You are on page 1of 11

Water Research 225 (2022) 119122

Contents lists available at ScienceDirect

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

High-performance reductive decomposition of trichloroacetamide by the


vacuum-ultraviolet/sulfite process: Kinetics, mechanism and combined
toxicity risk
Huahan Huang a, b, Xinrui Liang a, b, Qingsong Li a, b, *, Jing Deng c, Jing Zou d, Xia Li e,
Xiaoyan Ma c, Guoxin Li a, Guoyuan Chen a
a
Water Resources and Environmental Institute, Xiamen University of Technology, Xiamen 361005, China
b
Key Laboratory of Water Resources Utilization and Protection, Xiamen city, Xiamen 361005, China
c
College of Civil Engineering and Architecture, Zhejiang University of Technology, Hangzhou 310014, China
d
College of Civil Engineering, Huaqiao University, Xiamen 361021, China
e
College of Civil Engineering, Fuzhou University, Fuzhou 350116, China

A R T I C L E I N F O A B S T R A C T

Keywords: Trichloroacetamide (TCAcAm) is among of the nitrogenous disinfection by-products (N-DBPs) with high cyto­
Advanced reduction process toxicity and genotoxicity, which is usually detected at low concentration (μg/L) in drinking water. In this study,
Hydrous electrons advanced reduction process (ARP) based on vacuum ultraviolet (VUV) was employed to eliminate TCAcAm.
Nitrogenous disinfection by-products
Compared with VUV, VUV/sulfide, and VUV/ferrous iron processes, VUV/sulfite process demonstrated excellent
Kintecus
Reductants
performance for TCAcAm decomposition, the higher removal of TCAcAm could be achieved by VUV/sulfite
process (85.6 %) than VUV direct photolysis (13.5 %) due to the production of a great number of reactive species.
The degradation of TCAcAm followed the pseudo-first-order kinetics well in VUV/sulfite process, and the pseudo-
first-order rate constant (kobs) increased with increasing sulfite concentration. Reactive species quenching ex­
periments demonstrated that e−aq, SO⋅−3 and H⋅ were involved in the degradation of TCAcAm. The in situ generated
e−aq, SO⋅−3 and HO⋅ via VUV/sulfite process were identified by electron paramagnetic resonance (EPR), and the e−aq
was proved to be the dominated species (relative contribution: 83.5 %) for TCAcAm decomposition. The second-
order rate constant of TCAcAm with e−aq was determined to be 2.41 × 1010 M− 1 s− 1 for the first time based on
competitive kinetic method. The complete TCAcAm degradation could be achieved at pH > 8.3, while TCAcAm
degradation efficiency decreased to 11.9 % at pH 5.8. TCAcAm decay could be divided into two stages: rapid
growth (sulfite dosage: 0.25-1.0 mM) and slow growth (sulfite dosage: 1.0-4.0 mM). The yield of e−aq was
controlled by sulfite dosage, and the predict yield of e−aq increased from 3.69 × 10− 14 to 2.58 × 10− 12 M with
increasing the sulfite dosage from 0.25 to 4.0 mM by Kintecus 6.80, which resulted in an increase in TCAcAm
removal. Meanwhile, the presence of dissolved oxygen (DO), chloride (Cl− ), bicarbonate (HCO−3 ) and humic acid
(HA) posed negative influence on TCAcAm decomposition to various degrees. Dichloroacetamide (DCAcAm),
trichloroacetic acid (TCAA), dichloroacetic acid (DCAA) and Cl− were identified as intermediate products,
indicated that reductive dechlorination and hydrolysis coexisted during the degradation of TCAcAm in VUV/
sulfite process.

1. Introduction to eliminate them (Sedlak and Gunten, 2011; Bao et al., 2012). However,
the undesirable disinfection by-products (DBPs) are occurred during the
Waterborne pathogens, such as viruses, protozoa, and other micro­ drinking water treatment. Nowadays, over 800 DBPs have been recog­
organisms, pose an enduring threat to human health (Chen et al., 2021; nized and identified, which include nitro­
Du et al., 2021; Tang et al., 2020). In order to reduce diseases caused by genous/chlorinated/carbonaceous/brominated DBPs (Bond et al., 2011;
waterborne pathogens, drinking water disinfection process is employed Fang et al., 2021). Among these DBPs, nitrogenous disinfection

* Corresponding author at: Water Resources and Environmental Institute, Xiamen University of Technology, Xiamen 361005, China.
E-mail address: leetsingsong@sina.com (Q. Li).

