You are on page 1of 11

Article

pubs.acs.org/est

UV/H2O2 and UV/PDS Treatment of Trimethoprim and


Sulfamethoxazole in Synthetic Human Urine: Transformation
Products and Toxicity
Ruochun Zhang,† Yongkui Yang,† Ching-Hua Huang,‡ Na Li,§ Hang Liu,† Lin Zhao,*,† and Peizhe Sun*,‡

School of Environmental Science and Engineering, Tianjin University, Tianjin 300072, China

School of Civil and Environmental Engineering, Georgia Institute of Technology, Atlanta, Georgia 30332, United States
§
Tianjin Institute of Agriculture Quality Standards and Testing Technology, Tianjin 300381, China
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Elimination of pharmaceuticals in source-sepa-


Downloaded via NATL TAIWAN UNIV on June 20, 2021 at 10:37:38 (UTC).

rated human urine is a promising approach to minimize the


pharmaceuticals in the environment. Although the degradation
kinetics of pharmaceuticals by UV/H2O2 and UV/peroxydisul-
fate (PDS) processes has been investigated in synthetic fresh and
hydrolyzed urine, comprehensive evaluation of the advanced
oxidation processes (AOPs), such as product identification and
toxicity testing, has not yet been performed. This study identified
the transformation products of two commonly used antibiotics,
trimethoprim (TMP) and sulfamethoxazole (SMX), by UV/
H2O2 and UV/PDS in synthetic urine matrices. The effects of
reactive species, including •OH, SO4•−, CO3•−, and reactive
nitrogen species, on product generation were investigated.
Multiple isomeric transformation products of TMP and SMX
were observed, especially in the reaction with hydroxyl radical. SO4•− and CO3•− reacted with pharmaceuticals by electron
transfer, thus producing similar major products. The main reactive species deduced on the basis of product generation are in
good agreement with kinetic simulation of the advanced oxidation processes. A strain identified as a polyphosphate-accumulating
organism was used to investigate the antimicrobial activity of the pharmaceuticals and their products. No antimicrobial property
was detected for the transformation products of either TMP or SMX. Acute toxicity employing luminescent bacterium Vibrio
qinghaiensis indicated 20−40% higher inhibitory effect of TMP and SMX after treatment. Ecotoxicity was estimated by
quantitative structure−activity relationship analysis using ECOSAR.

■ INTRODUCTION
Pharmaceutical micropollutants in the environment have been
which involved only physical separation of the pharmaceuticals
and further treatment is still needed. Ozone has been investigated
regarded as a major threat to the ecosystem due to their adverse for destruction of pharmaceuticals in urine, but high doses of
biological effect and the potential of inducing drug-resistant ozone (150 mg·L−1 or higher) were needed to reach satisfactory
bacteria. Because pharmaceuticals discharged in wastewater reduction.14 Therefore, more effective processes are needed to
cannot be effectively removed by traditional wastewater fully destruct the pharmaceuticals in urine.
treatment technologies, this has resulted in widespread Advanced oxidation processes (AOPs) are among the most
occurrence of pharmaceuticals in the aquatic environment effective and promising processes in eliminating organic
worldwide.1−3 Among all the wastewater streams entering into pollutants. Radical-based AOPs, in particular with hydroxyl
wastewater treatment plants (WWTPs), the treatment of source- radical (•OH)15−17 and sulfate radical (SO4•−),18−22 have been
separated urine has been viewed as a promising approach to successfully applied to destruct micropollutants in wastewater,
minimize the harm of the pharmaceuticals, because urine not drinking water, groundwater, and sediment−slurry system. •OH
only contributes a significant portion of the pharmaceuticals in (E° (•OH/H2O) = 1.9−2.7 V)23 is of low selectivity and can
municipal WWTPs,4,5 it also carries high toxic potential to react with many pharmaceuticals rapidly (k = 108−1010 M−1·
aquatic organisms.6−8 Moreover, recovering nutrients from s−1).24 SO4•− is also a strong oxidant (E° (SO4•−/SO42−) = 2.5−
urine, as a new resource recovery strategy,9 also requires
elimination of pharmaceuticals from urine. To date, nano- Received: November 13, 2015
filtration membranes,10 strong-base anion exchange polymer Revised: January 22, 2016
resins,11 electrodialysis,12 and struvite precipitation13 have been Accepted: February 3, 2016
investigated for removing pharmaceuticals from urine, all of Published: February 3, 2016

© 2016 American Chemical Society 2573 DOI: 10.1021/acs.est.5b05604


Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

3.1 V)25 and reacts with pharmaceuticals mainly through electron have been investigated and samples after AOP treatments are
transfer. In urine matrices, which are more complicated and normally more biocompatible.75−77
uniquely challenging compared to typical municipal wastewaters The objective of this study was to better understand the UV/
and surface waters because of the high concentrations of the H2O2 and UV/PDS processes in eliminating pharmaceuticals in
organic components and inorganic salts, the efficacy of UV/ synthetic source-separated urine solutions with emphasis on the
H2O2 and UV/peroxydisulfate (PDS) processes in eliminating product identification and toxicity evaluation. TMP and SMX
pharmaceuticals and metabolites was evaluated in our previous were selected because they are widely used antibiotics and have
research.26 Sulfamethoxazole (SMX), acetyl-sulfamethoxazole been shown high reactivity toward different reactive species in
(acetyl-SMX) (a metabolite of SMX), and trimethoprim (TMP) AOPs in our previous study.26 In the present study, products
were studied due to their high frequency of detection in the generated by specific radicals were identified, and degradation of
environment. Systematic kinetic studies have elucidated the TMP and SMX by mixed reactive species was illustrated in
direct photolysis and indirect photolysis of the above synthetic urine solutions. Toxic effects of the parent compounds
pharmaceuticals and metabolites’ degradation rates. In addition and the transformation products were evaluated by testing their
to •OH and SO4•−, carbonate radical (CO3•−) and reactive antimicrobial property and bioluminescence inhibition. Quanti-
nitrogen species (RNS) also showed reactivity toward the tative structure−activity relationship (QSAR) analysis was also
pharmaceuticals and acted as the main reactive species under applied to estimate the ecotoxicity. The major goal was to achieve
particular conditions. CO3•− is electrophilic and also sufficient comprehensive evaluation of AOPs treatment of pharmaceuticals
for degrading pharmaceuticals (E° (CO3•−/CO32−) = 1.63 V at in urine matrices, which can help develop better urine treatment
pH 8.4).27 Reactivity of CO3•− toward pharmaceuticals has been strategies. Research on reacting mechanism by different reactive
reported but is still in its early stage.28,29 RNS, such as amino species and corresponding toxic effects would improve the
radical, nitrogen dioxide radical, nitric oxide radical, and understanding of AOPs performances.
peroxynitrite, etc., are weaker oxidants with relatively high
selectivity.26 However, kinetics studies alone were not sufficient
to evaluate the performance of AOP processes.
■ MATERIALS AND METHODS
Materials. Sources of chemicals and materials are provided in
Although AOPs are powerful oxidation processes, often they the Supporting Information (SI) Text S1.
cannot completely mineralize pollutants within the typical Experimental Setup. UV, UV/H2O2, and UV/PDS experi-
operation time and a number of products are formed and may ments were conducted in a reactor with the same setup as in the
persist in the system.30 The product formation may vary with previous research.26 Potassium ferrioxalate was employed as
different treatment processes, oxidants, background constituents chemical actinometer78 and UV fluence rate was measured at
involved, and reaction mechanisms.31−34 Product identification 1.78 × 10−6 Einstein·L−1·s−1 in this study. Reaction solutions
can provide crucial clues for reaction pathways.35,36 For TMP were prepared with 100 μM TMP or SMX in 100 mL of fresh
and SMX, products generated by chlorination,37−40 ozona- urine, hydrolyzed urine, or phosphate buffer (PB) solution at pH
tion,30,41,42 photochemical reactions,43−45 photocatalytic treat- 6 or 9. The synthetic human urine composition was adapted from
ment,46−48 electrochemical processes,49 and in other oxidation the previous recipe (SI Table S1).26 The high initial
systems50,51 were identified and utilized to propose correspond- concentration of TMP and SMX was chosen in order to explore
ing reaction pathways. However, to our knowledge, products of transformation products. H2O2 and potassium PDS were added
SMX by UV/PDS and TMP by UV/H2O2 and UV/PDS have at 3 mM in PB solutions to generate •OH- and SO4•−-dominant
not been reported. Information is also scarce regarding oxidation systems, respectively. To create CO3•−- and RNS-dominant
products of pharmaceuticals by CO3•− and RNS. Product systems, 0.5 M sodium bicarbonate and 1 M ammonia solution
analysis coupled with measurement such as total organic carbon were added into PB solution at pH 9, respectively, in UV/H2O2
(TOC) can be complementary to assess the overall efficiency of system. Simulation of radical concentrations under different
the treatment processes. experimental conditions was conducted by Gepasi 3.0. The rate
In recent years, evaluation of toxicological effect has been constants for reactions considered in the simulation were
recommended52 and applied to investigate the biological impact obtained from our previous publication, which considered all
of pharmaceuticals in various environmental matrices.53−57 relevant reactions under AOP conditions. This model has been
Transformation products may retain the properties of the parent successfully employed in a number of studies and proved to be
compounds or potentially become more biologically active. reliable to predict radical concentrations.26,79 Total simulation
Therefore, toxicity tests of the products have been increasingly time was set as 300 s because most reactive species were close to
applied as part of the evaluation of the performance of treatment the pseudosteady state. Small amounts of concentrated NaOH
processes.58,59 At least two or multiple bioassays are commonly and perchloric acid were added to adjust the pH when needed.
needed to reveal the overall toxic potential.60,61 Cytotoxicity,62 Analytical Methods. A Waters AcQuity UPLC system
genotoxicity,63 and estrogenicity64 have been applied to test the equipped with a PDA detector and a BEH C18 column (2.1 × 50
toxic effect of the pharmaceuticals and transformation mm, 1.7 μM) was used to monitor the loss of the parent
products.58 Luminescent bacteria, particularly Vibrio f i- compounds. The UPLC system was connected to a TOF mass
scheri,65−67 are frequently employed to assess acute toxicity. spectrometer (Premier, Micromass, UK) with electrospray
Crustaceans, algae,41 fish,68 and zebrafish embryos69 are also interface to analyze transformation products. Structure identi-
widely used as toxicity indicators. Vibrio qinghaiensis, a fresh fication was achieved based on the fragmentation pattern. The
water luminescent bacterium, is now gaining increasing popular- elemental composition was deduced from the exact m/z values
ity.65,70 As for antibiotics, the antibacterial property of the obtained from the ESI-TOF-MS system (micrOTOF-Q II,
compounds is commonly monitored using bacterial strains of E. Bruker, Germany). The detailed chromatographic and MS
coli71,72 and Bacillus subtilis,73,74 while few studies have applied conditions are summarized in SI Text S2.
functional bacteria as indicators. Biodegradability of pharma- Toxicity Analysis. The antimicrobial properties of TMP and
ceuticals and their transformation products in activated sludge SMX and their transformation products were tested using a
2574 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

