5‾, and the formation of less reactive SO5●‾ radicals [56]:

22 HSO5‾ + HO● → SO5●‾ + OH‾ (4)


23 HSO5‾+ SO4●‾→ SO42‾ + H+ (5)
24 Thus, the optimum CPMS was found to be 0.66 mM. Wang and Chu [41] also observed a
25 similar trend during the degradation of 2,4,5-trichlorophenoxy acetic acid (0.1 mM) by UV
26 activated PMS in the presence of iron. Ji et al. [57] also found that the CuO activated PMS is
27 effective at 1.5 mM PMS with 100 % phenol (50 mg/L) removal and 86 % TOC removal in
28 60 min. The phenol removal efficiency decreased when PMS dosage was 1.25 mM.
29 (Fig. 2)
30 The degradation of BPA by UV-C/PMS system can be explained by pseudo-first order
31 reaction kinetics (ln (C/C0) = -kt) [19, 44]. The nonlinear regression analysis shows a good fit
32 to the experimental data during the initial 60 min of reaction time (inset of Fig. 2).The
33 apparent first order rate constant for the removal of BPA at its initial concentration of 0.22

7
1 mM, varied from 0.003 min-1 to 0.025 min-1 with an increase in the PMS dosage from an
2 initial 0.16 mM to an optimum dosage of 0.66 mM, but decreased to 0.011 min -1 at 1.62 mM
3 PMS dosage. Based on the above results, 0.66 mM PMS concentration was used in further
4 experimental study.
5
6 3.2 Effect of BPA concentration
7 Fig. 3 shows the variation of residual fractional BPA concentration with the time of oxidation
8 for four initial BPA concentration values of 0.04 mM (10 mg/L), 0.13 mM (30 mg/L), 0.22
9 mM (50 mg/L) and 0.31 mM (70 mg/L). As the initial concentration increased, the residual
10 BPA concentration also increased: 0.1, 0.6, 1.65 and 21 mg/L for an initial BPA
11 concentration of 10, 30, 50 and 70 mg/L, respectively. The retardation in BPA removal may
12 be attributed to the requirement of a larger number of SO4●‾ radicals by the BPA molecules
13 and the intermediates present in the solution and the competition between BPA molecules
14 and the intermediates for free radicals. These results are consistent with those reported in
15 literature [15]. The initial degradation rates of the photooxidation process were estimated
16 from the slope of the plot of BPA concentration versus time at t = 0 for the initial 60 min of
17 reaction time (Inset of Fig. 3). It can be seen that the initial degradation rate of BPA increased
18 from 2.2 x 10-4 mmol L-1 min-1 to 10.0 x 10-4 mmol L-1 min-1 with an increase in the initial
19 BPA concentration from 0.04 mM to 0.31 mM. This is because of the availability of the
20 larger number of BPA molecules to oxidizing free radicals and thereby enhancing the rate of
21 reaction [58]. In accordance with these observations, the degradation time for > 90 % BPA
22 removal is also found to have decreased for the initial BPA concentration < 0.13 mM.
23 (Fig. 3)
24 3.3 Effect of initial solution pH on BPA degradation
25 The pH of the reaction mixture has profound effect on the total radical concentration and
26 their formation in SO4●‾ based AOPs [40]. In order to investigate the effect of initial pH of
27 reaction mixture on BPA degradation by PMS under UV-C irradiation, experiments were
28 performed at five initial pH values of BPA reaction mixture (pHo: 3.0, 5.0, 7.0, 10.0 and
29 12.0).
30 It is seen that the BPA removal increased with an increase in the initial pH of the reaction
31 mixture (Fig. 4): 74.3 % (pHo= 3), 88.3 % (pHo= 5), 97.1 % (pHo= 7), 97.9 % (pHo= 10), and
32 98.8 % (pHo= 12) as against 96.7 % BPA removal at the natural reaction mixture pH. The
33 time course of BPA removal at different pH values follows the order: pH o3 <pHo5 <pHN ≈
34 pHo7 <pHo10 <pHo12.

8
1 The variation in removal with pHo can be explained on the basis of three factors (a)
2 absorption coefficient, (b) conversion of SO4●‾radicals to HO● radicals with pH, and (c) the
3 rate of reaction of radicals with deprotonated BPA.
4 As the pH of the reaction system increases, the absorption coefficient of PMS (єPMS)
5 increases (єPMS = 13.8 - 149.5 M-1cm-1; pH 6-12), with the enhanced photolysis of PMS,
6 which, in turn, resulted in the increased concentration of SO4●‾and HO● radicals [40]. The
7 BPA molecule has two dissociation constants: pKa1 = 9.6 and pKa2 = 10.2 [2]. BPA can
8 dissociate as Bis(OH)(O)‾or Bis(O)2‾in the reaction mixture. As indicated by the BPA
9 speciation diagram (Supporting Information Material Fig. S1), the first deprotonation of BPA
10 starts at about pH 7 and the second one at about pH 9. At pH < pKa1, undissociated BPA (i.e.
11 Bis(OH)2) is predominantly present in the solution. In the pH range, pKa1 < pH < pKa2, the
12 predominant species is Bis(OH)(O)‾, and at pH > pKa2, the majority species is the Bis(O)2‾.
13 The phenolate ions are reported to have higher reaction rate constant in comparison to neutral
14 molecules of BPA [59]. It may, therefore, be suggested that all the above three factors act in a
15 synergistic manner to effect the removal of BPA from the solution.
16 (Fig. 4)
17 It has also been reported, that the degradation of butylated hydroxyl anisole increases as the
18 pH increases from 3 to11 in a PS/UV system [60], and the degradation of nitrobenzene
19 improves as the pH increases from 7 to 12 in a thermally activated PS system [61] due to
20 conversion of SO4●‾ to HO● at alkaline pH [40]. During the course of oxidation, the pH of the
21 reaction mixture decreases from its initial value with its value dropping down to ~ 3 for 5 <
22 pHo< 10 and to ~ 9 for pH 12. This is because of the acidic byproducts obtained at the end of
23 the oxidation (as explained in section 4).
24 It can be seen that the BPA removal increases as pHo increases from the acidic region to the
25 alkaline region, although the difference is not appreciable. On the basis of above results,
26 further study was carried out at the natural solution pH (pHN), because of the marginal
27 increment in BPA removal with an increase in pHo from the natural pH of the solution to the
28 alkaline region.
29 3.4 Effect of co-existing water matrix chemicals
30 The components of the water matrix in ionized form also affect the fate of the target
31 compound. The effect of water matrix components such as Cl‾ and HCO3‾ ions and HA on
32 BPA removal by UV-C/PMS oxidation was studied. The experiments were conducted in the
33 concentration range of 5-20 mM for Cl‾ and HCO3‾ ions, and in the concentration range of 5-
34 20 mg/L for HA. The results indicated that the BPA degradation was not significantly