https://doi.org/10.1016/j.watres.2022.119122
Received 15 July 2022; Received in revised form 26 August 2022; Accepted 14 September 2022
Available online 14 September 2022
0043-1354/© 2022 Elsevier Ltd. All rights reserved.
H. Huang et al. Water Research 225 (2022) 119122

by-products (N-DBPs) have attracted considerable attention in view of 2. Materials and methods
their high potential ecological risk (Bond et al., 2011; Chen et al., 2018;
Krasner et al., 2006; Shah and Mitch, 2012). Trichloroacetamide 2.1. Materials
(TCAcAm), a newly recognized unregulated N-DBPs, which generally
forms in lower amount than regulated DBPs (Nihemaiti et al., 2017). All reagents were of analytical grade and used without further pu­
While relevant studies revealed that it showed two orders of magnitude rification. TCAcAm (99 %) and dichloroacetamide (DCAcAm) (98.9 %)
higher cytotoxic and genotoxic in mammalian cell assays than regulated were obtained from CATO Research Chemicals Inc (Guangzhou, China).
DBPs (Li et al., 2012; Chu et al., 2016; Richardson et al., 2007). Besides, Methyl-tert-butyl ether (MTBE) (AR, > 99.9 %) was used for extraction
TCAcAm can also induce hepatotoxicity, and even results in metabolic and purchased from Aladdin Industrial Inc (Shanghai, China). Anhy­
disfunction by disrupting the metabolic function of the gut microbiota drous sodium sulfate (Na2SO4, 99 %) was used as the enhanced
(Zhang et al., 2015; Deng et al., 2014). Thence, it is necessary to elim­ extracting agent and purchased from Aladdin Industrial Inc (Shanghai,
inate TCAcAm from drinking water to safeguard drinking water. China). Nitrate (NO−3 ) and nitrite (NO−2 ) as e−aq and H⋅ scavengers were
Pre-disinfection removal of DBP precursors and post-formation supplied by Aladdin Industrial Inc (Shanghai, China). Nitrogen gas (N2,
removal of DBPs have become the crucial control technologies for > 99.99 %) was provided by Hong Hua Industrial Gases Co. (Xiamen,
DBPs (Zhang et al., 2020). It’s reported that UV/persulfate process prior China). Sodium sulfite (Na2SO3, 99 %) was obtained from Aladdin In­
to disinfection is capable of reducing the formation of haloacetamides dustrial Inc (Shanghai, China). Sodium sulfide nonahydrate
(HAcAms), as well as other N-DBPs (Chu et al., 2015). Meanwhile, many (Na2S⋅9H2O, 98 %) and iron (II) sulfate heptahydrate (FeSO4⋅7H2O, 99
processes, such as UV/H2O2, UV/persulfate, which principle based on %) were purchased from MACKLIN (Shanghai, China). The other re­
oxidation (UV-AOPs), have been developed to tackle the post-formation agents were obtained from Sinopharm Chemical Reagent Co., Ltd
DBPs (e.g., brominated trihalomethanes, brominated acetic acids, (Shanghai, China). All solutions were prepared with Milli-Q water pu­
chloroacetonitriles) (Ling et al., 2016; Jo et al., 2011; Li et al., 2022). rified using a Milli-Q Millipore system (18 MΩ cm− 1).
However, the trichlorinated DBPs, such as TCAcAm, can be barely All experiments were conducted in a quartz cylindrical photoreactor
eliminated by UV-AOPs due to the low reactive with HO⋅ (Chuang et al., (total volume 1 L, inner diameter 100 mm, outer diameter 108 mm,
2016; Wang et al., 2009). According to the polarity characteristics of height 170 mm) (Fig. 1). Both VUV (4 W, 185 nm + 254 nm, length 106
TCAcAm (Chu et al., 2016; Luo et al., 2015), advanced reduction pro­ mm, diameter 15 mm) and UV (4 W, 254 nm, length 106 mm, diameter
cesses (ARPs) is found to be an effective technology, which can degrade 15 mm) lamps, which were purchased from Beijing Aerospace Hongda
the highly toxic halogenated organic compounds into low toxic organic Optoelectronics Technology Co. Ltd, were used for the TCAcAm
compounds and halide ions (Luo et al., 2015; Milh et al., 2021). The degradation.
most commonly ARPs are UV/sulfite (SO2− 2−
3 ), UV/dithionite (S2O4 ),
2−
UV/iodide and UV/benzoquinone, UV/sulfide (S ) and UV/ferrous
iron (Gu et al., 2017; Liu et al., 2013; Yazdanbakhsh et al., 2021). 2.2. Experimental procedures
Among these, sulfite is the most popular choice due to its virtues of
low-cost and environmentally friendly (weak mutagenic or other gen­ The photonic intensity was estimated to be 0.392 μEinstein/s (for the
otoxic effects) (Rao et al., 2021). In addition, residual sulfite in water 254 nm UV) and 0.063 μEinstein/s (for the 185 nm VUV), respectively,
can be desalinized by aeration (Wu et al., 2021). according to KI/KIO3 and the production rate of H2O2 methods (Li et al.,
VUV can emit ultraviolet light with wavelength of 185 and 254 nm 2020; Yang et al., 2018; Zhang et al., 2022). Details for the determina­
simultaneously. Compared with UV irradiation, VUV has the advantages tion of photon intensity at 254 and 185 nm were shown in Text S1 and
of hydrolysis and ionization directly to generate reactive species, such as Text S2, respectively. The photoreactor was wrapped with aluminum
hydroxyl radicals (HO⋅), hydrogen atoms (H⋅), and hydrated electrons foil to avoid the interference from other light sources.
(e−aq) Eqs. (1)-((2)) (Gonzalez et al., 2004; Buxton et al., 1988a). Due to The temperature was maintained at 25 ◦ C. Before starting the ex­
the high-energy photons production, VUV is more capable for degrading periments, the VUV and UV lamps were ignited and warmed up at least
chloroacetaldehyde, trichloroacetic acid and haloacetonitriles than UV 30 min to guarantee the stable energy output, unless otherwise noted,
(Moussavi and Rezaei, 2017; Gonzalez et al., 2004; Kiattisaksiri et al., the solution was purged with N2 at a pressure of 0.15 Mpa to blow off the
2016; Pan et al., 2022). Relevant literatures confirmed that the VUV/­ DO from the water firstly (20 min), and then continued to charge N2 to
sulfite process was superior to the UV/sulfite process in degradation of
haloacetic acid and perfluorooctane sulfonate (Gu et al., 2017; Zhang
et al., 2022). Therefore, VUV/sulfite may be a promising process for the
abatement of TCAcAm. However, there is no report on TCAcAm
decomposition by VUV/sulfite process, thus leaving some knowledge
gaps to be filled.