polyphosphate-accumulating bacterium identified as Aeromonas.


Optical density (absorbance at 600 nm) was used as an indication
of bacterial growth. The acute toxicity assay was carried out
against freshwater luminescent bacterium Vibrio qinghaiensis.
Because ammonia and bicarbonate show high toxicity to Vibrio
qinghaiensis, the acute toxicity of products by CO3•− and RNS
were not tested. QSAR analysis calculated by the Ecological
Structure−Activity Relationship Model (ECOSAR) program30
was also employed to estimate the acute and chronic toxicity for
fish, daphnid, and green algae. Detailed information for toxicity
analysis is described in SI Text S3.

■ RESULTS AND DISCUSSION


Contribution of Reactive Species. With UV/H2O2 and
UV/PDS processes in synthetic urine solutions, •OH, SO4•−,
CO3•−, and RNS were major reactive species to degrade TMP
and SMX.26 To investigate the products generated by each
reactive species, experimental conditions where certain reactive
species dominated were designed. On the basis of the simulation
results, the concentrations of above-mentioned reactive species
under UV/H2O2 and UV/PDS conditions with different
components in solutions are presented in SI Table S2. The
dominant reactive species generated by UV/H2O2 and UV/PDS
in PB solution are •OH and SO4•−, respectively, for both pH 6
and pH 9. Although •OH can react with SO4•− and generate
•OH so that •OH concentration would increase when pH was
higher, it was still not sufficient (1.21 × 10−15 M) for an
observable degradation for TMP. More than 99% of TMP was
degraded by SO4•−. By adding excess sodium bicarbonate in UV/
H2O2 system, the CO3•− concentration was 4 orders of
magnitude higher than that of •OH. On the basis of the
second-order rate constants of TMP and SMX with •OH and
CO3•−,26 more than 90% of TMP and SMX was destructed by
Figure 1. Degradation of (a) TMP and (b) SMX in UV, UV/H2O2, and
CO3•− in indirect photolysis. Because the second-order rate UV/PDS systems at pH 6 and 9 in 5 mM PB solutions, and UV/H2O2
constants of pharmaceuticals with RNS were not available, RNS systems with 0.5 M NaHCO3 or 1 M NH4OH. Oxidant concentration
contribution was not possible to evaluate quantitatively. But was 3 mM. UV fluence = UV fluence rate × exposure time.
according to the simulation result shown in SI Table S2, when
excess ammonia solution was added in UV/H2O2 system, the
•OH concentration was 3.72 × 10−15 M which was insufficient to Therefore, evaluation of the product abundance was on the basis
yield an observable degradation of TMP and SMX. Meanwhile, of peak area (shown in SI).
concentrations of several RNS were relatively high thus they may Transformation Products of TMP by Hydroxyl Radical. The
play a role in degrading pharmaceuticals with higher reactivity. degradation of TMP by hydroxyl radical (i.e., under UV/H2O2
Therefore, RNS was assumed to be the dominant reactive species condition, SI Table S2) produced 13 main products at pH 6 in
under this condition. PB solution (Figure 2). Hydroxylation was the most prominent
Identification of Transformation Products of TMP. On mechanism, accompanied by demethylation and carbonylation.
the basis of the previous results,26 direct photolysis of TMP TP 307-3 (m/z 307, C14H19N4O4) was the most abundant
under low-pressure UV irradiation was negligible. Under AOP product based on the peak area (Figure 2, SI Figure S2a). The
conditions, TMP was mainly destructed by •OH, SO4•−, and addition of 16 Da to the molecular weight of the parent
CO3•−. Therefore, products produced by these radicals were compound suggested a transformation pathway of hydroxylation.
identified. Degradations of TMP by different radicals are shown The position of the hydroxyl group was proposed (Figure 2) due
in Figure 1a. The accurate m/z values obtained by the ESI-TOF- to the presence of the fragment ions m/z 277, m/z 259, and m/z
MS system are shown in SI Table S3. On the basis of accurate 123 (SI Figure S3), which were in accordance with the
mass, empirical chemical formulas of each transformation fragmentation pattern reported by Sirtori et al.80 TP 307-3 was
product were proposed (SI Table S3). In all cases, no cleavage likely produced by direct •OH addition to the benzene moiety.81
of the methylene group or aromatic rings was observed (Figure Two more m/z 307 products, TP 307-1 and TP 307-2, were also
2). This suggested inadequate mineralization of oxidation detected at much lower concentrations. TP 307-1 was a major
processes. Indeed, the decrease of TOC was negligible under product by SO4•− and will be discussed in a later section.
all tested conditions within corresponding reaction time (data TP 325 (m/z 325, C14H21N4O5) was another major product
not shown). Because standards of the transformation products with peak area similar to that of TP 307-3 (Figure 2, SI Figure
are not commercially available, accurate quantification of each S2a). The fragment ions m/z 221 and m/z 143 suggested that
product is impossible to achieve. On the basis of the structural one oxygen atom was added to C8 atom as a hydroxyl group and
similarity of the parent compound and the transformation one oxygen atom was added to C13 atom (SI Figure S4).
products, the MS signal responses were assumed to be similar. Previous research with ozone and nitrifying activated sludge
2575 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

Figure 2. Proposed structures of transformation products of TMP under different conditions. A: UV/H2O2 at pH 6 (•OH-dominant); B: UV/H2O2 at
pH 9 (•OH-dominant); C: UV/PDS at pH 6 (SO4•−-dominant); D: UV/PDS at pH 9 (SO4•−-dominant); E: UV/H2O2 with 0.5 M NaHCO3 (CO3•−-
dominant). Conditions A−D were conducted in 5 mM PB solutions.