9
1 influenced at 5 mM Cl‾ ion concentration. However, as the concentration of Cl− ions was
2 increased to 10 and 20 mM, an inhibition in degradation of BPA was observed (Supporting
3 Information Material Fig. S2 (a)). It is thermodynamically feasible for PMS (1.85 V) to
4 oxidize chloride ions (Cl‾) into less reactive chlorine species viz. Cl2/2Cl‾ (1.36 V) and
5 Cl/HOCl (1.48 V) [24, 62]. Therefore, the observed higher degree of inhibition to BPA
6 degradation at 10-20 mM Cl− concentration than that at 5 mM is due to the consumption of
7 sulfate radicals by Cl− ions [29], and the formation of less reactive chlorine species Cl2,
8 HOCl, Cl• and Cl2•− [24, 62, 63]. The chemical reactions involved in (but not limited to) UV-
9 C/PMS/BPA system in the presence of Cl‾ ions can be given as follows [24, 64]:
10 HSO5‾ + Cl‾ →SO42‾ + HOCl (6)
11 HSO5‾ + 2 Cl‾ + H+ → SO42‾ + Cl2‾ + H2O (7)
12 SO4•− + Cl‾ → SO42‾ + Cl• + H2O (8)
13 Cl• + Cl‾ ↔ Cl2•‾ (9)
14 Cl2•‾ + Cl2•‾→ Cl2 + 2 Cl‾ (10)
+
15 Cl2 + H2O → HOCl + H + Cl‾ (11)
+
16 HOCl → H + OCl‾ (pKa=7.5) (12)
17 Similar trends have also been observed by Yuan et al. [64] for sulfate radical-based
18 (Co/PMS) AOP of Acid Orange 7.
19
20 HA is shown to act as a photosensitizer, light filter and quencher of free radicals. The overall
21 effect of HA on the photodegradation of BPA will, therefore, depend on the competition
22 between the BPA and the HA molecules [45]. It has also been reported that the differences in
23 the quality and the type of the HA (or NOM) would give different results on the degradation
24 of organics [65, 66]. In the present study, the HA is found to adversely affect the oxidation of
25 BPA (Supporting Information Material Fig. S2 (b)). The degradation of BPA in the presence
26 of 10 mg/L HA was 54.0 % as compared to 96.7 % without HA. An increase in the HA
27 concentration further increases the degree of inhibition. Similar trends were also observed by
28 Zhang et al. [13]. At high concentration, the HA may be in competition with BPA for SO4●‾
29 and HO● radicals, thereby reducing the availability of oxidizing radicals for BPA degradation.
30 This slows down the BPA degradation rate. The increased concentration of HA also lowers
31 the solution transparency, thereby inhibiting the UV-C irradiation to pass through the reaction
32 mixture.
33 The bicarbonate ion (HCO3‾) is an efficient hydroxyl and sulfate radical scavenger [53, 67].
34 Therefore, the presence of bicarbonate ions would have an adverse effect on BPA
35 degradation. The results of the present study also support the above inference (Supporting

10
1 Information Material Fig. S2 (c)). The redox potential of bicarbonate ion is 1.78 V and its
2 capability for scavenging SO4●‾ and HO● radicals, are given by equations (13) and (14) [53]:
3 SO4●‾ + HCO3‾ → SO42‾ + CO3●‾ + H+ (k11 = 1.6 x 106 M-1s-1) (13)

4 HO + HCO3‾ → CO3●‾ + H2O 6
(k12 = 8.5 x 10 M s )-1 -1
(14)
5 So the scavenging or reactive radical species (SO4●‾ or HO●) is responsible for the BPA
6 removal inhibition. A comparative chart for the effect of water matrix components on the
7 BPA removal by UV-C/PMS system after 360 min of irradiation is illustrated in Fig. 5. In
8 comparison to 96.7 % removal of BPA by UV-C/PMS alone, the removal of BPA increases
9 to 99.0 % in the presence of 5 mM Cl- ion, but decreases to 54 % and 70 % in the presence of
10 5 mM HCO3−ion and 5 mg/L HA (inset of Fig. 5), respectively. This means that the BPA
11 degradation is not affected much in the chloride bearing water, whereas the presence of
12 HCO3− and HA have measurable adverse effect on BPA degradation by PMS under UV-C
13 irradiation.
14 (Fig. 5)
15
16 3.5 Degree of mineralization and PMS consumption
17 TOC is used as a measure of mineralization of target organic compounds. The requirement of
18 Oxone for the oxidation of 50 mg/L of BPA was calculated to be 4.85 g/L from the following
19 equation (Eq. 15) as [68].
20 C15H16O2 + 72 HSO5− → 72 SO4− + 44 H2O + 15 CO2 (15)
21 The effect of PMS dosage on the degree of mineralization of BPA (50 mg/L) was studied at
22 the PMS dosage of 0.16, 0.66 and 1.62 mM (Inset of Fig. 6). It is seen that at 0.16 mM
23 dosage of PMS, 64 % removal of BPA and 58 % TOC removal were obtained with a
24 consumption of 83 % of PMS in the oxidation process.
25 With an increase in the PMS dosage from 0.66 mM to 1.62 mM, the BPA degradation
26 decreased from 96.7 % to 90 % and the TOC removal was also observed to decrease from
27 72.5 % to 67 %, respectively, with the PMS consumption being 35 % and 37 %, respectively.
28 Yang et al. [69] also reported ~ 40 % PMS consumption for the degradation of the azo dye
29 (AO7) under 254 nm UV irradiation, when applied in the PMS: AO7 molar ratio of 10:1.
30 Fig.6 shows the TOC removal corresponding to BPA degradation during the course of
31 oxidation over a period of 360 min. During the oxidation process, the solution pH dropped
32 rapidly within the initial 30 min of reaction, and thereafter, followed a gradual decrease
33 (Supporting Information Material Fig. S3).The decrease in pH is because of the formation of
34 H+, HSO5‾ species and the oxidation intermediates which are acidic in nature. The

