You are on page 1of 10

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

Available online at www.sciencedirect.com

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

Aquatic photodegradation of sunscreen agent p-aminobenzoic acid in the presence of dissolved organic matter
Lei Zhou a, Yuefei Ji a,b, Chao Zeng a, Ya Zhang a, Zunyao Wang a, Xi Yang a,*
a b

State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210046, PR China Lyon 1, UMR CNRS 5256, Institut de recherches sur la catalyse et lenvironnement de Lyon (IRCELYON), 2 Avenue Albert Einstein, Universite F-69626 Villeurbanne, France

article info
Article history: Received 16 July 2012 Received in revised form 20 September 2012 Accepted 23 September 2012 Available online 4 October 2012 Keywords: Photolysis Sunscreen agent p-aminobenzoic acid Dissolved organic matter Density functional theory computation

abstract
Dissolved organic matter (DOM) is an important photosensitizer for the phototransformation of organic contaminants in sunlit natural waters. This article focuses on the photolysis kinetics and mechanism of sunscreen agent p-aminobenzoic acid (PABA) in the presence of four kinds of DOM; Suwannee River fulvic acid (SRFA), Suwannee River humic acid (SRHA), Nordic Lake fulvic acid (NOFA) and Nordic Lake humic acid (NOHA). It is evident that direct photolysis of PABA is highly pH-dependent because different species of PABA have different electrical densities on the ring system. The presence of four kinds of DOM inhibits the photolysis of PABA primarily due to their light screening effect. Meanwhile, a complex interaction involving energy transfer, triplet carbonyl group induced electron transfer, and amino acid induced proton abstraction between PABA and DOM is veried by competition kinetics experiments and density functional theory (DFT) computation. In addition, DOMinduced singlet oxygen (1O2) and hydroxyl radical (OH) are determined to play an insignicant role in PABA photolysis by competition dynamics method. Photoproducts identication using solid phase extractioneliquid chromatographyemass spectrometry (SPEeLCeMS) techniques reveals that the distribution of the photoproducts could not be affected by the addition of DOM. Two photodegradation pathways of PABA are temporarily proposed, in which the di(tri)-polymerization of intermediates are the dominant pathway whereas the oxidation of amino group to nitryl followed by hydroxylation is a minor process. Our ndings reveal that direct photolysis is the dominant transformation pathway of PABA in natural sunlit waters, while the presence of DOM could evidently inuence such process by light screening effect, energy transfer, electron transfer and proton abstraction mechanism. The ndings in this study provide useful information for understanding of interaction between DOM and organic contaminants. 2012 Elsevier Ltd. All rights reserved.

1.

Introduction

Pharmaceuticals and personal care products (PPCPs) are emerging contaminants that have drawn extensive attention

due to their potential hazardous effect on ecological system and the human beings (Yu and Wu, 2012). Since PPCPs had been designed to be lipophilic and biologically persistent, quite a lot of them resist to biodegradation (Halling-Sorensen

* Corresponding author. Tel.: 86 25 8968 0357. E-mail address: yangxi@nju.edu.cn (X. Yang). 0043-1354/$ e see front matter 2012 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.watres.2012.09.045

154

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

et al., 1998; Daughton and Ternes, 1999; 2000). Consequently, a wide range of PPCPs have been detected throughout the world in surface water, soil, sludge, sh and even the human body (Balmer et al., 2005; Buser et al., 2006; Giokas et al., 2007; Calafat et al., 2008). Sunscreens have been regarded as one of the important PPCPs due to their increasing use for prevention of skin cancer. They are discharged barely modied when entering wastewater treatment plants, while those used in recreational water activities may enter natural waters directly (Balmer et al., 2006; Diaz-Cruz and Barcelo, 2009). Humans are exposed to sunscreens mainly through two pathways, directly through dermal exposure by using cosmetics, and indirectly through the food chain (Schlumpf et al., 2008; Swan et al., 2008). Evidence is accumulating that sunscreens have endocrine activities on both the animal and human body (Stevenson and Davies, 1999; Schlumpf et al., 2004; Eichenseher, 2006; Blitz and Norton, 2008). However, the long term effects of sunscreens on the environment and humans are still unclear so far. Since sunscreens absorb solar energy, clarifying their photolysis behaviors in natural waters is of great importance for understanding their environmental fate and ecology risk assessment. Photodegradation of sunscreens in natural waters involves direct photolysis when their absorption spectrums overlap with the solar irradiation spectrum and indirect photolysis by reaction with reactive oxygen species, e.g., singlet oxygen (1O2) and hydroxyl radical (OH) formed by the photolysis of dissolved organic matter (DOM). As an important photosensitizer, DOM is widely present in natural waters and signicantly affects light-induced transformation of organic contaminants (Kohn et al., 2007; Guerard et al., 2009; Page et al., 2011). Direct reactions with triplet excited state DOM (3DOM*) by energy and/or electron transfer may also contribute to the fate of organic contaminants (Gerecke et al., 2001; Boreen et al., 2005; Housari et al., 2010). In some cases, DOM may also quench the triplet excited states and inhibit the transformation of organic contaminants (Wenk et al., 2011). Since patented in 1943, p-aminobenzoic acid (PABA) has been widely used in both sunscreen and pharmaceutical products (Burger, 1967; Osgood et al., 1982; Allen et al., 1995; He et al., 2005). The continuous input of PABA into the environment through personal care applications has caused particular concerns since previous studies demonstrated that PABA can increase photosensitivity and lead to DNA damage (Brash and Haseltine, 1982; Dromgoole and Maibach, 1990). Therefore, the investigation of PABA photolysis in natural surface waters is of great importance for understanding its fate and aquatic risk on ecological system. A preliminary study of the effect of nitrate, bicarbonate and DOM on PABA photolysis was performed (Mao et al., 2011). However, there is still a need for detailed information concerning the complex inuence of DOM on PABA photolysis. In this study we investigated the role of four kinds of DOM; Suwannee River fulvic acid (SRFA), Suwannee River humic acid (SRHA), Nordic Lake fulvic acid (NOFA) and Nordic Lake humic acid (NOHA), on the photolysis of PABA under simulated solar irradiation in aquatic solutions. Verication of the mechanisms of the selected DOM on PABA photolysis was carried out via competition kinetics, photoproducts