processes proposed addition of carbonyl group on C13 atom condition, SI Table S2) produced 6 main products at pH 6 in PB
with hydrogenation of C9−C10 double bond.82,83 In contrast, solution (Figure 2). TP 307-1 (m/z 307, C14H19N4O4) (Figure 2,
based on our MS2 spectrum of TP 325, there was no strong SI Figure S7a) was a major product, which was an isomer of the
evidence to confirm that the oxygen was added as a carbonyl most prominent product produced by •OH (TP 307-3). The
group. In addition, in a strong oxidation process, hydrogenation fragment ions, m/z 289 and m/z 274, implied that a hydroxyl
of a double bond was difficult to achieve. Therefore, instead of group was added to the bridge methylene group (SI Figure S8).
forming a carbonyl group, the oxygen was more likely added as a The same MS 2 spectrum was also found in previous
hydroxyl group. The same hypothesis also applies to TP 341 (m/ studies.30,80,83 TP 305 (m/z 305, C14H17N4O4) was another
z 341, C14H21N4O6), with a hydroxyl group addition to TP 325 major product by SO4•− (Figure 2, SI Figure S7a). It was
on the benzene ring (implied by the fragment ion m/z 197 previously identified as a product under solar-Fenton46 and TiO2
instead of m/z 181, SI Figure S5). Further study is still needed for photocatalysis80 conditions. Both TP 307-1 and TP 305 were
structural confirmation. likely generated via electron transfer mechanism (SI Figure S9).
Proposed structures of other products are shown in Figure 2. The aromatic moieties (i.e., diaminopyrimidine or benzene) on
Structural identification is discussed in SI Text S4. Overall, all the TMP lose one electron to SO4•− forming a radical intermediate
products by •OH were not accumulated in the solution and with a positive charge. The nonpaired electron is then stabilized
likely degradable by •OH (SI Figure S2a). on C7 atom, forming a carbon-center radical, which is known to
At pH 9, TP 307-3 (m/z 307, C14H19N4O4) was also the major transform to superoxide with the presence of dissolved oxygen.
product in the PB solution, whereas the amount of TP 325 (m/z Through bimolecular interaction, the superoxide intermediate
325, C14H21N4O5) was less than that at pH 6 (SI Figure S2b). In transforms to a hydroxyl moiety yielding TP 307-1 and a
addition, all the m/z 323 products were not observed at pH 9. carbonyl moiety yielding TP 305. TP 307-1 and TP 305 were also
Overall, most products of TMP at pH 9 were of lower relative found in the •OH-dominant system, but at much lower
abundance and generated via hydroxylation with addition of a abundance. TP 323-1,-2,-4 and TP 325 were also generated by
single hydroxyl group. The difference in product speciation at pH the reaction with SO4•−, whereas other products found in the
6 and pH 9 was likely due to the reactivity of the products of •OH-dominant system (i.e., TP 277-1, TP 277-2, TP 307-2, TP
TMP. It was expected that products of TMP reacted faster with 307-3, TP 295, and TP 341) were not observed. These
hydroxyl radical at higher pH due to two possible reasons. First, differences implied different reaction pathways of SO4•− and
as reported in our previous study,26 the second-order rate •OH. Unlike the •OH products, the products produced by
constant of TMP with hydroxyl radical was higher at pH 9 than SO4•− did not degrade throughout the reaction, indicating they
that at pH 6. Because the majority of productsresulted from slight may be persistent in UV/PDS process (SI Figure S7a).
modification of parent TMP, it was expected the product also has In the UV/PDS process at pH 9 instead of pH 6, the m/z 323
the same trend. Second, hydroxylated TMPs (such as TP 323) products were hardly detected, similar to that observed in UV/
were likely composed of secondary −OH moiety, which is H2O2 process at the two different pHs. TP 305 overweighed TP
known to degrade faster under AOP condition at higher pH (SI 307-1 in abundance and became the dominant product at pH 9
Figure S9).84 (SI Figure S7b). As shown in SI Figure S9, TP 307-1 further
Transformation Products of TMP by Sulfate Radical. The reacted with sulfate radical and transformed to TP 305. Although
degradation of TMP by sulfate radical (i.e., under UV/PDS such transformation was slow at pH 6, this reaction was
2576 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

Figure 3. Proposed structures of transformation products of SMX under different conditions. A: UV/H2O2 at pH 6 (•OH-dominant); B: UV/H2O2 at
pH 9 (•OH-dominant); C: UV/PDS at pH 6 (SO4•−-dominant); D: UV/PDS at pH 9 (SO4•−-dominant); E: UV/H2O2 with 0.5 M NaHCO3 (CO3•−-
dominant); F: UV/H2O2 with 1 M NH3 (RNS-dominant); G: UV at pH 6; H: UV at pH 9. Conditions A−D, G, and H were conducted in 5 mM PB
solutions.

accelerated at pH 9 because the elimination of superoxide radical more than 50% TMP removal, since the degradation was
was catalyzed by basic conditions.84 inhibited to a large extent. Samples were taken at each time
Transformation Products of TMP by Carbonate Radical. interval and a spectrum that can show all the dominant products
The degradation of TMP by carbonate radical (i.e., under UV/ is shown in SI Figures S13−S16 for each case.
H2O2/NaHCO3 condition, SI Table S2) produced 4 main In terms of product identification, in fresh urine with UV/
products at pH 9 in PB solution (Figure 2). Conditions at pH 6 H2O2 process, the presence of TP 277-2 and TP 307-3 confirmed
were not considered because carbonate radical only dominated in that the oxidation by •OH played an important role (SI Figure
hydrolyzed urine (pH 9).26 TP 305 (m/z 305, C14H17N4O4) was S13). With UV/PDS process, the main product of SO4•− TP
also a dominant and stable product in the CO3•−-dominant 307-1 stayed as the most abundant product, indicating the
system (Figure 2, SI Figure S11), as by SO4•−, suggesting similar contribution of SO4•− to TMP degradation (SI Figure S14). In
electron transfer mechanism of these two radicals (SI Figure S9). hydrolyzed urine, by UV/H2O2, the presence of TP 307-1, TP
TP 307-1 (m/z 307, C14H19N4O4) was found in CO3•−- 305, and TP 277-3 proved the degradation role of CO3•− (SI
dominant system but not as significant as it was in SO4•−- Figure S15). With UV/PDS process, the abundance of TP 307-1
dominant system (at pH 9). This difference can be explained by implies that besides CO3•−, other reactive species, such as SO4•−
an additional pathway which produced TP 305. As shown in SI and RNS, may also play a role (SI Figure S16). Indeed, when
Figure S9, besides the same mechanism as SO4•−, CO3•− can CO3•− was the dominant reactive species in the system, TP 305
transfer one of its oxygen to the carbon-center radical forming a was more abundant than TP 307-1. The major reactive species
carbonyl moiety (i.e., TP 305).85 Two new products (TP 277-3 deduced from the product presence in the present study are in
and TP 291) were generated exclusively by CO3•−. TP 277-3 (m/ good agreement with the simulation results of corresponding
z 277, C13H17N4O3) was an isomer of TP 277-1 and TP 277-2 contributing reactive species in synthetic urine solution.26
which were produced by •OH. The retention time of TP 277-3 Identification of Transformation Products of SMX.
on UPLC was close to that of TMP (SI Table S3), suggesting a Degradation of SMX is profiled in Figure 1b, and accurate m/z
structure similar to the parent compound, which is also values of transformation products are shown in SI Table S4.
supported by the remarkable resemblance between the MS2 Because of the high direct photolysis rate of SMX, the majority of
fragmentation patterns of TMP and TP 277-3 (SI Figure S12). SMX was degraded via direct photolysis in the UV/H2O2 and
The fragment ions m/z 261 and m/z 247 corresponded to the UV/PDS processes. As a result, the products of indirect
loss of a methoxyl group on C2 or C4 atom. The lack of fragment photolysis of SMX were not as prominent as the case of TMP.
ion m/z 123 suggested the hydroxyl group was added to the On the basis of both experimental and simulation results,26 SMX
diaminopyrimidine ring or the methylene group. An isomer of was of higher reactivity toward radical species than TMP and can
TMP, TP 291 (m/z 291, C14H19N4O3), was observed and the be degraded by •OH, SO4•−, CO3•−, and RNS. Products of SMX
retention time was longer than all the products and parent TMP. were analyzed respectively for each reactive species. The
The different fragmentation patterns of TP 291 and TMP corresponding radical-dominant systems were generated under
indicated a large difference between the structures, but current the same conditions as for TMP product analysis. Direct
information was not sufficient to identify the structure of TP 291. photolysis products were generated in PB solutions.
Transformation Products of TMP in Synthetic Urine Transformation Products by Direct Photolysis. By direct
Solution. After gathering the information on transformation photolysis, an isomer of SMX, SP 254 (m/z 254, C10H11N3O3S)
products of TMP by individual radicals, product identification was produced as a major product (Figure 3, SI Figure S17); it was
was performed in synthetic urine solution. To identify products produced due to photoisomerization.45 SP 99-1 (m/z 99,
in urine matrices, in hydrolyzed urine, the reaction time was set C4H7N2O) was likely 3-amino-5-methylisoxazole (Figure
the same as the experiment identifying products for CO3•− which 3),41,45 which was produced due to the cleavage of the
was regarded as the dominant radical. In fresh urine, reaction sulfonamide bond. It was observed under every tested condition.
time was set longer than in buffer solution (pH = 6) to reach An isomer of SP 99-1 (SP 99-2, m/z 99, C4H7N2O) was detected
2577 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