11
1 intermediates which are formed during the BPA oxidation also contributed to the residual
2 TOC in the reaction mixture [70, 53].
3 After 60 min of reaction time, 81.9 % BPA was removed as against 24.4 % TOC removal.
4 After 180 min, the BPA degradation was found to be 93.0 % with 51.6 % TOC removal.
5 After 360 min of irradiation, the BPA removal was 96.7 % with a TOC removal of 72.5 %.
6 This showed BPA decrement of only 14.8 % with a TOC decrement of 48 % in the later 300
7 min of oxidation. The ratio of TOC removal to that of BPA improves from about 0.29 after
8 60 min to about 0.75 after 360 min of reaction. This means that the intermediates get
9 mineralized, albeit slowly.
10 (Fig.6)
11 4 Transformation products
12 To the best of the authors’ knowledge, the mechanism of BPA degradation by UV-C
13 activated PMS has not yet been reported. The extract obtained from the reaction mixture
14 using DCM as a solvent was subjected to GC/MS analysis. Methanol [70], acetonitrile [71]
15 ethyl acetate [11, 72], diethyl ether and DCM [12, 73, 74] have been used as solvents to
16 extract the oxidation products from AOP reaction mixture. Each solvent has its limitation in
17 extracting some chemicals and intermediates and leaving out some others. Therefore, it is
18 possible that some chemicals and intermediates may not have been extracted by DCM. The
19 reaction products at very low concentration may also have escaped detection by the GC/MS.
20 The GC/MS mass spectra of these intermediates were evaluated on the basis of literature
21 information using the NIST and Wiley libraries. The FTIR analysis confirmed the formation
22 of intermediates through identification of bonds and functional groups.
23 GC chromatogram at 180 min reaction time revealed seven stable intermediates (Supporting
24 Information Material Fig. S4). The number of peaks representing intermediates decreased to
25 five as also their magnitudes at 360 min reaction time. Relative abundance of the degradation
26 intermediates was calculated by normalizing the peak areas at the defined reaction time to the
27 highest peak area (Supporting Information Material Fig. S5). The stable intermediates along
28 with some unstable compounds which were identified from the rigorous analysis of the data
29 are reported in Table 1.
30 Among the identified products, one product having a molecular weight of 206 is identified
31 for the first time. Some of the identified compounds have lower molecular weights (acids,
32 3,5-dimethylethyl phenol, phenol) than BPA, while some have higher molecular weights (1-
33 isopropyl-1-phenoxy-2-isopropyl-4- phenol, and quinone of dihydroxylated BPA) than BPA.

12
1 The formation of compounds having similar molecular weight characteristics are also
2 reported in the literature [11, 71].
3 The UV-spectra of the freshly prepared reaction mixture comprising of 0.22 mM BPA and
4 0.66 mM PMS initially were obtained at different reaction times over a time period of 0-360
5 min. For the freshly prepared solution, the absorbance spectrum shows a band with a
6 maximum wave length (λmax) around 276 nm. This spectral band is the manifestation of
7 BPA. There is intense absorption because of PMS. The spectra of the reaction mixture after
8 different reaction times show the absence of BPA and an increase in the absorbance over the
9 entire range of the wavelength span from 190 nm to 500 nm. This is the characteristic of
10 absorbance and dispersion of light due to presence of almost invisible tiny particles
11 (Supporting Information Material Fig. S6) [75]. This illustrates that the depletion of BPA
12 with reaction time is accompanied by the increasing amount of a brownish solid product.
13 The FTIR analysis of the filtrate confirmed the formation of various products as the spectra
14 shows the presence of benzene and aromatic ring stretch (< 1000 cm-1), along with acidic
15 (2800-3200 cm-1), carbonyl (1660-1651 cm-1), and hydroxyl functional groups (3300-3400
16 cm-1). The FTIR spectrum of the filtered brownish solid product (Supporting Information
17 Material inset of Fig. S6) shows peaks at 3408 cm-1 (-OH stretching, vibrations), 2969 cm-1
18 (-OH stretching of acidic groups as the 2400-3000 cm-1 medium intensity band is described
19 for acidic compounds [76], 1612 cm-1 (benzene ring, aromatic groups), 1505 cm-1 (phenyl
20 ring), 1221 cm-1 (C-O-C stretching), 832 (band of para substituted aromatic compounds), and
21 567 cm-1 attributed to C-H stretching out of the benzene plane. These characteristics indicate
22 the formation of solid polyphenols. Similar characteristics were also observed during the
23 degradation of phenol by thermally activated PMS [75], and for polyphenol polymers
24 containing phenylene and oxyphenylene units [77].
25 A number of studies reported on the toxicity of the degradation byproducts [78-80]. It has
26 been concluded that the phenolic moiety is responsible for the estrogenic activity in
27 comparison to quinone like and aromatic ring opened products [16, 81]. Fang et al. [81]
28 showed that the substituted group (–OH) on the aromatic structure imparted binding affinity
29 to receptors through H-bond interaction. In the present study, only one hydroxylated BPA
30 intermediate (stable) was identified while the others were the quinones and the ring opened
31 structures. These compounds are stated to be less estrogenic as compared to BPA [16, 81,
32 82].
33
34

13
1 5. Degradation mechanism and pathway
2 The radical identification study performed in this work demonstrates that the UV-C activation
3 of PMS produces SO4●‾ radicals predominantly. The sulfate radicals react with the substrate
4 via direct electron transfer and/or by addition/elimination reaction, and form HO● adducts via
5 hydrolysis [42, 82]. But the possibility of H-abstraction cannot be ruled out, and the nature,
6 ring position, and H-bond formation capacity of the substituted phenols will dictate the H-
7 abstraction efficiency by different radicals [75, 83].
8 The sulfate radicals react with aromatic ring and olefinic double bonds, thereby forming
9 sulfate radical adducts having very short life (< 200 ns) [82, 49]. The sulfate adducts can also
10 be formed due to electron transfer from aromatic ring to SO4●‾ by releasing a sulfate group
11 (SO42‾) [84, 85]. The sulfate radical adducts, thus formed, produce hydroxycyclohexadienyl
12 (HCHD) radicals by nucleophilic attack of water [22, 86, 87]. The elimination of water
13 molecule from HCHD radical (under acidic conditions) is responsible for the production of
14 highly oxidizing phenoxyl BPA (PhxBPA) radical (Fig. 7) [22, 88]. Alternative path of
15 PhxBPA radical formation can be via H-abstraction from BPA molecule as stated earlier.
16 Addition of oxygen (from air and/or decomposition of PMS) to HCHD radical will form
17 dioxygen radical adducts which are further responsible for the formation of hydroxylated
18 BPA byproducts such as 1,2- benzenediol (catechol) and hydroquinone [22, 87] by
19 elimination of HO2•.
20 (Fig.7)
21 Also the β-scission (C-C) in the PhxBPA radical, the phenoxy (PhO) and isopropynyl phenol
22 (IPP) radicals are formed (Fig. 7) which is responsible for further degradation of BPA to
23 phenol, hydroquinone, benzoquinone, and simpler organic acids [16, 89]. The addition/
24 coupling of these radicals to each other or with themselves or with neutral BPA molecules
25 generate oligomeric/high molecular weight compounds [12] (as is also evidenced by the
26 formation of polyphenols observed through FTIR spectroscopy). The detected species with
27 m/z = 320 is due to the bimolecular combination of PhxBPA followed by the elimination of
28 IPP radicals. The combination of IPP radicals followed by the removal of water molecules
29 results in the formation of a byproduct identified at m/z = 270 in this study [12].
30 The monohydroxylated BPA (m/z = 244) is formed by the reaction of isopropenylphenol with
31 hydroquinone or as a result of HO● addition on the aromatic ring of BPA [71].
32 Monohydroxylated BPA serves as a precursor for the formation of its quinone (m/z = 242)
33 and the quinone of dihydroxylated BPA (m/z = 257). This may be due to the SO4●‾ radicals
34 which are predominantly present in the reaction media and have high reactivity towards