identication, as well as theoretical computation for the possible reactions.

2.
2.1.

Experiment section
Chemicals and materials

PABA ( p-aminobenzoic acid, 99%) and furfuryl alcohol (FFA, 99%) were purchased from SigmaeAldrich (St. Louis, Missouri). Four kinds of DOM; SRFA (1R101F), SRHA (1R101H), NOFA (1R105F) and NOHA (1R105H) of reference grade were obtained from International Humic Substance Society (IHSS). Sorbic acid (99%), and p-chlorobenzoic (PCBA, 99%) were supplied by TCI (Japan). Benzophenone (BP, 99%) was purchased from J&K (China). HPLC grade methanol and acetic acid were obtained from Tedia Company (Faireld, USA). Deionized (DI) water was prepared from a Millipore Milli-Q System. All the solutions were prepared by dissolving the reagents directly in DI water without further purication, stored under 4  C and used within one month.

2.2.

Irradiation experiments

Steady-irradiation experiments were performed in the XPA-2 photochemical reactor (Nanjing Xujiang Motor Factory, China), equipped with a 1000 W xenon arc lamp (Beijing Electric Light Source Institute, China). The lamp was placed inside a quartz cooling well. Capped pyrex reactor (l > 280 nm) containing 50 mL reaction solution was placed in a merry-goground unit at a xed distance. Aliquots of 1 mL were sampled at different irradiation times before analysis.

2.3.

Analytical methods

The concentrations of PABA, FFA and PCBA were analyzed by HPLC. A description of the HPLC systems including HPLC conditions, detection wavelength and separation columns are given together in Table 1. Concentrations of DOM were expressed as total organic carbon (TOC, mg C L1) and determined by a Shimadzu 5000A analyzer (Japan). Absorbance scans of PABA and DOM solutions for the wavelength of 280e330 nm were collected on a UVeVis spectrometer (UV2450, Shimadzu, Japan). Identication of PABA photoproducts was carried out using the SPEeLCeMS method. The solutions after irradiation were concentrated by solid phase extraction (SPE) workstation using HLB-C18 cartridges (6 CC/200 mg, Waters Oasis). Prior to extraction, the HLB-C18 cartridges were activated by 5 mL methanol and 10 mL DI water in sequence. Subsequently, 100 mL reaction solutions were percolated through the cartridges at ows of 0.5 mL min1; then the cartridges were purged completely dry by puried N2. The nal extracted photoproducts were obtained by eluting with 2 2 mL methanol and analyzed by LCeMS. LCeMS analysis was conducted on a Thermo Finnigan Surveyor Modular HPLC system equipped with a Finnigan LCQ Advantage MAX ion trap mass spectrometer. The LC separation was achieved by using an Agilent SB-C18 column (5 mm,

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

155

Table 1 e HPLC analysis parameters. Compound Separationa column


TC-C18 Zobrax SB-C18 Zobrax SB-C18

Flow rate (mL/min)


1.0 1.0 1.0

Mobile phase composition Water


80% 40% 40%
b

Methanol
20% 60% 60%

Injection volume (mL)


10 20 20

Detection wavelength (nm)


282 216 240

PABA FFA PCBA

a The specication of the columns: 5 mm, 250 mm 4.6 mm, all the columns were purchased from Aglient. b the water phase contains 0.5% (by volume) acetic acid.

150 mm 4.6 mm). The mobile phase was a mixture of water (A) and methanol (B) under a linear gradient from 5% B to 95% B in 25 min with a ow rate of 0.2 mL min1. An electrospray interface (ESI) was used for the MS and MS2 measurement in negative ionization mode and full scan acquisition between m/z 50e400. The spray voltage was 4.5 kV and an ion-transfer capillary temperature was 200  C. Nitrogen was used as desolvation and nebulizer gas at a ow rate of 35 arbitrary unit and argon was used as a collision gas at 0.25 MPa.

2.4.

Calculation

The calculation of light screening factor Sl and the quantum yield V of the photolysis are described in Text S1 and Text S2 (Supplementary Material).

2.5.