Figure 4. Inhibition on the tested strain (identified as Aeromonas) by parent compound standard solutions and samples after treatment: (a) TMP and
samples at pH 6, (b) TMP and samples at pH 9, (c) SMX and samples at pH 6, (d) SMX and samples at pH 9. Initial concentration of TMP and SMX in
AOP samples (i.e., 0% removal) was 100 μM. The samples were diluted by a factor of 2 in the inhibition test (i.e., 50 μM at 0% removal). Lines represent
the inhibition of the remaining parent compounds, and dots represent the observed inhibition by the samples. Errors represent the standard deviation (n
= 3).

as a product only by direct photolysis at pH 9 (Figure 3, SI Figure additional product was found except for the direct photolysis
S17b), which was probably due to the cleavage of the products. However, at pH 6, the abundance of SP 254 decreased
sulfonamide bond of SP 254. dramatically compared with that by direct photolysis (SI Figure
Transformation Products of SMX by Hydroxyl Radical. In S20a). At pH 9, no isomer product was observed (SI Figure
addition to the direct photolysis products, two products with m/z S20b). Our preliminary results showed that PDS did not react
270 (SP 270, SP 270-2, C10H11N3O4S) (Figure 3, SI Figure S18) with SP 254, suggesting SP 254 or some intermediates essential
were produced in •OH-dominant system (i.e., under UV/H2O2 for forming SP 254 may be degraded by SO4•−.
conditions). The 16 Da higher molecular weight suggested Transformation Products of SMX by Carbonate Radical
hydroxylation occurred on the parent compound. The fragment and RNS. In CO3•−-dominant system (UV/H2O2/NaHCO3),
ion m/z 172, compared with the fragment ion m/z 156 of SMX, there was no product found other than the direct photolysis
indicated that a •OH was added to the benzene ring (SI Figure products. As by SO4•−, the production of SP 254 was also
S19). A product of SMX with m/z 270 was also observed in TiO2 suppressed (SI Figure S21), suggesting the presence of PDS was
photocatalysis system.47 SP 262 was found exclusively in the not the reason for the decrease of SP 254 production.
•OH-dominant system. The concentration profile of SP 262 was Considering the similarity of the reacting mechanisms of
consistent with that of SP 99-1 (SI Figure S18), suggesting a SO4•− and CO3•−, SP 254 or some intermediates essential for
correlation between the two products. Considering SP 99-1 forming SP 254 might be prone to destruction by electron
represented the isoxazole ring of SMX, the other moiety of SMX transfer mechanism.
after cleavage of the sulfonamide bond might recombine with On the basis of the kinetic simulation (SI Table S2), the
small fragments and generate SP 262. However, the speculation oxidation by •OH can be neglected when 1 M ammonia was
needs further evidence to be confirmed. All the products added into the solution with UV/H2O2 process. Therefore, in
degraded more rapidly at pH 9 than at pH 6, especially for SP 99- addition to the direct photolysis product, SP 270-2 was also
1 and SP 262, which accumulated in the solution at pH 6 (SI observed and speculated to be a product by reaction with RNS
Figure S18). (SI Figure S22). As peroxynitrite was identified as the major RNS
Transformation Products of SMX by Sulfate Radical. In in hydrolyzed urine under AOP conditions, formation of
UV/PDS process in PB solution (SO4•−-dominant system), no hydroxylation products of SMX was likely partly due to the
2578 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

Figure 5. Impact on Vibrio qinghaiensis luminescence by parent compound standard solutions and samples after treatment: (a) TMP and samples at pH
6, (b) TMP and samples at pH 9, (c) SMX and samples at pH 6, (d) SMX and samples at pH 9. Initial concentration of TMP and SMX in AOP samples
(i.e., 0% removal) was 100 μM. The samples were diluted by a factor of 2 in the inhibition test (i.e., 50 μM at 0% removal). Lines represent the inhibition
of the remaining parent compounds and dots represent the observed inhibition by the samples. Errors represent the standard deviation (n = 3).

reaction with peroxynitrite. Indeed, oxidation of phenolic Antimicrobial Property. Aeromonas has been accepted as a
compound by peroxynitrite was reported through one-electron polyphosphate accumulating organism (PAO) and the strain
oxidiation processes that do not involve free hydroxyl radicals, used in this study has been proved to remove phosphorus
which yielded hydroxylated phenolic products.86 satisfactorily. It was employed as a representative of the
Transformation Products of SMX in Synthetic Urine microorganism in the biological wastewater treatment system
Solution. In urine matrices, the reaction time was set the same to test the antimicrobial property of the parent compounds and
as the experiment identifying products for the corresponding transformation products, so that the toxic potential of the
main reactive species and spectra that can show all the dominant transformation products on wastewater treatment systems can be
delineated.
products are shown in SI Figures S23−S26. In fresh urine with
The antimicrobial property of the parent compounds was
UV/H2O2, SP 270-2 was observed, indicating •OH was measured. The inhibition was calculated from the following
contributing (SI Figure S23). SP 99 and SP 254 were also equation:
found, which was consistent with the product identification result
under UV/H2O2 condition in buffer solution. With UV/PDS, SP ⎛ OD600 (Sample) ⎞
99-1 was the only observed product, which is similar to the case Inhibition (%) = ⎜1 − ⎟ × 100
in PB solution at pH 6 (SI Figure S24). In terms of hydrolyzed ⎝ OD600 (Control) ⎠
urine, SP 254 was hardly detected, which was expected because
where OD600 (Sample) was the absorbance at 600 nm of the
SO4•− and CO3•− were the major reactive species in these
sample aliquot taken at each selected time interval from a direct
systems (SI Figures S25 and S26). photolysis or AOP reaction, and OD600 (Control) was the
Toxicity Evaluation. Growth inhibition of TMP and SMX absorbance of DI water replacing the sample.
and the transformation products on a target strain was tested to Because standards of transformation products are not
investigate the change of antimicrobial activity during degrada- commercially available, it is difficult to test the toxicity of the
tion. Bioluminescence inhibition was monitored for acute products separately. However, it is beneficial to test the inhibition
toxicity. To comprehensively evaluate the toxicity, QSAR effect of a mixture of transformation products and remaining
analysis was also performed to predict the ecotoxicity of parent compounds because they practically coexist in the AOP
individual transformation products. systems and may exert toxicity jointly. Concentrations of the
2579 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