14
1 hydroxylated intermediates. The byproduct identified at m/z = 206 is observed for the first
2 time in the reaction mixture after AOP. This might be formed due to resonance stabilization
3 of PhO radical and the alkyl chain of the IPP radical.
4 The detection of low molecular weight (m/z = 168, m/z = 116) acidic compounds, such as
5 oxalic, acetic and formic acids, may be the result of further oxidation and ring opening of the
6 aromatic compounds.
7 Similar pathways were also reported for the phenol degradation by UV/H2O2, UV/PS and
8 UV/PMS AOP [22] for TCA degradation under UV activated persulfate [42] and Fenton like
9 treatment of BPA [12].
10 It can, therefore, be concluded that when SO4●− radical oxidizes the BPA molecule, one
11 electron is transferred from the BPA molecule to produce C-centered BPA radical cation
12 (BPA●+). This carbon centered radical is also known as HCHD radicals. Hydrolysis of the
13 radical cation forms OH adducts, (OH) (BPA●+). H abstraction generates a hydroxyl BPA
14 radical, ((OH)BPA●) which gets degraded to smaller molecules. The reaction process can
15 proceed as follows:
16 SO4●− + BPA → BPA●+ + SO42− (16)
17 BPA●+ + H2O → (OH) BPA● + H+ (17)
18 HO● + BPA→ (OH)BPA● (18)
19 (OH)BPA●+ → intermediates → CO2 + H2O (19)
20 The degradation pathway and mechanism are illustrated in Fig. 8. Several researchers [22, 87,
21 90] have reported the hydroxycyclohexadienyl and phenoxy radical pathways in SO4●− based
22 AOPs for phenol. The formation of quinone of dihydroxylated BPA, isopropyl phenol, and
23 phenol proves the involvement of hydroxycyclohexadienyl and phenoxy radicals. Further
24 attack of sulfate and hydroxyl radicals may be responsible for the formation of lower
25 molecular weight phenols and acids.
26 (Fig.8)
27 6. Conclusions
28 The results obtained in this study indicate that the AOP using PMS under UV-C irradiation is
29 a feasible treatment method for the elimination of BPA from aqueous phase. It is found that
30 the percent of BPA removal decreases with an increase in the initial BPA concentration and
31 at the higher PMS concentration than its optimum value. The optimum dosage of PMS was
32 observed to be 0.66 mM for 0.22 mM BPA concentration. At the optimum treatment
33 conditions, the removal of BPA and TOC was found to be 96.7 % and 72.5 %, respectively.
34 The degradation of BPA under UV-C/PMS system followed the pseudo-first order reaction

15
1 kinetics, under optimum conditions, with the rate constant being 0.025 min-1 in 60 min of
2 UV-C irradiation.
3 The BPA degradation increases with an increase in the initial pH of the solution which is
4 ascribed to the absorption coefficient, conversion of SO4•− radicals to HO• and the speciation
5 of BPA with increasing pH. The studies on the effect of water matrix components indicated
6 no inhibitory effect of Cl− up to 5 mM Cl− concentration on BPA removal. The addition of 5
7 mM HCO3− and 5 mg/L HA resulted in a decrease of BPA degradation from 96.7 % to 54 %,
8 and to 70 %, respectively. Thus, it can be concluded that the presence of chlorides in water
9 (at low concentrations) does not affect the BPA degradation. However, HA and bicarbonates,
10 even at low concentrations, adversely affect the removal of BPA.
11 From the identification of intermediates, the hydroxylated products along with quinone(s) are
12 found. The BPA degradation with sulfate radicals is found to proceed via one electron
13 transfer reaction mechanism. The degradation pathway reveals that HCHD and phenoxyl type
14 radicals are probably involved in BPA degradation, followed by the formation of sulfate
15 radical adducts. The possible BPA degradation mechanism involved hydroxylation, oxidative
16 skeletal rearrangement, demethylation, dehydration and ring opening.
17
18 Acknowledgements

19 One of the authors (Sharma J.) gratefully acknowledges the financial support provided by the
20 Ministry of Human Resource Development (MHRD), Govt. of India, for carrying out this
21 work and also thanks Mr. P. Kundu, for his help in the improvement of the graphical
22 representation.

23

24

25

26

27

28

29

16
1 References

2 [1] M. Furhacker, S. Scharf, H. Weber, Bisphenol A: emission from point sources,


3 Chemosphere 41 (2000) 751-756.

4 [2] C.A. Staples, P.B. Dorn, G.M. Klecka, S.T. Oblock, L.R. Harris, A Review of the
5 environmental fate, effects and exposures of Bisphenol A, Chemosphere 36 (1998) 2149-
6 2173.
7
8 [3] T. Yamamoto, A. Yasuhara, H. Shiraishi, O. Nakasugi, Bisphenol A in hazardous waste
9 landfill leachates, Chemosphere 42 (2001) 415-418.
10
11 [4] S.A. Snyder, P. Westerhoff, Y. Yoon, D.L. Sedlak, Pharmaceuticals, personal care
12 products, and endocrine disruptors in water: Implications for the water industry, Environ.
13 Eng. Sci. 20 (2003 ) 449-469.
14
15 [5] H. Fromme, T. Kuchler, T. Otto, K. Pilz, J. Muller, A. Wenzel, Occurrence of phthalates
16 and Bisphenol-A and F in the environment, Water Res. 36 (2002) 1429-1438.
17
18 [6] J.A. Rogers, L. Metz, V.W. Yong, Review: Endocrine disrupting chemicals and immune
19 responses: A focus on bisphenol-A and its potential mechanisms, Mol. Immunol. 53 (2013)
20 421-430.
21
22 [7] J. G. Teeguarden, S. Hanson-Drury, A systematic review of Bisphenol A "low dose"
23 studies in the context of human exposure: a case for establishing standards for reporting
24 "low-dose" effects of chemicals, Food Chem. Toxicol. 62 (2013) 935-948.
25
26 [8] C.D. Metcalfe, T.L. Metcalfe, Y. Kiparissis, B. Koenig, C. Khan, R.J. Hughes, T.R.
27 Croley, R.E. March, T. Potter, Estrogenic potency of chemicals detected in sewage treatment
28 plant effluents as determined by in vivo assays with Japanese medaka (Oryziaslatipes),
29 Environ. Toxicol. Chem. 20 (2001) 297-308.
30
31 [9] K. Krishnan, M. Gagne, A. Nong, L.L. Aylward, S.M. Hays, Biomonitoring equivalents
32 for bisphenol A (BPA), Regul. Toxicol. Pharm. 58 (2010) 18-24.
33