Computational analysis

DFT calculations were performed employing the Gaussian 09 software package. Initial geometries of the investigated compounds were optimized using the hybrid density functional B3LYP method with the 6311 G* basis set and then conrmed by the frequency calculations. The integral equation formalism polarized continuum model (IEFPCM), based on the self-consistent-reaction-eld (SCRF) method, was used to consider the bulk solvent effect of water (Tomasi et al., 2005). The effect of hydrogen bonds was also taken into account. Vertical electron energy and vertical ionization energy were computed using the difference of separate selfconsistent eld (SCF) energies of the neutral and corresponding anionic and cationic systems. The excitation energy from the ground state to lowest singlet and triplet excited state (neutral state, protonation state and deprotonation state) were also obtained from differences of SCF energies calculated at different states.

PABA in aqueous solutions at various pH values. As seen, photolysis reactions of PABA in different pH solutions all follow pseudo-rst-order kinetics. The rate constant of PABA photolysis, k (s1), accelerates with increasing pH value (Table S1). For instance, k value is determined to be (3.46 0.14) 106, (12.57 0.18) 106 and (14.96 0.33) 106 s1 at pH 4.8, 7.8 and 9.2, respectively. In pH 3.0 solution, the removal of PABA caused by direct photolysis is negligible after 6 h of irradiation. It is apparent that PABA direct photolysis was highly pH-dependent because high pH of PABA solution can lead to rapid PABA photolysis. PABA, (with pKa1 of 2.38 and pKa2 of 4.85), has three different species at varying pH, as shown in Fig. 1 (B) (Cox and Stewart, 1976; IUPAC, 1979). When at pH > 7.0, anion form of PABA is expected to be dominant. Our nding is in line with a previous study by Shaw et al. wherein the photodegradation of PABA was highly pH dependent and the deprotonated PABA radical cation was an important intermediate for the formation of some products (Shaw et al., 1992). Electrons from negatively charged carboxylate group are capable of delocalizing to aromatic ring which enrich electron density of ring, which leads to the blue shift of PABA UVeVis absorption spectrum, as shown in Figure S2. Thus, the high photo-activity of anionic PABA was most probably attributed to its high electrical density on the ring system (Mao et al., 2011). To further verify the relationship between photoactivity and solution pH, density functional theory (DFT) calculation of lowest singlet excitation energies of PABA moieties was performed. The calculated singlet energy values (eV) of different PABA forms are 4.43 (anionic) < 4.59 (neutral) < 4.78 (cationic) (Table S2), indicating anionic PABA is more susceptible to photolysis than neutral or cationic moiety due to its lower excitation energy. Therefore, considering the high photoactivity of PABA anionic moiety and pH value of the natural surface water, we selected pH 7.8 in the following work.

3.2.

Effect of DOM on the kinetics of PABA photolysis

3.
3.1.

Results and discussion


Direct photolysis of PABA at various pH values
The presence of DOM in aqueous solutions signicantly inuences the light-induced transformation of organic compounds due to the abundant chromophoric moieties. The effect of DOM on photolysis may be complex, which includes light screening, reactive oxygen species (ROS) generation or quenching, etc (Minella et al., 2011). Photolysis of PABA by steady-state irradiation methods in the presence of four kinds of DOM was carried out to clarify the effect of different DOM on PABA photodegradation. The observed photodegradation

The UVeVis absorbance spectrum of aqueous PABA under different pH values indicates the possibility that direct photolysis of PABA could occur due to the overlap of PABA UVeVis absorbance spectrum and the emission spectrum of our solar simulator (l > 280 nm) (Figure S2, Supplementary Material). Fig. 1(A) shows the kinetics of direct photolysis of

156

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

Fig. 1 e The direct photolysis of PABA (10 mg LL1) at different pH under simulated solar irradiation (A) and the fraction of PABA moieties at different pH value (B); pH value of the solutions was adjusted by NaOH and HCl.

rate constants of PABA in the presence of DOM at different concentrations are listed in Table 2. It can be perceived that the presence of DOM generally inhibits direct photolysis of PABA. Moreover, the quantum yields of PABA photolysis with DOM are less than that of direct photolysis without DOM (V0 0.0039 0.0001), as shown in Table S3.

3.2.1.

Light screening effect of DOM

UV absorbance spectrum of the four kinds of DOM at different concentrations indicates that DOM can absorb sunlight which could weaken the radiant ux necessary for PABA photolysis (Figure S5). To investigate the effect of DOM on PABA photolysis via indirect pathways, inner correction was applied to normalize the light intensity and eliminate the effect caused by absorbing sunlight (Mao et al., 2011). Table 3 displays the integrated light screening factor of four DOMs under different concentrations. As can be seen, the Sl value decreases remarkably with the increasing of DOM concentration, which can be explained by the competition for photons between PABA and DOM. On the other hand, the possibility of photosensitized degradation of PABA in the presence of DOM may also exist. Therefore, considering the light screening effect, and observed degradation rate constant of PABA photolysis in the presence of DOM, kobs, can be described by the following equation. kobs kD Sl kDOM (1)

photosensitized effect of DOM in the presence of PABA. A negative kDOM value indicates an inhibition effect of DOM on PABA photodegradation except light screening; otherwise, DOM enhances the removal of PABA when kDOM value is positive. Fig. 2 displays the kDOM values of PABA photodegradation in the presence of four different DOMs at varying concentrations. At low concentrations of DOM (e.g., 1, 2, 5 mg C L1), kDOM values are normally negative, indicating that direct photolysis dominated over the degradation process while DOM-induced indirect photolysis played a minor role. However, at high concentrations (e.g., 10, 20 mg C L1), kDOM values are positive, suggesting that DOM-induced enhancement of PABA photolysis is more important than inhibition in the presence of these DOMs. It is interesting to note that in the case of SRHA, the inhibition effect is not as pronounced as other two DOMs at low DOM concentrations.