parent compounds were measured by UPLC, and corresponding products alone was increasing because of the decrease of toxicity
inhibition ratios were calculated from the reference curve (SI from the remaining SMX (dot−dash line in Figure 5). At the last
Figure S1). Therefore, the observed inhibition of the sample time point at pH 6, the luminescence of the samples treated by
(dots in Figure 4) subtracted by the inhibition by the parent UV/H2O2 was back to the initial level, whereas the samples
compounds (lines in Figure 4) represents the inhibition by the treated by UV/PDS retained its toxicity. At pH 9, toxicity of
products. samples treated by UV/H2O2 and UV/PDS treatment increased
To be noted, because the presence of bicarbonate could with the degradation of SMX. SO4•−-produced products
promote the bacteria growth, the inhibitions by remaining parent generated approximately 20% higher toxicity than that of •OH.
compounds in CO3•− samples are shown separately (dashed Reduced flavin mononucleotide (FMNH2) is necessary for
lines in Figure 4). For almost all the data points, the observed Vibrio qinghaiensis to luminesce.89 Hydroxylated compounds are
inhibition equaled, or was slightly higher than, the inhibition of prone to combine with FMNH2 through hydrogen bond to block
the retaining parent compound calculated by the reference curve. the bonding between FMNH2 and luciferase (the most
This indicated that toward the tested strain, negligible significant catalyzer of Vibrio qinghaiensis for luminescing).90
antimicrobial property was retained for the transformation Most products of TMP were hydroxylated compounds, which
products throughout the reaction. resulted in higher acute toxicity of the treated samples. Moreover,
TMP interferes with the normal bacterial metabolism pathway products by SO4•− were more abundant and accumulated in
by binding to dihydrofolate reductase, and SMX acts by solutions, whereas products by •OH were degraded in •OH-
inhibiting bacterial utilization of para-aminobenzoic acid dominant system, which was likely a reason toxicity of UV/PDS
(PABA) because of their structural similarity with dihydrofolic treated samples was relatively higher. For SMX, the major
acid and PABA, respectively.87 After treatment, the parent transformation products (e.g., SP 254 and SP 270) retained
compounds were hydroxylated, carbonylated, demethylated, −NH2 group. In addition, due to the cleavage of the sulfonamide
isomerized, or broke down. The structurally modified products bond to form SP 99 and the rest moiety (probably aniline-3-
may not able to bind to the receptors as the parent compounds sulfonic acid), the number of the −NH2 group was elevated.
did, thus PABA and DHF were utilized without inhibition. The −NH2 group is known to interact with FMNH2, thus the
structural transformation of parent TMP and SMX by AOPs is aminated products led to higher acute toxicity against Vibrio
likely to disable the competitive antagonism and therefore qinghaiensis.
deprive the products of antibiotic properties. Although no growth inhibition toward Aeromonas was
Acute Toxicity. Microtox test has been the most frequently observed, increasing acute toxicity of products with degradation
applied method to measure the acute toxicity of toxic substances of parent compounds was observed for both TMP and SMX.
in environmental studies.88 When the target microorganism Therefore, single bioassay was not sufficient to comprehensively
contacts with toxic substances, its bioluminescence decreases due evaluate the toxicity of the products.
to disruption of normal metabolism. Although marine bacterium Ecotoxicity. To estimate the impact of the parent
Vibrio f ischeri is the most commonly used indicator of toxicity, it pharmaceuticals and transformation products on various species,
is not suitable in fresh water studies. Therefore, freshwater QSAR analysis was applied to predict the ecotoxicity by
luminescent bacterium Vibrio qinghaiensis was selected in this ECOSAR program.
study. Multiple classes were identified for TMP based on specific
Figure 5 shows the acute toxicity of the parent compounds and structure features when running ECOSAR. In previous research,
the products. The y-axis, L/L0, reflects the comparison of the 48-h Half Effective Concentration (EC50) value for D. magna91
luminescence of the sample with the initial luminescence (i.e., and 96-h Half Lethal Concentration (LC50) value for O. latipes68
before treatment at t = 0). Data points above the dashed line were reported. The class aniline (unhindered) had the closest
(standing for no change in the luminescence) indicate that the corresponding toxicity value and thus was selected for prediction.
luminescence increased after treatment, which suggests a decline Because of the structural similarity of TMP and its products, the
in toxicity. Likewise, data points below the dashed line indicate same class was selected for its products. The results are shown in
lower luminescence and higher toxicity. Dot−dash lines in Figure SI Table S5. Compounds showed different toxicity levels for
5 represent the L/L0 of the samples containing different amounts different species, in which, daphnid was the most sensitive
of parent pharmaceuticals. species for TMP and the products. For fish and green algae, LC50
Compared with antibiotic-free samples, TMP, at lower than values for most transformation products were higher than that
100 μM concentration, exhibited almost no inhibition to the for TMP. However, for daphnid, LC50 values for most products
Vibrio qinghaiensis (SI Figure S27). However, at both pH 6 and were half of the value for TMP, suggesting higher toxicity than
pH 9, the luminescence of the samples decreased to TMP (SI Table S5a). In terms of chronic toxicity, the difference
approximately 80% of the initial luminescence, indicating slight between species was reduced compared to the results of acute
toxicity of the products by •OH. For products by SO4•−, a toxicity. For daphnid and green algae, all the products exhibited
distinct inhibition was observed because 40% of the lower toxicity than TMP, whereas for fish, most products were
luminescence was suppressed, indicating higher toxicity. more toxic (SI Table S5b).
SMX inhibited approximately 25% of luminescence of Vibrio For SMX, the class aniline (unhindered) was also selected
qinghaiensis at 100 μM at neutral pH compared with antibiotic- based on its closest corresponding toxicity value with the
free samples (SI Figure S27). With the decrease of SMX, the reported values.68,91 Unlike TMP, the acute and chronic toxicity
inhibitory effects on Vibrio qinghaiensis decreased (dot−dash line for three species showed the same trend for SMX and the
in Figure 5c and d). As for the samples at pH 6 and 9, the products (SI Table S6). Except for SP 99, all the other products
luminescence of the samples after direct photolysis reduced by showed lower toxicity than SMX. Daphid was also the most
around 40% compared with the initial luminescence. The sensitive species.
observed toxicity of the samples remained almost unchanged Environmental Significance. Source-separated urine is a
with the degradation of SMX, suggesting that the toxicity of the complex matrix where different types of reactive species
2580 DOI: 10.1021/acs.est.5b05604
Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

interacted with pharmaceuticals simultaneously under UV/H2O2 (6) Mix or NoMix? A Closer Look at Urine Source Separation; Eawag
and UV/PDS conditions. By elucidation of the transformation Swiss Federal Institute of Aquatic Science and Technology, 2007.
products and mechanisms, this study demonstrated significant (7) Escher, B. I.; Bramaz, N.; Richter, M.; Lienert, J. Comparative
product variations of TMP and SMX when they were attacked by ecotoxicological hazard assessment of beta-blockers and their human
metabolites using a mode-of-action-based test battery and a QSAR
different reactive species. Especially, the transformation products
approach. Environ. Sci. Technol. 2006, 40 (23), 7402−7408.
of pharmaceuticals by carbonate radical and RNS were (8) Lienert, J.; Güdel, K.; Escher, B. I. Screening method for
investigated for the first time, which provided more insight on ecotoxicological hazard assessment of 42 pharmaceuticals considering
the radical chemistry in aqueous phase. The final products human metabolism and excretory routes. Environ. Sci. Technol. 2007, 41
detected in the synthetic urine after treatment by AOPs was able (12), 4471−4478.
to delineated by the simulation results of radical concentrations (9) Latifian, M.; Holst, O.; Liu, J. Nitrogen and Phosphorus Removal
and the transformation products generated by the dominant from Urine by Sequential Struvite Formation and Recycling Process.
radicals. Clean: Soil, Air, Water 2014, 42 (8), 1157−1161.
Toxicity evaluation showed that the UV/H2O2 and UV/PDS (10) Pronk, W.; Palmquist, H.; Biebow, M.; Boller, M. Nanofiltration
processes were able to eliminate the antibacterial properties from for the separation of pharmaceuticals from nutrients in source-separated
TMP and SMX on the functional bacteria in wastewater urine. Water Res. 2006, 40 (7), 1405−1412.
(11) Landry, K. A.; Sun, P.; Huang, C.-H.; Boyer, T. H. Ion-exchange
treatment plants, which suggests AOP treatment lowers the
selectivity of diclofenac, ibuprofen, ketoprofen, and naproxen in
impact of source-separated urine on the performance of WWTP. ureolyzed human urine. Water Res. 2015, 68, 510−521.
However, it is interesting to observe higher acute toxicity of (12) Pronk, W.; Zuleeg, S.; Lienert, J.; Escher, B.; Koller, M.; Berner,
transformation products than their parent compounds. Notably, A.; Koch, G.; Boller, M. Pilot experiments with electrodialysis and
although our previous study suggested that UV/PDS was more ozonation for the production of a fertiliser from urine. Water Sci.
favorable than UV/H2O2 for the removal of parent pharmaceut- Technol. 2007, 56 (5), 219−227.
icals in source-separated urine,26 the toxicity results in this study (13) Kemacheevakul, P.; Chuangchote, S.; Otani, S.; Matsuda, T.;
indicated higher acute toxicity of the transformation products Shimizu, Y. Phosphorus Recovery: Minimization of Amount of
generated by UV/PDS. Therefore, it is suggested that a Pharmaceuticals and Improvement of Purity in Struvite Recovered
comprehensive evaluation of both kinetics and toxic effect from Hydrolyzed Urine. Environ. Technol. 2014, 35 (23), 3011−3019.
should be considered when evaluating treatment processes for (14) Dodd, M. C.; Zuleeg, S.; von Gunten, U.; Pronk, W. Ozonation of
source-separated urine for resource recovery and waste minimization:
degrading target pollutants.