17
1 [10] N.H. Ince, I.G. Apikayan, Synthetic endocrine disruptors in the environment and water
2 remedieation by advanced oxidation processes, J. Env. Manag. 85 (2007) 816-32.
3
4 [11] H. Katsumata, S. Kawabe, S. Kaneco, T. Suzuki, K. Ohta, Degradation of bisphenol A in
5 water by the photo-Fenton reaction, J. Photochem. Photobio. A: Chem. 162 (2004) 297-305.
6
7 [12] J. Poerschmann, U. Trommler, T. Gorecki, Aromatic intermediate formation during
8 oxidative degradation of Bisphenol A by homogeneous sub-stoichiometric Fenton reaction,
9 Chemosphere 79 (2010) 975-986.
10
11 [13] K. Zhang, N. Gao, Y. Deng, T.F. Lin, Y.S. Ma, M. , Degradation of Bisphenol -A using
12 ultrasonic irradiation assisted by low concentration hydrogen peroxide, J. Environ. Sci. 23
13 (2011) 31-36.
14
15 [14] M. Bistan, T. Tisler, A. Pintar, Catalytic and Photocatalytic Oxidation of Aqueous
16 Bisphenol A Solutions: Removal, Toxicity, and Estrogenicity, Ind. Eng. Chem. Res. 51
17 (2012) 8826-8834.
18
19 [15] S.H. Yoon, S. Jeong, S. Lee, Oxidation of bisphenol A by UV/S2O82- : Comparison with
20 UV/H2O2, J. Env. Tech. 33 (2012) 123-128.
21
22 [16] J. Zhang, B. Sun, X. Guan, Oxidative removal of bisphenol A by permanganate:
23 Kinetics, pathways and influences of co-existing chemicals, Sep. Purif. Technol. 107 (2013)
24 48-53.
25
26 [17] J. Lee, H. Park, J. Yoon, Ozonation characteristics of bisphenol a in water, Environ.
27 Technol. 24 (2003) 241-248.
28
29 [18] R.A. Torres, C. Petrier, E. Combet, F. Moulet, C. Pulgarin, Bisphenol A Mineralization
30 by Integrated Ultrasound-UV-Iron (II) Treatment, Environ. Sci. Technol. 41 (2006) 297-302.
31
32 [19] Y.-F. Huang, Y.-H. Huang, Identification of produced powerful radicals involved in the
33 mineralization of bisphenol A using a novel UV-Na2S2O8/H2O2-Fe(II,III) two-stage oxidation
34 process, J. Hazard. Mater. 162 (2009) 1211-1216.

18
1
2 [20] X. Liu, T. Zhang, Y. Zhao, L. Fang, Y. Shao, Degradation of atenolol by
3 UV/Peroxymonosulfate: Kinetics, effects of operational parameters and mechanism,
4 Chemosphere 93 (2013) 2717-2724.
5
6 [21] T.W. Chan, N.J.D. Graham, W. Chu, Degradation of iopromide by combined UV
7 irradiation and peroxydisulfate, J. Hazard. Mater. 181 (2010) 508-513.
8
9 [22] T. Olmez-Hanci, I. Arslan-Alaton, Comparison of sulfate and hydroxyl radical based
10 advanced oxidation of phenol, Chem. Eng. J. 24 (2013) 10-16.
11
12 [23] H. Hori, A. Yamamoto, E. Hayakawa, S. Taniyasu, N. Yamashita, S. Kutsuna, H.
13 Kiatagawa, R. Arakawa, Efficient decomposition of environmentally persistent
14 perfluorocarboxylic acids by use of persulfate as a photochemical oxidant, Environ. Sci.
15 Technol. 39 (2005) 2383-2388.
16
17 [24]G.P. Anipsitakis, T.P. Tufano, D.D. Dionysiou, Chemical and microbial decontamination
18 of pool water using activated potassium peroxymonosulfate, Water Res. 42 (2008) 2899-
19 2910.
20 [25] R.H. Waldemer, P.G. Tratnyek, R.L. Johnson, J.T. Nurmi, Oxidation of chlorinated
21 ethenes by heat-activated persulfate: kinetics and products Environ. Sci. Technol. 41 (2007)
22 1010-1015.
23
24 [26] J. Cong, G. Wen, T. Huang, L. Deng, J. Ma, Study on enhanced ozonation degradation
25 of para-chlorobenzoic acid by peroxymonosulfate in aqueous solution. Chem. Eng. J. 264
26 (2015) 399-403.
27
28 [27] J.A. Khan, X. He, N.S. Shah, H.M. Khan, E. Hapeshi, D. Fatta-Kassinos, D.D.
29 Dionysiou, Kinetic and mechanism investigation on the photochemical degradation
30 of atrazine with activated H2O2, S2O8 2− and HSO5−, Chem. Eng. J. 252 (2014) 393–403.
31
32 [28] W.M. Latimer, Oxidation Potentials. Englewood Cliffs, NJ: Prentice-Hall, Inc.; 1952.
33