3.2.2.

Possible mechanism of DOM-induced PABA photolysis

where kD is the direct degradation rate constant in the presence of PABA alone, kDOM is the rate constant caused by the

Previous studies have demonstrated that DOM can act as both sensitizer and inhibitor in the photolysis of organic pollutants (Gerecke et al., 2001; Boreen et al., 2005; Housari et al., 2010; Wenk et al., 2011). As stated above, it can be concluded that both of the two kinds of roles are involved in DOM-induced PABA photodegradation. The possible reactions of PABA photodegradation in the presence of DOM were temporarily proposed: (1) DOM promoted or inhibited PABA photolysis by generating or quenching ROS. (2) Interaction between PABA and DOM.

Table 2 e Observed photolysis rate constant of PABA (kPABA) in the presence of four kinds of DOM at varying concentrations, CPABA [ 10 mg LL1, pH [ 7.8. CDOM (mg C L1)
1 2 5 10 20

Table 3 e Light screening factor of four kinds of DOM for wavelengths 280e330 nm. CDOM (mg C L1)
1 2 5 10 20

k(106 s1) SRFA


9.47 8.35 8.09 10.37 7.39

Sl SRFA
0.94 0.87 0.74 0.60 0.40

NOFA

SRHA

NOHA

NOFA
0.93 0.87 0.74 0.56 0.36

SRHA
0.90 0.87 0.68 0.50 0.38

NOHA
0.93 0.90 0.77 0.63 0.42

0.09 9.67 0.24 9.09 0.18 8.03 0.18 11.06 0.20 10.14

0.22 12.44 0.35 12.00 0.35 0.18 10.78 0.26 8.18 0.11 0.14 9.47 0.14 7.34 0.19 0.15 9.33 0.14 9.74 0.10 0.10 7.43 0.17 8.35 0.08

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

157

(Haag et al., 1984; Zhan et al., 2006). However, no obvious change is observed, suggesting 1O2-induced PABA degradation unlikely occurs during photolysis. A similar phenomenon is also observed after the addition of 2-propanol, a molecular quencher of $OH (Aguer and Richard, 1996), indicating that $OH played an insignicant role in the SRFAinduced photolysis. Therefore, our observation implies a weak feasibility of 1O2/$OH-induced PABA elimination in the presence of DOM. Thus, we can conclude that DO is a good energy quencher for 3SRFA*, and 3SRFA*-induced PABA removal may occur.

3.2.2.2. Part 2. calculations for interaction between PABA and SRFA. The possible photoinduced energy and electron transfer reactions of PABA in the presence of SRFA are listed in Table 4. DFT calculation was adopted to predict the possibilities of these reactions (Table S3). The analog structure of SRFA was employed for computation, as shown in Fig S7 (Leenheer et al., 1990). For the autoionization between PABA and 1PABA */3PABA * (reaction 1), DG1 DG2 (5.55/4.59 eV) < 0, indicate the spontaneity of these reactions. Moreover, since DG3 (9.98/8.06 eV) < 0, PABA$and PABA$could also be generated by the autoionization of 1PABA */3PABA *. The triplet state of SRFA, acts as electron donor and accepter in reactions 3 and 4, respectively. DG4 and DG5 were calculated according to the VIET1 (3.36 eV) and VEAT1 (4.54 eV) values of SRFA, we calculated for these two reactions (Zhang et al., 2010). Results with DG4 (0.04 eV) > 0, and DG5 (2.34 eV) < 0, show that 3SRFA* acts more as an electron acceptor than donor in the photolysis of PABA. Since anilines are electron-rich, it is reasonable that PABA shows a strong electron donating ability. The ET1 value of SRFA was 2.37 eV, which is lower than ET1 (PABA) (3.47 eV). Therefore, PABA could not be sensitized by 3 SRFA* through energy transfer. On the contrary, SRFA can accept energy from 3PABA*. In terms of energy transfer pathway, SRFA could inhibit the photolysis by scavenging the triplet state PABA. The computational results indicate that there is a strong possibility for the energy and electron transfer reactions in the PABA photolysis in the presence of SRFA.

Fig. 2 e Values of kDOM in four DOMs at varying concentrations.

A series of experiments and computations were carried out to check the three proposed reaction processes, and SRFA was selected as the reference DOM. The results obtained are shown as following.

3.2.2.1. Part 1. the effect of dissolved oxygen and possible formation of $OH. Dissolved oxygen (DO) plays an important
role in the photolysis of organic contaminants in the presence of DOM. The possible product 1O2 might favor the photolysis of PABA (Allen et al., 1995, 1996). Meanwhile, DO is also an energy quencher of singlet/triplet state DOM (Zepp et al., 1985). The absence of DO increases the decay of PABA in the presence of SRFA, suggesting a detrimental effect of DO on PABA photolysis (Fig. 3). The inhibition effect of DO is most likely attributed to the competition for 3SRFA* with PABA, which is consistent with previous ndings (Chen et al., 2009). Moreover, the effect of 1O2 on PABA photolysis is further evaluated by adding FFA, a widely used chemical trap for 1O2