(1)
hv=185nm
H2 O + hv ̅̅̅̅̅̅̅̅̅̅̅→ H⋅ + HO⋅ Φ = 0.33

(2)
hv=185nm
H2 O + hv ̅̅̅̅̅̅̅̅̅̅̅→ H+ + HO⋅ + e−aq Φ = 0.045

In order to inspect the advantage and better understand the VUV/


sulfite process, the objective of this study is to: (i) examine the effec­
tiveness of degrading TCAcAm under VUV irradiation using three
different reductants (sulfite, sulfide, ferrous iron). (ii) recognize the
dominant reactive species and determine their corresponding contri­
butions to the TCAcAm degradation. (iii) evaluate the influence of key
parameters, including pH, sulfite dosage, dissolved oxygen (DO), and
water matrices on the TCAcAm degradation and predict the yields of
reactive species (e−aq, H⋅) under different conditions in VUV/sulfite Fig. 1. Schematic of the quartz cylindrical photoreactor. (1. Nitrogen source, 2.
ampler, 3. Quartz photoreactor bottle cap, 4. Quartz photoreactor bottle body,
process. (iv) elucidate the degradation pathway of TCAcAm in VUV/
5. UV lamp, 6. Quartz sleeves, 7. Fixed suction cups, 8. Round aeration discs, 9.
sulfite process.
Rotor, 10. Magnetic stirrer.)

2
H. Huang et al. Water Research 225 (2022) 119122

prevent oxygen from re-entering and maintain the solution at low DO


concentration (< 1 mg/L) during the reaction. 0.1 M phosphate buffer
was introduced to adjust the solution initial pH. 9 mL sample was
withdrawn at a pre-set time intervals (0, 0.5, 1, 2, 3, 4, 6, 8 min) for
analysis.