19
1 [29]C. Liang, Z.-S. Wang, N. Mohanty, Influences of carbonate and chloride ions on
2 persulfate oxidation of trichloroethylene at 20 C, Sci. Total Environ. 370 (2006) 271-277.
3
4 [30] K.-C. Huang, R.A. Couttenye, G.E. Hoag, Kinetics of heat-assisted persulfate oxidation
5 of methyl tert-butyl ether (MTBE), Chemosphere 49 (2002) 413-420.
6
7 [31] G.P. Anipsitakis, D.D. Dionysiou, Radical generation by the interaction of transition
8 metals with common oxidants, Environ. Sci. Technol. 38 (2004) 3705−3712.
9
10 [32] G.P. Anipsitakis, E. Stathatos, D.D. Dionysiou, Heterogeneous activation of oxone using
11 Co3O4, J. Phys. Chem. B. 109 (2005) 13052-13055.
12
13 [33] Yi-F. Huang, Yao-H. Huang, Behavioral evidence of the dominant radicals and
14 intermediates involved in Bisphenol A degradation using an efficient Co2+/PMS oxidation
15 process, J. Hazard. Mater. 167 (2009) 418-426.
16
17 [34] X. Chen, X. Qiao, D. Wang, J.Lin, J. Chen, Kinetics of oxidative decolorization and
18 mineralization of Acid Orange 7 by dark and photoassisted Co2+-catalyzed
19 peroxymonosulfate system, Chemosphere 67 (2007) 802–808.
20
21 [35] S.K. Ling, S. Wang, Y. Peng, Oxidative degradation of dyes in water using Co 2+/H2O2
22 and Co2+/peroxymonosulfate, J. Hazard. Mater. 178 (2010) 385-389.
23
24 [36] Q.J. Yang, H. Choi, D.D. Dionysiou, Nanocrystalline cobalt oxide immobilized on
25 titanium dioxide nanoparticles for the heterogeneous activation of peroxymonosulfate, Appl.
26 Catal. B: Environ. 74 (2007) 170-178.
27
28 [37] P. R. Shukla, S. Wang, H. Sun, H. M. Ang, M. Tade, Activated carbon supported cobalt
29 catalysts for advanced oxidation of organic contaminants in aqueous solution, Appl. Catal. B:
30 Environ. 100 (2010) 529-534.
31

20
1 [38] P. Shukla, H. Sun, S. Wang, H. M. Ang, M.O. Tade, Nanosized Co3O4/SiO2 for
2 heterogeneous oxidation of phenolic contaminants in waste water, Sep. Purif. Technol. 77
3 (2011) 230-236.
4
5 [39] E.A. Betterton, M.R. Hoffmann, Kinetics and mechanism of the oxidation of aqueous
6 hydrogen sulfide by peroxymonosulfate, Environ. Sci. Technol. 24 (1990) 1819-1824.
7
8 [40] Y.-H. Guan, J. Ma, X.-C.Li, J.-Y. Fang, L.-W. Chen, Influence of pH on the formation
9 of sulfate and hydroxyl radicals in the UV/Peroxymonosulfate system, Environ. Sci. Technol.
10 45 (2011) 9308-9314.
11
12 [41] Y.R. Wang, W. Chu, Photo-assisted degradation of 2,4,5-trichlorophenoxyacetic acid by
13 Fe(II)-catalyzed activation of Oxone process: The role of UV irradiation, reaction mechanism
14 and mineralization, Appl. Catal. B: Environ. 123-124 (2012) 151-161.
15
16 [42] X. Gu, S. Lu, Z. Qiu, Q. Sui, Z. Miao, K. Lin, Y. Liu, Q. Luo, Comparison of
17 photodegradation performance of 1,1,1-trichloroethane in aqueous solution with the addition
18 of H2O2 or S2O82- oxidants, Ind. Eng. Chem. Res. 51 (2012) 7196-7204.
19
20 [43] L.R. Bennedsen, J. Muff, E.G. Sogaard, Influence of chloride and carbonates on the
21 reactivity of activated persulfate, Chemosphere 86 (2012) 1092-1097.
22
23 [44] M. Sanchez-Polo, M.M. Abdel Daiem, R. Ocampo-Perez, J. Rivera-Utrilla, A.J. Mota,
24 Comparative study of the photodegradation of bisphenol A by HO●, SO4●‾ and CO3●‾
25 /HCO3‾ radicals in aqueous phase, Sci. Total Environ. 463-464 (2013) 423-431.
26
27 [45] M. Zhan, X. Yang, Q. Xian, L. Kong, Photosensitized degradation of bisphenol A
28 involving reactive oxygen species in the presence of humic substances, Chemosphere 63
29 (2006) 378–386.
30
31 [46] W. Songlin, Z. Ning, W. Si, Z. Qi, Y. Zhi, Modeling the oxidation kinetics of sono-
32 activated persulfate’s process on the degradation of humic acid, Ultrason. Sonochem. 23
33 (2015) 128-134.
34

21
1 [47] M. Deborde, S. Rabouan, P. Mazellier, J.-P. Duguet, B. Legube, Oxidation of bisphenol
2 A by ozone in aqueous solution, Water Res. 42 (2008) 4299-4308.
3
4 [48] R.A. Torres, F. Abdelmalek, E. Combet, C. Petrier, C. Pulgarin, A comparative study of
5 ultrasonic cavitation and Fenton's reagent for bisphenol A degradation in deionised and
6 natural waters, J. Hazard. Mater. 146 (2007) 546-551.
7
8 [49] M.G. Antoniou, A.A. de la Cruz, D.D. Dionysiou, Degradation of microcystin-LR using
9 sulfate radicals generated through photolysis, thermolysis and e− transfer mechanisms, Appl.
10 Catal. B: Environ. 96 (2010) 290-298.
11
12 [50] D. H. Lataye, I. M. Mishra, I. D. Mall, Adsorption of 2-picoline onto bagasse fly ash
13 from aqueous solution, Chem. Eng. J. 138 (2008) 35-46.
14
15 [51] A. Gonzalez-Casado, N. Navas, M. del Olmo, J.L. Vilchez, Determination of Bisphenol
16 A in water by micro liquid-liquid extraction followed by silylation and gas chromatography-
17 mass spectrometry analysis, J. Chromatogr. Sci. 36 (1998) 565-570.
18
19 [52] M. de1 Olmo, A. Gonz6lez-Casado, N.A. Navas, J.L. Vilchez, Determination of
20 bisphenol A (BPA) in water by gas chromatography-mass spectrometry, Anal. Chim. Acta
21 346 (1997) 87-92.
22
23 [53] A. Ghauch, A.M. Tuqan, Oxidation of bisoprolol in heated persulfate/H2O systems:
24 Kinetics and products, Chem. Eng. J. 183 (2012) 162-171.
25
26 [54] A. Ghauch, Al M. Tuqan, N. Kibbi, Ibuprofen removal by heated persulfate in aqueous
27 solution: A kinetics study, Chem. Eng. J. 197 (2012) 483-492.
28
29 [55] Z. Wang, R. T. Bush, L. A. Sullivan, C. Chen, J. Liu, Selective oxidation of arsenite by
30 peroxymonosulfate with high utilization efficiency of oxidant, Environ. Sci. Technol. 48
31 (2014) 3978-3985.
32
33 [56] P. Maruthamuthu, P. Neta, Radiolytic chain decomposition of peroxomonophosphoric
34 and peroxomonosulfuric acids, J. Phys. Chem. 81 (1977) 937-940.