3.2.2.3. Part 3. verication of the possible mechanisms. To


verify the computational inference, sorbic acid and benzophenone were selected as substitutes of SRFA to check the energy and electron transfer reactions, respectively. Sorbic acid, with high singlet excitation energy (4.64 eV) and moderately low triplet excitation energy (2.46 eV), which is close to that of SRFA, is difcult to interact with most singlet sensitizers except the energetic ones, thus, quenching a variety of triplets seems relatively easy (Zepp et al., 1985; Velosa et al., 2007). Grebel et al. have determined the stead-state concentration of the triplet excited state of organic compounds by using sorbic acid as a quantitative probe (Grebel et al., 2011). The signicant part of the carbon atom in DOM is present as carbonyl group (Canonica et al., 1995). The triplet state aromatic ketones, e.g., benzophenone (BP), have been chosen to model DOM triplet states having oxidative character in previous studies (Canonica et al., 2000, 2006).

Fig. 3 e The effect of dissolved oxygen and possible formation of $OH in the presence of SRFA, CSRFA [ 10 mg C LL1; CPABA [ 10 mg LL1; nitrogen was purged for 30 min; pH [ 7.8.

158

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

Table 4 e Possible photoinduced energy and electron transfer reaction of PABA and corresponding evaluating criteria in the presence of SRFA. No
1 2 3 4 5 6
1 3

Reaction
PABA */ PABA * PABA / PABA
$

Criterion
PABA
$

PABA */3PABA * / PABA$ PABA$ SRFA* PABA / SRFA PABA$ 3 SRFA* PABA / SRFA PABA$ 3 SRFA* PABA / SRFA 3PABA* SRFA 3PABA* / 3SRFA* PABA
1 3

DG1 VEAS0VIEb < 0 DG2 VEAS1/T1VIES0 < 0c DG3 VEA S1/T1 VIES1/T1 < 0 DG4 VEAS0 VIET1(SRFA) < 0 DG5 VEA T1(SRFA) VIES0 < 0 ET1(SRFA) > ET1(PABA) ET1(SRFA) < ET1(PABA)

a VEA (vertical electron energy) is the energy required to remove an electron from a molecule to evaluate the electron accepting ability (EAA). A low VEA value implies a strong EAA b VIE (vertical ionization energy) is the energy required for a molecule to accept an electron to evaluate the electron donating ability (EDA); a low VIE value implies a strong EDA. c DG1 and DG2 are for the electron transfer reactions with PABAS0 as electron acceptor and donor, respectively.

The different content of chromospheres function groups may be responsible for the discrepancy of the PABA degradation efciency, especially the amino acids moiety. According to the website of IHSS, amino acids are important compositions of DOM (IHSS website). Suzuki et al. have demonstrated that the basic amino acids, including histidine, lysine and arginine, can accelerate the photoreaction of ketoprofen (Suzuki et al., 2012). Here, we used glycine and histidine to test the effect of amino acid on PABA photolysis (Fig. 4). Fig. 4 presents the photolysis of PABA with the addition of sorbic acid (20 mM) and BP (50 mM). The presence of sorbic acid decreases the photolysis of PABA, suggesting that the energy transfer from 3PABA* to sorbic acid is involved in the photolysis. On the other hand, it can be seen that 3BP* shows strong oxidation ability toward PABA, which conrms the computational results. Therefore, we can conclude that both of the energy and electron transfer reactions between PABA and SRFA occur in the photolysis progress. Moreover, both glycine (20 mM) and histidine (20 mM) show prominent enhancement effects on PABA photolysis, which is consistent with a previous study. Amino acid is regarded as a key species for the proton abstraction reaction of contaminants. Consequently, the photolysis of PABA can be accelerated by the addition of amino acids.

In conclusion, we deduce that there are three kinds of mechanism among the interaction between PABA and DOM: 1) energy transfer, 2) triplet carbonyl group induced electron transfer, and 3) amino acid induced proton abstraction.

3.2.2.4. Part 4. difference among four selected DOMs at low concentrations. From Fig. 2, the selected DOMs have different
effects on PABA photolysis at low DOM concentrations. The difference could be explained by the combined action of three different mechanisms. It is suggested that amino acid content in SRHA and NOHA is higher than that in SRFA and NOFA (IHSS website). It can be expected that DOMs with the higher content of amino acid group usually exhibited promoted effect on PABA photodegradation. Thus, for SRHA and NOHA, all of the three mechanisms contribute to the PABA removal. While for SRFA and NOFA, energy transfer and triplet carbonyl group induced oxidation dominate due to the very low amino acid content. There is the possibility that the status of energy and electron transfer is related to the steady-concentration of 3DOM*. Energy transfer reaction from 3PABA* to DOM seems signicant at low levels, while, triplet carbonyl group-induced oxidation is dominating at higher DOM concentrations associated with high [3DOM*]SS. The proton abstraction effect is just related to the total amino acid compositions. For NOHA, we speculate that the energy transfer effect is stronger than proton abstraction at a low concentration, resulting in a similar apparent phenomenon to SRFA and NOFA. On the contrary, for SRHA, the enhancement effect of proton abstraction may be dominant.

3.3.

Identication of PABA photoproducts

Fig. 4 e The effect of sorbic acid, benzophenone, glycine and histidine on the photolysis of PABA.