process modeling, reaction chemistry, and operational considerations.
Environ. Sci. Technol. 2008, 42 (24), 9329−9337.
ASSOCIATED CONTENT (15) Katsoyiannis, I. A.; Canonica, S.; von Gunten, U. Efficiency and
*
S Supporting Information energy requirements for the transformation of organic micropollutants
The Supporting Information is available free of charge on the by ozone, O3/H2O2 and UV/H2O2. Water Res. 2011, 45 (13), 3811−
ACS Publications website at DOI: 10.1021/acs.est.5b05604. 3822.
(16) Yao, H.; Sun, P.; Minakata, D.; Crittenden, J. C.; Huang, C.-H.
Text S1−S4, Tables S1−S6, and Figures S1−S28 (PDF) Kinetics and Modeling of Degradation of Ionophore Antibiotics by UV

■ AUTHOR INFORMATION
Corresponding Authors
and UV/H2O2. Environ. Sci. Technol. 2013, 47 (9), 4581−4589.
(17) Wols, B. A.; Hofman-Caris, C. H. M.; Harmsen, D. J. H.;
Beerendonk, E. F. Degradation of 40 selected pharmaceuticals by UV/
H2O2. Water Res. 2013, 47 (15), 5876−5888.
*Phone: 86-22-27401154; e-mail: zhaolin@tju.edu.cn. (18) Ahmed, M. M.; Brienza, M.; Goetz, V.; Chiron, S. Solar photo-
*Phone: 1-404-358-4858; e-mail: sunpeizhe@gatech.edu. Fenton using peroxymonosulfate for organic micropollutants removal
Notes from domestic wastewater: Comparison with heterogeneous TiO2
The authors declare no competing financial interest. photocatalysis. Chemosphere 2014, 117, 256−261.


(19) Ayoub, G.; Ghauch, A. Assessment of bimetallic and trimetallic
ACKNOWLEDGMENTS iron-based systems for persulfate activation: Application to sulfame-
thoxazole degradation. Chem. Eng. J. 2014, 256, 280−292.
This work was supported by a project from National Natural (20) Ji, Y.; Ferronato, C.; Salvador, A.; Yang, X.; Chovelon, J.-M.
Science Foundation of China (21276182).


Degradation of ciprofloxacin and sulfamethoxazole by ferrous-activated
persulfate: Implications for remediation of groundwater contaminated
REFERENCES by antibiotics. Sci. Total Environ. 2014, 472, 800−808.
(1) Zhang, T.; Li, B. Occurrence, Transformation, and Fate of (21) Anipsitakis, G. P.; Dionysiou, D. D. Degradation of Organic
Antibiotics in Municipal Wastewater Treatment Plants. Crit. Rev. Contaminants in Water with Sulfate Radicals Generated by the
Environ. Sci. Technol. 2011, 41 (11), 951−998. Conjunction of Peroxymonosulfate with Cobalt. Environ. Sci. Technol.
(2) Benotti, M. J.; Trenholm, R. A.; Vanderford, B. J.; Holady, J. C.; 2003, 37 (20), 4790−4797.
Stanford, B. D.; Snyder, S. A. Pharmaceuticals and Endocrine Disrupting (22) Rastogi, A.; Al-Abed, S. R.; Dionysiou, D. D. Sulfate radical-based
Compounds in U.S. Drinking Water. Environ. Sci. Technol. 2009, 43 (3), ferrous−peroxymonosulfate oxidative system for PCBs degradation in
597−603. aqueous and sediment systems. Appl. Catal., B 2009, 85 (3), 171−179.
(3) Luo, Y.; Guo, W.; Ngo, H. H.; Nghiem, L. D.; Hai, F. I.; Zhang, J.; (23) Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B.;
Liang, S.; Wang, X. C. A review on the occurrence of micropollutants in Tsang, W. Critical review of rate constants for reactions of hydrated
the aquatic environment and their fate and removal during wastewater electrons, hydrogen atoms and hydroxyl radicals. J. Phys. Chem. Ref. Data
treatment. Sci. Total Environ. 2014, 473−474 (0), 619−641. 1988, 17 (2), 513−886.
(4) Winker, M.; Tettenborn, F.; Faika, D.; Gulyas, H.; Otterpohl, R. (24) Wols, B. A.; Hofman-Caris, C. H. M. Review of photochemical
Comparison of analytical and theoretical pharmaceutical concentrations reaction constants of organic micropollutants required for UV advanced
in human urine in Germany. Water Res. 2008, 42 (14), 3633−3640. oxidation processes in water. Water Res. 2012, 46 (9), 2815−2827.
(5) Winker, M.; Faika, D.; Gulyas, H.; Otterpohl, R. A comparison of (25) Neta, P.; Huie, R. E.; Ross, A. B. Rate constants for reactions of
human pharmaceutical concentrations in raw municipal wastewater and inorganic radicals in aqueous solution. J. Phys. Chem. Ref. Data 1988, 17
yellowwater. Sci. Total Environ. 2008, 399 (1), 96−104. (3), 1027−1284.