22
1
2 [57] F. Ji, C. Li, L. Deng, Performance of CuO/PMS system: Heterogeneous catalytic
3 oxidation of phenol at ambient conditions, Chem. Eng. J. 178 (2011) 239-243.
4
5 [58] N. S. Shah, X. He, H. M. Khan, J. A. Khan, K. E. O’Shea, D. L. Boccelli, D. D.
6 Dionysiou, Efficient removal of endosulfan from aqueous solution by UV-C/peroxides: A
7 comparative study, J. Hazard. Mater. 263 (2013) 584-592.
8
9 [59] Y. Lee, U. von Gunten, Quantitative structure reactivity relationships (QSARs) for the
10 transformation of organic micropollutants during oxidative water treatment, Water Res. 46
11 (2012) 6177 - 6195.
12
13 [60] T.K. Lau, W. Chu, N.J.W. Graham, The aqueous degradation of butylated
14 hydroxyanisole by UV/S2O82- : Study of reaction mechanisms via dimerization and
15 mineralization, Environ. Sci. Technol. 41 (2007) 613-619.
16
17 [61] C. Liang, H.-W. Su, Identification of sulfate and hydroxyl radicals in thermally
18 activated persulfate, Ind. Eng. Chem. Res. 48 (2009) 5558-5562.
19
20 [62] Z. Wang, R. Yuan, Y. Guo, L. Xu, J. Liu, Effects of chloride ions on bleaching of azo
21 dyes by Co2+/oxone regent: Kinetic analysis, J. Hazard. Mater. 190 (2011) 1083-1087.
22
23 [63] G.-D. Fang, D.D. Dionysiou, Y. Wang, S.R. Al-Abed, D.-M. Zhou, Sulfate radical-based
24 degradation of polychlorinated biphenyls: Effects of chloride ion and reaction kinetics, J.
25 Hazard. Mater. 227-228 (2012) 394-401.
26
27 [64] R. Yuan, S.N. Ramjaun, Z. Wang, J. Liu, Effects of chloride ion on degradation of Acid
28 Orange 7 by sulfate radical-based advanced oxidation process: Implications for formation of
29 chlorinated aromatic compounds, J. Hazard. Mater. 196 (2011) 173-179.
30
31 [65] Y. Chen, K. Zhang, Y.G. Zuo, Direct and indirect photodegradation of estriol in the
32 presence of humic acid, nitrate and iron complexes in water solutions, Sci. Total Environ. 463
33 (2013) 802-809.
34

23
1 [66] S. K. Atkinson, V. L. Marlatt, L. E. Kimpe, D. R. S. Lean, V. L. Trudeau, J. M. Blais,
2 Environmental factors affecting ultraviolet photodegradation rates and estrogenicity of
3 estrone and ethinylestradiol in natural waters, Arch. Environ. Contam.Toxicol. 60 (2011) 1-
4 7.
5
6 [67] G.V. Buxton, C.L. Greenstock, W.P. Helman, A.B. Ross, Critical review of rate constant
7 for reaction hydrated electrons, hydrogen atoms and hydroxyl radicals (HO• /O−) in aqueous
8 solution, J. Phys. Chem. Ref. Data 17 (1988 ) 513-886.
9
10 [68] W-D. Oh, S-K. Lua, Z. Dong, T-T. Lim, High surface area DPA-hematite for efficient
11 detoxification of bisphenol A via peroxymonosulfate activation, J. Mater. Chem. A, 2 (2014)
12 15836-15845.
13
14 [69] S. Yang, .P. Wang, X. Yang, L. Shan, W. Zhang, X. Shao, R. Niu, Degradation
15 efficiencies of azo dye Acid Orange 7 by the interaction of heat, UV and anions with
16 common oxidants: Persulfate, peroxymonosulfate and hydrogen peroxide, J. Hazard. Mater.
17 179 (2010) 552–558.
18
19 [70] T. Olmez-Hanci, I. Arslan-Alaton, B. Genc, Bisphenol A treatment by the hot persulfate
20 process: Oxidation products and acute toxicity, Journal of Hazardous Materials 263 (2013)
21 283-290.
22
23 [71] M. Molkenthin, T. Olmez-Hanci, M.R. Jekel, I. Arslan-Alaton, Photo-Fenton-like
24 treatment of BPA: Effect of UV light source and water matrix on toxicity and transformation
25 products, Water Res. 47 (2013) 5052-5064.
26
27 [72] K.S. Tay, N. Abd. Rahman, Mhd. R.B. Abas, Degradation of bisphenol A by ozonation:
28 rate constants, influence of inorganic anions, and by-products, Maejo Int. J. Sci. Technol. 6
29 (2012) 77-94.
30
31 [73] C. Li, X.Z. Li, N. Graham, N.Y. Gao, The aqueous degradation of bisphenol A and
32 steroid estrogens by ferrate, Water Res. 42 (2008) 109-120.
33

24
1 [74] K. Lin, W. Liu, D.Jaygan, Oxidative removal of bisphenol A by manganese dioxide:
2 Efficacy, Products, and Pathways, Environ. Sci. Technol. 43 (2009) 3860-3864.
3
4 [75] V.C. Mora, J.A. Rosso, D.O. Martire, M.C. Gonzalez, Phenol depletion by thermally
5 activated peroxydisulfate at 70 C, Chemosphere 84 (2011) 1270-1275.
6
7 [76 ] http://chemistry.oregonstate.edu/courses/ch361-464/ch362/irinterp.htm as assessed on
8 11 June 2014.
9
10 [77] M. Akita, D. Tsutsumi, M. Kobayashi, H. Kise, Structural change and catalytic activity
11 of horseradish peroxidase in oxidative polymerization of phenol, Biosci. Biotech. Bioch. 65
12 (2001) 1581-1588.
13
14 [78] K. Chiang, T.M. Lim, L. Tsen, C.C. Lee, Photocatalytic degradation and mineralization
15 of bisphenol A by TiO2 and platinized TiO2, Appl. Catal. A: Gen. 261 (2004) 225-237.
16
17 [79] P.J. Chen, K.G. Linden, D.E. Hinton, S. Kashiwada, E.J. Rosenfeldt, S.W. Kullman,
18 Biological assessment of bisphenol A degradation in water following direct photolysis and
19 UV advanced oxidation, Chemosphere 65 (2006) 1094-1102.
20
21 [80] Y. Ohko, I. Ando, C. Niwa, T. Tatsuma, T. Yamamura, T. Nakashima, Y. Kubota, A.
22 Fujishima, Degradation of Bisphenol A in water by TiO2 photocatalyst, Environ. Sci.
23 Technol. 35 (2001) 2365-2368.
24
25 [81 ] H. Fang, W. Tong, L.M. Shi, R. Blair, R. Perkins, W. Branham, B.S. Hass, Q. Xie, S.L.
26 Dial, C.L. Moland, D.M. Sheehan, Structure-activity relationships for a large diverse set of
27 natural, synthetic, and environmental estrogens, Chem. Res. Toxicol. 14 (2001) 280-294.
28
29 [82] R.O.C. Norman, P.M. Storey, P.R. West, Electron spin resonance studies. Part XXV.
30 Reactions of the sulphate radical anion with organic compounds, J. Chem. Soc. B: Phys. Org.
31 (1970) 1087-1095.
32