The structural identication of the products during photolytic process was based on the analysis of the total ion chromatogram (TIC) and the corresponding mass spectrum. Table S4 presents the main fragment ion of PABA products and the proposed molecular structure in the presence of SRFA. Based on the analyzed results from LCeMS/MS spectrum, two main possible photodegradation pathways of PABA in the presence of SRFA upon simulated solar irradiation were temporarily proposed (Fig. 5). Pathway 1 includes the dipolymerization and tri-polymerization of an intermediate,

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

159

Fig. 5 e Possible photodegradation pathways of PABA in the presence of SRFA upon simulated solar irradiation.

160

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

and the hydroxylation product of PABA, p-hydroxybenzoic (PHBA), leading to the formation of a series of polymerized benzoic acid derivatives with biphenyl or terphenyl structure. It was assumed that PHBA could quickly convert to various reactive molecular radicals upon irradiation; these radicals could further react with each other, and consequently result in polymerization. Intermediate product PHBA was observed in direct photolysis of PABA in a previous report, and PHBA may be generated by the reaction of ionization intermediates of PABA and H2O (Beltran-Heredia et al., 2001a). However, it should be noted that the PHBA was not found in our experiment, which was most likely due to the high photo-reactivity of PHBA. Previous studies found that PHBA decay very fast under irradiation, and the direct photolysis constant rate is about 5.3 104 s1 at the same condition, which was much larger than PABA (Beltran-Heredia et al., 2001b). Thus, the transformation from PABA to PHBA is expected to be the rate control step in pathway 1. Interestingly, after 36 h irradiation, only three photoproducts (3, 4 and 6) were detected by LCeMS/MS compared with 180 h irradiation. Therefore, it is possible that prolonging the irradiation time favored the polymerization of PHBA. In addition, pathway 2 is most likely to be a minor degradation pathway because photoproducts 1 and 2 could only be determined at relative longer irradiation time.

Appendix A. Supplementary data


Supplementary data related to this article can be found at http://dx.doi.org/10.1016/j.watres.2012.09.045.

references

4.

Conclusions

In this study, the kinetics and mechanism of PABA photolysis in the presence of four kinds of DOM were investigated by experiments and theoretical calculation. In general, direct photolysis is the main pathway of PABA photodegradation in natural waters. The photolysis of PABA obeys pseudo-rstorder kinetics and the rate constant is remarkably enhanced with the increasing solution pH value, probably due to the rich electrical density on the ring system associated with anionic form PABA. In the presence of DOM, PABA photolysis is remarkably decreased primarily due to light screening effect. A complex interaction between DOM and PABA is veried via both experiment and computational analysis, which includes energy transfer from 3PABA* to DOM, triplet carbonyl groupinduced electron transfer, and amino acid-induced proton abstraction. In addition, 1O2 and $OH are unlikely to play a key role in the DOM-induced photolysis. Direct photolysis is the dominant transformation pathway of PABA in natural sunlit waters, while the presence of DOM could evidently inuence such process by light screening effect, energy transfer, electron transfer and proton abstraction mechanism. SPEeLCeMS analysis results reveal that photoproducts of PABA photolysis with or without DOM are almost the same. Photodegradation pathways include the di(tri)-polymerization and the oxidation of the intermediate PHBA.

Acknowledgment
This work was supported by National Natural Science Foundation of China (20977045 and 21177056) and Fundamental Research Funds for the Central Universities (1112021101).

Aguer, J.P., Richard, C., 1996. Reactive species produced on irradiation at 365 nm of aqueous solutions of humic acids. Journal of Physical Chemistry A 93 (2e3), 193e198. Allen, J.M., Engenolf, S., Allen, S.K., 1995. Rapid reaction of singlet molecular-oxygen (O-1(2)) with p-aminobenzoic acid (Paba) in aqueous-solution. Biochemical and Biophysical Research Communications 212 (3), 1145e1151. Allen, J.M., Gossett, C.J., Allen, S.K., 1996. Photochemical formation of singlet molecular oxygen (1O2) in illuminated aqueous solutions of p-aminobenzoic acid (PABA). Journal of Physical Chemistry B 32, 33e37. Balmer, M.E., Buser, H.R., Muller, M.D., Poiger, T., 2005. Occurrence of some organic UV lters in wastewater, in surface waters, and in sh from Swiss lakes. Environmental Science & Technology 39 (4), 953e962. Balmer, M.E., Buser, H.R., Poiger, T., 2006. Entry pathways of UV lters from sunscreens to Swiss lakes. Chimia 60 (1e2), 95. Beltran-Heredia, J., Torregrosa, J., Dominguez, J.R., Peres, J.A., 2001a. Comparison of the degradation of p-hydroxybenzoic acid in aqueous solution by several oxidation processes. Chemosphere 42 (4), 351e359. Beltran-Heredia, J., Torregrosa, J., Dominguez, J.R., Peres, J.A., 2001b. Kinetics of the oxidation of p-hydroxybenzoic acid by the H2O2/UV system. Industrial & Engineering Chemistry Research 40 (14), 3104e3108. Blitz, J.B., Norton, S.A., 2008. Possible environmental effects of sunscreen run-off. Journal of the American Academy of Dermatology 59 (5), 898. Boreen, A.L., Arnold, W.A., McNeill, K., 2005. Triplet-sensitized photodegradation of sulfa drugs containing six-membered heterocyclic groups: identication of an SO2 extrusion photoproduct. Environmental Science & Technology 39 (10), 3630e3638. Brash, D.E., Haseltine, W.A., 1982. Uv-Induced mutation hotspots occur at DNA damage hotspots. Nature 298 (5870), 189e192. Burger, A., 1967. West, Es e textbook of biochemistry. Journal of Medicinal Chemistry 10 (2), 296. Buser, H.R., Balmer, M.E., Schmid, P., Kohler, M., 2006. Occurrence of UV lters 4-methylbenzylidene camphor and octocrylene in sh from various swiss rivers with inputs from wastewater treatment plants. Environmental Science & Technology 40 (5), 1427e1431. Calafat, A.M., Wong, L.Y., Ye, X.Y., Reidy, J.A., Needham, L.L., 2008. Concentrations of the sunscreen agent benzophenone-3 in residents of the United States: national health and nutrition examination survey 2003e2004. Environmental Health Perspectives 116 (7), 893e897. Canonica, S., Jans, U., Stemmler, K., Hoigne, J., 1995. Transformation kinetics of phenols in water: photosensitization by dissolved natural organic material and aromatic ketones. Environmental Science & Technology 29, 1822e1831. Canonica, S., Hellrung, B., Wirz, J., 2000. Oxidation of phenols by triplet aromatic ketones in aqueous solution. Journal of Physical Chemistry A 104 (6), 1226e1232. Canonica, S., Hellrung, B., Muller, P., Wirz, J., 2006. Aqueous oxidation of phenylurea herbicides by triplet aromatic ketones. Environmental Science & Technology 40 (21), 6636e6641.