2581 DOI: 10.1021/acs.est.5b05604


Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

(26) Zhang, R.; Sun, P.; Boyer, T. H.; Zhao, L.; Huang, C.-H. media: persistence, toxicity and photoproducts assessment. Chemo-
Degradation of Pharmaceuticals and Metabolite in Synthetic Human sphere 2009, 77 (10), 1292−1298.
Urine by UV, UV/H2O2, and UV/PDS. Environ. Sci. Technol. 2015, 49 (46) Michael, I.; Hapeshi, E.; Osorio, V.; Perez, S.; Petrovic, M.;
(5), 3056−3066. Zapata, A.; Malato, S.; Barcelo, D.; Fatta-Kassinos, D. Solar photo-
(27) Zuo, Z.; Cai, Z.; Katsumura, Y.; Chitose, N.; Muroya, Y. catalytic treatment of trimethoprim in four environmental matrices at a
Reinvestigation of the acid−base equilibrium of the (bi) carbonate pilot scale: Transformation products and ecotoxicity evaluation. Sci.
radical and pH dependence of its reactivity with inorganic reactants. Total Environ. 2012, 430 (14), 167−173.
Radiat. Phys. Chem. 1999, 55 (1), 15−23. (47) Hu, L.; Flanders, P. M.; Miller, P. L.; Strathmann, T. J. Oxidation
(28) Liu, Y.; He, X.; Duan, X.; Fu, Y.; Dionysiou, D. D. Photochemical of sulfamethoxazole and related antimicrobial agents by TiO2
degradation of oxytetracycline: Influence of pH and role of carbonate photocatalysis. Water Res. 2007, 41 (12), 2612−2626.
radical. Chem. Eng. J. 2015, 276, 113−121. (48) Ding, S.; Niu, J.; Bao, Y.; Hu, L. Evidence of superoxide radical
(29) Wols, B. A.; Harmsen, D. J. H.; Beerendonk, E. F.; Hofman-Caris, contribution to demineralization of sulfamethoxazole by visible-light-
C. H. M. Predicting pharmaceutical degradation by UV (MP)/H2O2 driven Bi2O3/ Bi2O2CO3/ Sr6Bi2O9 photocatalyst. J. Hazard. Mater.
processes: A kinetic model. Chem. Eng. J. 2015, 263, 336−345. 2013, 262 (22), 812−818.
(30) Kuang, J.; Huang, J.; Wang, B.; Cao, Q.; Deng, S.; Yu, G. (49) Moreira, F. C.; Garcia-Segura, S.; Boaventura, R. A. R.; Brillas, E.;
Ozonation of trimethoprim in aqueous solution: Identification of Vilar, V. J. P. Degradation of the antibiotic trimethoprim by
reaction products and their toxicity. Water Res. 2013, 47 (8), 2863− electrochemical advanced oxidation processes using a carbon-PTFE
2872. air-diffusion cathode and a boron-doped diamond or platinum anode.
(31) Lam, M. W.; Mabury, S. A. Photodegradation of the Appl. Catal., B 2014, 160−161 (6), 492−505.
pharmaceuticals atorvastatin, carbamazepine, levofloxacin, and sulfame- (50) Mahdi Ahmed, M.; Barbati, S.; Doumenq, P.; Chiron, S. Sulfate
thoxazole in natural waters. Aquat. Sci. 2005, 67 (2), 177−188. radical anion oxidation of diclofenac and sulfamethoxazole for water
(32) Ahmed, M. M.; Chiron, S. Solar photo-Fenton like using decontamination. Chem. Eng. J. 2012, 197 (14), 440−447.
persulphate for carbamazepine removal from domestic wastewater. (51) Anquandah, G. A. K.; Sharma, V. K.; Knight, D. A.; Batchu, S. R.;
Water Res. 2014, 48 (1), 229−236. Gardinali, P. R. Oxidation of Trimethoprim by Ferrate(VI): Kinetics,
(33) Zhou, Z.; Jiang, J.-Q. Treatment of selected pharmaceuticals by Products, and Antibacterial Activity. Environ. Sci. Technol. 2011, 45 (24),
ferrate(VI): Performance, kinetic studies and identification of oxidation 10575−10581.
products. J. Pharm. Biomed. Anal. 2015, 106, 37−45. (52) Agerstrand, M.; Berg, C.; Bjorlenius, B.; Breitholtz, M.;
(34) Anipsitakis, G. P.; Dionysiou, D. D.; Gonzalez, M. A. Cobalt- Brunstrom, B.; Fick, J.; Gunnarsson, L.; Larsson, D. G. J.; Sumpter, J.
mediated activation of peroxymonosulfate and sulfate radical attack on P.; Tysklind, M.; Ruden, C. Improving environmental risk assessment of
phenolic compounds. implications of chloride ions. Environ. Sci. Technol. human pharmaceuticals. Environ. Sci. Technol. 2015, 49 (9), 5336−5345.
(53) Pino, M. R.; Val, J.; Mainar, A. M.; Zuriaga, E.; Espanol, C.; Langa,
2006, 40 (3), 1000−1007.
E. Acute toxicological effects on the earthworm Eisenia fetida of 18
(35) Sun, P.; Pavlostathis, S. G.; Huang, C.-H. Photodegradation of
common pharmaceuticals in artificial soil. Sci. Total Environ. 2015, 518,
Veterinary Ionophore Antibiotics under UV and Solar Irradiation.
225−237.
Environ. Sci. Technol. 2014, 48 (22), 13188−13196.
(54) van der Grinten, E.; Pikkemaat, M. G.; van den Brandhof, E.-J.;
(36) Zong, W.; Sun, F.; Sun, X. Oxidation by-products formation of
Stroomberg, G. J.; Kraak, M. H. S. Comparing the sensitivity of algal,
microcystin-LR exposed to UV/H2O2: Toward the generative
cyanobacterial and bacterial bioassays to different groups of antibiotics.
mechanism and biological toxicity. Water Res. 2013, 47 (9), 3211−3219.
Chemosphere 2010, 80 (1), 1−6.
(37) Dodd, M. C.; Huang, C.-H. Aqueous chlorination of the
(55) la Farre, M.; Perez, S.; Kantiani, L.; Barcelo, D. Fate and toxicity of
antibacterial agent trimethoprim: reaction kinetics and pathways. Water emerging pollutants, their metabolites and transformation products in
Res. 2007, 41 (3), 647−655. the aquatic environment. TrAC, Trends Anal. Chem. 2008, 27 (11),
(38) Wang, P.; He, Y.-L.; Huang, C.-H. Oxidation of Antibiotic Agent 991−1007.
Trimethoprim by Chlorine Dioxide: Reaction Kinetics and Pathways. J. (56) Dalzell, D. J. B.; Alte, S.; Aspichueta, E.; de la Sota, A.; Etxebarria,
Environ. Eng. 2012, 138 (3), 360−366. J.; Gutierrez, M.; Hoffmann, C. C.; Sales, D.; Obst, U.; Christofi, N. A
(39) Gao, S.; Zhao, Z.; Xu, Y.; Tian, J.; Qi, H.; Lin, W.; Cui, F. comparison of five rapid direct toxicity assessment methods to
Oxidation of sulfamethoxazole (SMX) by chlorine, ozone and determine toxicity of pollutants to activated sludge. Chemosphere
permanganate-A comparative study. J. Hazard. Mater. 2014, 274 (12), 2002, 47 (5), 535−545.
258−269. (57) Jesus Garcia-Galan, M.; Gonzalez Blanco, S.; Lopez Roldan, R.;
(40) Dodd, M. C.; Huang, C. H. Transformation of the antibacterial Diaz-Cruz, S.; Barcelo, D. Ecotoxicity evaluation and removal of
agent sulfamethoxazole in reactions with chlorine: kinetics, mechanisms, sulfonamides and their acetylated metabolites during conventional
and pathways. Environ. Sci. Technol. 2004, 38 (21), 5607−15. wastewater treatment. Sci. Total Environ. 2012, 437, 403−412.
(41) del Mar Gomez-Ramos, M.; Mezcua, M.; Agueera, A.; Fernandez- (58) Fatta-Kassinos, D.; Vasquez, M. I.; Kuemmerer, K. Trans-
Alba, A. R.; Gonzalo, S.; Rodriguez, A.; Rosal, R. Chemical and formation products of pharmaceuticals in surface waters and wastewater
toxicological evolution of the antibiotic sulfamethoxazole under ozone formed during photolysis and advanced oxidation processes -
treatment in water solution. J. Hazard. Mater. 2011, 192 (1), 18−25. Degradation, elucidation of byproducts and assessment of their
(42) Abellan, M. N.; Gebhardt, W.; Schroeder, H. F. Detection and biological potency. Chemosphere 2011, 85 (5), 693−709.
identification of degradation products of sulfamethoxazole by means of (59) Molkenthin, M.; Olmez-Hanci, T.; Jekel, M. R.; Arslan-Alaton, I.
LC/MS and -MS(n) after ozone treatment. Water Sci. Technol. 2008, 58 Photo-Fenton-like treatment of BPA: Effect of UV light source and
(9), 1803−1812. water matrix on toxicity and transformation products. Water Res. 2013,
(43) Luo, X.; Zheng, Z.; Greaves, J.; Cooper, W. J.; Song, W. 47 (14), 5052−5064.
Trimethoprim: Kinetic and mechanistic considerations in photo- (60) Marciocha, D.; Kalka, J.; Turek-Szytow, J.; Wiszniowski, J.;
chemical environmental fate and AOP treatment. Water Res. 2012, 46 Surmacz-Gorska, J. Oxidation of sulfamethoxazole by UVA radiation
(4), 1327−1336. and modified Fenton reagent: toxicity and biodegradability of by-
(44) Lekkerkerker-Teunissen, K.; Benotti, M. J.; Snyder, S. A.; van products. Water Sci. Technol. 2009, 60 (10), 2555−2562.
Dijk, H. C. Transformation of atrazine, carbamazepine, diclofenac and (61) Olmez-Hanci, T.; Dursun, D.; Aydin, E.; Arslan-Alaton, I.; Girit,
sulfamethoxazole by low and medium pressure UV and UV/H2O2 B.; Mita, L.; Diano, N.; Mita, D. G.; Guida, M. S2O82‑/UV-C and H2O2/
treatment. Sep. Purif. Technol. 2012, 96 (33), 33−43. UV-C treatment of Bisphenol A: Assessment of toxicity, estrogenic
(45) Trovó, A. G.; Nogueira, R. F.; Agüera, A.; Sirtori, C.; Fernández- activity, degradation products and results in real water. Chemosphere
Alba, A. R. Photodegradation of sulfamethoxazole in various aqueous 2015, 119, S115−S123.