25
1 [83] G. Merga, B.S.M. Rao, H. Mohan, J.P. Mittal, Reactions of OH and SO4•− with some
2 halobenzenes and halotoluenes: A radiation chemical study, J. Phys. Chem. 98 (1994) 9158-
3 9164.
4
5 [84] M.G. Antoniou, A.A. de la Cruz, D.D. Dionysiou, Intermediates and reaction pathways
6 from the degradation of Microcystin-LR with sulfate radicals, Environ. Sci. Technol. 44
7 (2010) 7238-7244.
8
9 [85] X. Xie, Y. Zhang, W. Huang, S. Huang, Degradation kinetics and mechanism of aniline
10 by heat-assisted persulfate oxidation, J. Environ. Sci. 24 (2012) 821-826.
11
12 [86 ] P. Neta, V. Madhavan, H. Zemel, R.W. Fessenden, Rate constants and mechanism of
13 reaction of sulfate radical anion with aromatic compounds, J. Am. Chem. Soc. 99 (1977)
14 163-164.
15
16 [87] G. Anipsitakis, D.D. Dionysiou, M.A. Gonzalez, Cobalt-mediated activation of
17 peroxymonosulfate and sulfate radical attack on phenolic compounds. Implications of
18 chloride ions, Environ. Sci. Technol. 40 (2006) 1000-1007.
19
20 [88] M. Foti, K.U. Ingold, J. Lusztyk, The Surprisingly High Reactivity of Phenoxyl
21 Radicals, J. Am. Chem. Soc. 116 (1994) 9440-9447.
22
23 [89] Y.-H. Cui, X.-Y. Li, G. Chen, Electrochemical degradation of bisphenol A on different
24 anodes, Water Res. 43 (2009) 1968-1976.
25
26 [90] T. Miyazaki, Y. Katsumura, M. Lin, Y. Muroya, H. Kudo, M. Taguchi, M. Asano, M.
27 Yoshida, Radiolysis of phenol in aqueous solution at elevated temperatures, Rad. Phys.
28 Chem. 75 (2006) 408-415.
29
30

26
Table captions

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


conditions

Table 1.

Compound/ formula Tentative Structure MW/(m/z) Detected Remaining


intermediates
Reported by/in
other AOPs
Bisphenol-A/C15H16O2 228 √

Formic acid (16.75 min) 116 √

p-phenol-2-hydroxy 168 √

carboxylic acid (19.2

min)

Biphenyldiol (24.9 min) 186 √

Quinone of MHBPA 242 √

(21.11 min)

Benzaldehyde/C7H6O 106 [70]

p-hydroxyacetophenone 136 [16], [71]

Benzophenone 182 [33]

Monohydroxylated BPA 244 √

(MHBPA) 24.1 min


Phenol,2,4-bis(1,1- 206 √

dimethyl ethyl)/C14H22O

(18.2 min)

Phenol,3,5-bis(1,1- 206 √

dimethyl(ethyl)/ C14H22O

2-hydroxy-4-(2-(3- 257 √

hydroxy-4-oxocyclohexa-

2,5-dienyl) propan-2-yl)

cyclohexa-2,5-dienone [or

quinone of dihydroxylated

BPA] (26.11 min)

quinone of dihydroxylated 257 √

BPA

4-isopropyl-1-phenoxy-2- 270 √

isopropyl-4-phenol (29.12

min)

Hydroxylated (4- 286 [12]

isopropyl-1-phenoxy-2-

isopropyl-4-phenol)

Coupling product 320 [12]


Figure and figure captions

S.No. Captions

Fig. 1. Effect of scavengers on the degradation of BPA at 29 ± 3 C: —●— [PMS/BPA],


—○— [TBA /PMS/BPA], —▼— [ethanol /PMS/BPA],
(Inset: PMS utilization efficiency for different systems at 360 min of reaction time)

Fig. 2. Effect of PMS concentration on the degradation efficiency of BPA. [BPA] o= 0.22
mM. [PMS]o = —●— 0.16 mM, —○— 0.32 mM, —▼— 0.66 mM, —∆— 0.98
mM, —■— 1.30 mM, —□— 1.62 mM (Inset: Pseudo-first–order kinetics)

Fig. 3. Time course of BPA degradation under UV/PMS system [PMS]o=0.66 mM; [BPA]o
= —●— 0.04 mM; —○— 0.13 mM; —▼— 0.22 mM; —∆— 0.31mM. (Inset:
initial rate vs BPA concentration)

Fig. 4. Effect of pH on the degradation efficiency of BPA. [BPA]o = 0.22 mM,


[PMS]o = 0.66 mM, —∆— pHN= 5.15, —○— pH = 3, —▼— pH = 5,
—■— pH = 7, —□— pH = 10, —●— pH = 12

Fig. 5. Effect of degradation of BPA using Cl-, HA and HCO3- 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; 10 mg/L; 20 mg/L)

Fig. 6. Time course of fractional removal of BPA and TOC


[BPA]o = 0.22 mM; [PMS]o = 0.66 mM. BPA removal; TOC removal
[Inset: BPA removal, TOC removal, PMS consumption at specific PMS
dosage]
Fig. 7. Possible formation of radical adducts during BPA oxidation by UV/PMS system

Fig. 8. Proposed pathway of BPA degradation by UV/PMS system

1
Fig. 1

2
Fig. 2

3
Fig. 3

4
Fig. 4

5
Fig. 5

6
Fig. 6

7
Fig. 7

8
Fig. 8

9
Revised Highlights

 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.
Graphical abstract

You might also like