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

161

Chen, Y., Hu, C., Hu, X.X., Qu, J.H., 2009. Indirect photodegradation of amine drugs in aqueous solution under simulated sunlight. Environmental Science & Technology 43 (8), 2760e2765. Cox, R.A., Stewart, R., 1976. Ionization of feeble organic-acids in dmso-water mixtures - acidity constants derived by extrapolation to aqueous state. Journal of the American Chemical Society 98 (2), 488e494. Daughton, C.G., Ternes, T.A., 1999. Pharmaceuticals and personal care products in the environment: agents of subtle change? Environmental Health Perspectives 107, 907e938. Daughton, C.H., Ternes, T.A., 2000. Special report: pharmaceuticals and personal care products in the environment: agents of subtle change? (vol 107, pg 907, 1999). Environmental Health Perspectives 108, 598. Diaz-Cruz, M.S., Barcelo, D., 2009. Chemical analysis and ecotoxicological effects of organic UV-absorbing compounds in aquatic ecosystems. Trac-Trends in Analytical Chemistry 28 (6), 708e717. Dromgoole, S.H., Maibach, H.I., 1990. Sunscreening agent intolerance e contact and photocontact sensitization and contact urticaria. Journal of the American Academy of Dermatology 22 (6), 1068e1078. Eichenseher, T., 2006. The cloudy side of sunscreens. Environmental Science & Technology 40 (5), 1377e1378. Elemental compositions and stable isotopic ratios of IHSS samples: http://www.humicsubstances.org/elements.html. Gerecke, A.C., Canonica, S., Muller, S.R., Scharer, M., Schwarzenbach, R.P., 2001. Quantication of dissolved natural organic matter (DOM) mediated phototransformation of phenylurea herbicides in lakes. Environmental Science & Technology 35 (19), 3915e3923. Giokas, D.L., Salvador, A., Chisvert, A., 2007. UV lters: from sunscreens to human body and the environment. Trac-Trends in Analytical Chemistry 26 (5), 360e374. Grebel, J.E., Pignatello, J.J., Mitch, W.A., 2011. Sorbic acid as a quantitative probe for the formation, scavenging and steady-state concentrations of the tripletexcited state of organic compounds. Water Research 45 (19), 6535e6544. Guerard, J.J., Miller, P.L., Trouts, T.D., Chin, Y.P., 2009. The role of fulvic acid composition in the photosensitized degradation of aquatic contaminants. Aquatic Science 71 (2), 160e169. Haag, W.R., Hoigne, J., Gassman, E., Braun, A.M., 1984. Singlet oxygen in surface waters .1. Furfuryl alcohol as a trapping agent. Chemosphere 13 (5e6), 631e640. Halling-Sorensen, B., Nielsen, S.N., Lanzky, P.F., Ingerslev, F., Lutzhoft, H.C.H., Jorgensen, S.E., 1998. Occurrence, fate and effects of pharmaceutical substances in the environment e a review. Chemosphere 36 (2), 357e394. He, Y.G., Wu, C.Y., Kong, W., 2005. Theoretical and experimental studies of water complexes of p- and o-aminobenzoic acid. Journal of Physical Chemistry A 109 (12), 2809e2815. Housari, F., Vione, D., Chiron, S., Barbati, S., 2010. Reactive photoinduced species in estuarine waters. Characterization of hydroxyl radical, singlet oxygen and dissolved organic matter triplet state in natural oxidation processes. Photochemical & Photobiological Sciences 9 (1), 78e86. International Union of Pure and Applied Chemistry. Commission on Equilibrium Data.; Serjeant, E. P.; Dempsey, B.; International Union of Pure and Applied Chemistry. Commission on Electrochemical Data, 1979. Ionisation Constants of Organic Acids in Aqueous Solution. Pergamon Press, Oxford ; New York, p xi, 989 pp. Kohn, T., Grandbois, M., McNeill, K., Nelson, K.L., 2007. Association with natural organic matter enhances the