2582 DOI: 10.1021/acs.est.5b05604


Environ. Sci. Technol. 2016, 50, 2573−2583
Environmental Science & Technology Article

(62) Richard, J.; Boergers, A.; vom Eyser, C.; Bester, K.; Tuerk, J. (80) Sirtori, C.; Agüera, A.; Gernjak, W.; Malato, S. Effect of water-
Toxicity of the micropollutants Bisphenol A, Ciprofloxacin, Metoprolol matrix composition on Trimethoprim solar photodegradation kinetics
and Sulfamethoxazole in water samples before and after the oxidative and pathways. Water Res. 2010, 44 (9), 2735−2744.
treatment. Int. J. Hyg. Environ. Health 2014, 217 (4−5), 506−514. (81) An, T.; Gao, Y.; Li, G.; Kamat, P. V.; Peller, J.; Joyce, M. V.
(63) Karci, A.; Arslan-Alaton, I.; Bekbolet, M.; Ozhan, G.; Alpertunga, Kinetics and mechanism of (*)OH mediated degradation of dimethyl
B. H2O2/UV-C and Photo-Fenton treatment of a nonylphenol phthalate in aqueous solution: experimental and theoretical studies.
polyethoxylate in synthetic freshwater: Follow-up of degradation Environ. Sci. Technol. 2014, 48 (1), 641−648.
products, acute toxicity and genotoxicity. Chem. Eng. J. 2014, 241 (4), (82) Radjenovic, J.; Godehardt, M.; Hein, A.; Farré, M.; Jekel, M.;
43−51. Barceló, D. Evidencing generation of persistent ozonation products of
(64) vom Eyser, C.; Boergers, A.; Richard, J.; Dopp, E.; Janzen, N.; antibiotics roxithromycin and trimethoprim. Environ. Sci. Technol. 2009,
Bester, K.; Tuerk, J. Chemical and toxicological evaluation of 43 (17), 6808−6815.
transformation products during advanced oxidation processes. Water (83) Eichhorn, P.; Ferguson, P. L.; Pérez, S.; Aga, D. S. Application of
Sci. Technol. 2013, 68 (9), 1976−1983. ion trap-MS with H/D exchange and QqTOF-MS in the identification
(65) Qi, C.; Liu, X.; Lin, C.; Zhang, X.; Ma, J.; Tan, H.; Ye, W. of microbial degradates of trimethoprim in nitrifying activated sludge.
Degradation of sulfamethoxazole by microwave-activated persulfate: Anal. Chem. 2005, 77 (13), 4176−4184.
Kinetics, mechanism and acute toxicity. Chem. Eng. J. 2014, 249, 6−14. (84) von Sonntag, C.; Schuchmann, H. P. The Elucidation of Peroxyl
(66) Sagi, G.; Csay, T.; Patzay, G.; Csonka, E.; Wojnarovits, L.; Takacs, Radical Reactions in Aqueous Solution with the Help of Radiation-
E. Oxidative and reductive degradation of sulfamethoxazole in aqueous Chemical Methods. Angew. Chem., Int. Ed. Engl. 1991, 30 (10), 1229−
solutions: decomposition efficiency and toxicity assessment. J. 1253.
(85) Crean, C.; Geacintov, N. E.; Shafirovich, V. Vladimir, S. Oxidation
Radioanal. Nucl. Chem. 2014, 301 (2), 475−482.
of guanine and 8-oxo-7,8-dihydroguanine by carbonate radical anions:
(67) Karci, A.; Arslan-Alaton, I.; Bekbolet, M. Advanced oxidation of a
insight from oxygen-18 labeling experiments. Angew. Chem., Int. Ed.
commercially important nonionic surfactant: Investigation of degrada-
2005, 44 (32), 5057−5060.
tion products and toxicity. J. Hazard. Mater. 2013, 263, 275−282.
(86) Pryor, W. A.; Squadrito, G. L. The chemistry of peroxynitrite: a
(68) Kim, Y.; Choi, K.; Jung, J.; Park, S.; Kim, P.-G.; Park, J. Aquatic
product from the reaction of nitric oxide with superoxide. Am. J. Physiol.
toxicity of acetaminophen, carbamazepine, cimetidine, diltiazem and six 1995, 268 (268), 699−722.
major sulfonamides, and their potential ecological risks in Korea. (87) Brogden, R.; Carmine, A.; Heel, R.; Speight, T.; Avery, G.
Environ. Int. 2007, 33 (3), 370−375. Trimethoprim: a review of its antibacterial activity, pharmacokinetics
(69) Plahuta, M.; Tisler, T.; Toman, M. J.; Pintar, A. Efficiency of and therapeutic use in urinary tract infections. Drugs 1982, 23 (6), 405−
advanced oxidation processes in lowering bisphenol A toxicity and 430.
oestrogenic activity in aqueous samples. Arh. Hig. Rada Toksikol. 2014, (88) Dirany, A.; Efremova Aaron, S.; Oturan, N.; Sires, I.; Oturan, M.
65 (1), 77−87. A.; Aaron, J. J. Study of the toxicity of sulfamethoxazole and its
(70) Zhang, Q.; Chen, J.; Dai, C.; Zhang, Y.; Zhou, X. Degradation of degradation products in water by a bioluminescence method during
carbamazepine and toxicity evaluation using the UV/persulfate process application of the electro-Fenton treatment. Anal. Bioanal. Chem. 2011,
in aqueous solution. J. Chem. Technol. Biotechnol. 2015, 90 (4), 701− 400 (2), 353−360.
708. (89) Jablonski, E.; Deluca, M. Studies of the control of luminescence in
(71) Keen, O. S.; Linden, K. G. Degradation of Antibiotic Activity Beneckea harveyi: properties of the NADH and NADPH:FMN
during UV/H2O2 Advanced Oxidation and Photolysis in Wastewater oxidoreductases. Biochemistry 1978, 17 (17), 672−8.
Effluent. Environ. Sci. Technol. 2013, 47 (22), 13020−13030. (90) Chen; Liu, F.; Duan, S. Xintian Molecular Modeling Study on the
(72) Wammer, K. H.; Lapara, T. M.; McNeill, K.; Arnold, W. A.; Three-dimensional Structure of the Luciferase Protein in Vibrio-
Swackhamer, D. L. Changes in antibacterial activity of triclosan and sulfa qinghaiensis sp.-Q67. Huaxue Xuebao 2013, 71 (7), 1035−1040.
drugs due to photochemical transformations. Environ. Toxicol. Chem. (91) Halling-Sørensen, B.; Lützhøft, H.-C. H.; Andersen, H. R.;
2006, 25 (6), 1480−1486. Ingerslev, F. Environmental risk assessment of antibiotics: comparison
(73) Sun, P.; Yao, H.; Minakata, D.; Crittenden, J. C.; Pavlostathis, S. of mecillinam, trimethoprim and ciprofloxacin. J. Antimicrob. Chemoth.
G.; Huang, C.-H. Acid-catalyzed transformation of ionophore veterinary 2000, 46 (s1), 53−58.
antibiotics: Reaction mechanism and product implications. Environ. Sci.
Technol. 2013, 47 (13), 6781−6789.
(74) Dodd, M. C.; Kohler, H.-P. E.; Von Gunten, U. Oxidation of
Antibacterial Compounds by Ozone and Hydroxyl Radical: Elimination
of Biological Activity during Aqueous Ozonation Processes. Environ. Sci.
Technol. 2009, 43 (7), 2498−2504.
(75) Malato, S.; Fernández-Ibáñez, P.; Maldonado, M. I.; Blanco, J.;
Gernjak, W. Decontamination and Disinfection of Water by Solar
Photocatalysis: Recent Overview and Trends. Catal. Today 2009, 147
(1), 1−59.
(76) Coelho, A. D.; Sans, C.; Aguera, A.; Gomez, M. J.; Esplugas, S.;
Dezotti, M. Effects of ozone pre-treatment on diclofenac: intermediates,
biodegradability and toxicity assessment. Sci. Total Environ. 2009, 407
(11), 3572−3578.
(77) Keen, O. S.; Love, N. G.; Aga, D. S.; Linden, K. G.
Biodegradability of iopromide products after UV/H2O2 advanced
oxidation. Chemosphere 2016, 144, 989−994.
(78) Harris, G. D.; Dean Adams, V.; Moore, W. M.; Sorensen, D. L.
Potassium ferrioxalate as chemical actinometer in ultraviolet reactors. J.
Environ. Eng. 1987, 113 (3), 612−627.
(79) Yang, Y.; Pignatello, J. J.; Ma, J.; Mitch, W. A. Comparison of
halide impacts on the efficiency of contaminant degradation by sulfate
and hydroxyl radical-based advanced oxidation processes (AOPs).
Environ. Sci. Technol. 2014, 48 (4), 2344−2351.

2583 DOI: 10.1021/acs.est.5b05604


Environ. Sci. Technol. 2016, 50, 2573−2583

You might also like