sunlight-mediated inactivation of MS2 coliphage by singlet oxygen. Environmental Science & Technology 41 (13), 4626e4632. Leenheer, J.A., Mcknight, D.M., Thurman, E.M., Maccarthy, P., 1990. Structural components and proposed structural models of fulvic-acid from the Suwannee River. Abstracts of Papers of the American Chemical Society 200, 48. Mao, L., Meng, C., Zeng, C., Ji, Y., Yang, X., Gao, S., 2011. The effect of nitrate, bicarbonate and natural organic matter on the degradation of sunscreen agent p-aminobenzoic acid by simulated solar irradiation. Science of the Total Environment 409 (24), 5376e5381. Minella, M., Rogora, M., Vione, D., Maurino, V., Minero, C., 2011. A model approach to assess the long-term trends of indirect photochemistry in lake water. The case of Lake Maggiore (NW Italy). Science of the Total Environment 409 (18), 3463e3471. Osgood, P.J., Moss, S.H., Davies, D.J.G., 1982. The sensitization of near-ultraviolet radiation killing of mammalian-cells by the sunscreen agent para-aminobenzoic acid. Journal of Investigative Dermatology 79 (6), 354e357. Page, S.E., Arnold, W.A., McNeill, K., 2011. Assessing the contribution of free hydroxyl radical in organic mattersensitized photohydroxylation reactions. Environmental Science & Technology 45 (7), 2818e2825. Schlumpf, M., Schmid, P., Durrer, S., Conscience, M., Maerkel, K., Henseler, M., Gruetter, M., Herzog, I., Reolon, S., Ceccatelli, R., Faass, O., Stutz, E., Jarry, H., Wuttke, W., Lichtensteiger, W., 2004. Endocrine activity and developmental toxicity of cosmetic UV lters e an update. Toxicology 205 (1e2), 113e122. Schlumpf, M., Durrer, S., Faass, O., Ehnes, C., Fuetsch, M., Gaille, C., Henseler, M., Hofkamp, L., Maerkel, K., Reolon, S., Timms, B., Tresguerres, J.A.F., Lichtensteiger, W., 2008. Developmental toxicity of UV lters and environmental exposure: a review. International Journal of Andrology 31 (2), 144e150. Shaw, A.A., Wainschel, L.A., Shetlar, M.D., 1992. The photochemistry of p-aminobenzoic acid. Photochem. Photobiology 55 (5), 647e656. Stevenson, C., Davies, R.J.H., 1999. Photosensitization of guaninespecic DNA damage by 2-phenylbenzimidazole and the sunscreen agent 2-phenylbenzimidazole-5-sulfonic acid. Chemical Research in Toxicology 12 (1), 38e45. Suzuki, T., Osanai, Y., Isozaki, T., 2012. Effect of basic amino acids on photoreaction of ketoprofen in phosphate buffer solution. Photochemistry and Photobiology 88, 884e888. Swan, S., Schlumpf, M., Soeborg, T., Schlumpf, M., Soto, A., Schlumpf, M., Soto, A., Schlumpf, M., Foster, P., Schlumpf, M., Jensen, A., Soto, A., Leffers, H., Gregorasczczuk, E., Bjerregaard, P., Swan, S., Schlumpf, M., Patisaul, H., 2008. Developmental toxicity of UV lters and environmental exposure: a review e panel discussion. International Journal of Andrology 31 (2), 150e151. Tomasi, J., Mennucci, B., Cammi, R., 2005. Quantum mechanical continuum solvation models. Chemistry Reviews 105 (8), 2999e3093. Velosa, A.C., Baader, W.J., Stevani, C.V., Mano, C.M., Bechara, E.J.H., 2007. 1,3-diene probes for detection of triplet carbonyls in biological systems. Chemical Research in Toxicology 20 (8), 1162e1169. Wenk, J., von Gunten, U., Canonica, S., 2011. Effect of dissolved organic matter on the transformation of contaminants induced by excited triplet states and the hydroxyl radical. Environmental Science & Technology 45 (4), 1334e1340. Yu, Y., Wu, L.S., 2012. Analysis of endocrine disrupting compounds, pharmaceuticals and personal care products in

162

w a t e r r e s e a r c h 4 7 ( 2 0 1 3 ) 1 5 3 e1 6 2

sewage sludge by gas chromatography-mass spectrometry. Talanta 89, 258e263. Zepp, R.G., Schlotzhauer, P.F., Sink, R.M., 1985. Photosensitized transformations involving electronic-energy transfer in natural-waters-role of humic substances. Environmental Science & Technology 19 (1), 74e81. Zhan, M.J., Yang, X., Xian, Q.M., Kong, L.G., 2006. Photosensitized degradation of bisphenol A involving reactive oxygen species

in the presence of humic substances. Chemosphere 63 (3), 378e386. Zhang, S.Y., Chen, J.W., Qiao, X.L., Ge, L.K., Cai, X.Y., Na, G.S., 2010. Quantum chemical investigation and experimental verication on the aquatic photochemistry of the sunscreen 2-phenylbenzimidazole5-sulfonic acid. Environmental Science & Technology 44 (19), 7484e7490.

You might also like