2.3. Analytical methods

3 g anhydrous Na2SO4 and 3 mL MTBE were introduced into 9 mL


sample and shaken by a circumferential oscillator (KS 130 B, IKA, Ger­
many) at 720 r/min for 10 min, and then left for 5 min. The concen­
trations of TCAcAm and haloacetic acids (HAAs) were determined by gas
chromatography (GC) which was coupled with an electron capture de­
tector (GC-ECD) (2010 plus, Shimadzu, Japan). 1 μL of the extract was
injected into a GC column (HP-5, 30.0 m × 0.25 mm × 0.25 μm). The
column oven was initially held at 45 ◦ C for 2 min, ramping to 220 ◦ C at
17 ◦ C/min and holding for 1 min. The injection port was controlled at
220 ◦ C and the ECD detector temperature was 290 ◦ C. For HAAs, the
Fig. 2. Degradation of TCAcAm using different reductants. Experimental con­
derivatization was performed before detection using 10 % sulfuric acid
ditions: [TCAcAm]0 = 200 μg/L, [sulfite]0 = [sulfide]0 = [ferrous iron]0 = 1
in methanol. Chloride ion concentrations in the sample were measured
mM, [DO] 0 < 1.0 mg/L, initial pH = 7.00 ± 0.05.
using ion chromatography (CIC-D160 SHINE ShengHan, China). The
H2O2 concentration was determined by UV-vis spectrophotometer (UV
Liu et al., 2014, 2013). Sulfide can absorb UV lights, resulting in the
2550, Shimadzu, Japan).
generation of excited state bisulfide ion (HS*− ) with a redox potential
(E) of -0.45 V (Yu et al., 2018; Melsheimer and Schlögl, 1997). Besides,
2.4. Kinetic modeling
UV irradiation of sulfite or ferrous iron produce e−aq (E = -2.9 V), while
sulfite generates additional other reactive species, such as H⋅ (E = -2.3
The concentrations of reactive species were simulated in VUV/sulfite
V) and SO⋅−3 (E = 0.8 V) (Zhang et al., 2022; Airey and Dainton, 1966).
process by kinetic modeling (Kintecus 6.80, chlorate iodine model)
The reduction ability of sulfite, sulfide and ferrous iron are relatively
(Ianni, 2021). The related reactions based on reactive species in
weak, which may be responsible for the poor TCAcAm removal (Song
VUV/sulfite process were summarized in Table S1, and additional de­
et al., 2013). In addition, VUV irradiation exhibited better TCAcAm
tails of the model were provided in Text S3.
degradation performance than UV irradiation at the same UV power (4
W), which probably because the TCAcAm degradation (under VUV) is
2.5. Trapping of reactive species by electron paramagnetic resonance
associated with direct photolysis by 185 + 254 nm, and indirect
(EPR)
photolysis by reactive species, which was generated by VUV via hy­
drolysis directly (Gonzalez et al., 2004). While under UV irradiation,
5, 5-dimethyl-1-pyrroline N-oxide (DMPO) was adopted to capture
TCAcAm can be decomposed only by direct photolysis by 254 nm,
reactive species, which were generated in VUV or VUV/sulfite process
however, the TCAcAm has a weak molar absorbance coefficient at 254
under N2-saturation conditions and by the EPR spectrometer (Bruker
nm (ε254 = 27.64 M− 1 cm− 1, detail information was provided in Text
EMXnano) (Liang et al., 2021). The additions of sulfite and DMPO were
S4). Only 13.5 % of TCAcAm removal was observed by VUV irradiation,
10 mM and 100 mM, respectively. In order to capture e−aq more easily in
which was lower than that of monochloroacetic acid (MCAA) and
the reactive species capture experiment, the initial pH was adjusted to
dichloroacetonitrile (DCAN) (Zhang et al., 2022; Wu et al., 2020). Two
9.2, unless otherwise noted. The magnetic field, modulation amplitude,
reasons may account for this phenomenon: (i) Low yield of e−aq in VUV
modulation frequency, microwave frequency, microwave power,
photolysis of H2O (Eqs. (1)-((2)). (ii) e−aq is consumed by H2O to generate
receiver gain, number of detections and sweep time were set as 344 ± 5
other reactive species (e.g., HO⋅, H⋅) Eqs. (3)-((5)) (Buxton et al.,
mT, 2 G, 100 KHz, 9.64 GHz, 10 mW, 40, 3, 30 s, respectively.
1988b).
With the addition of sulfite (1 mM), the TCAcAm decomposition was
3. Result and discussion
greatly enhanced during the VUV (or UV) process because the photolysis
of sulfite can generate an amount of reactive reductive species (e.g., e−aq,
3.1. Degradation of TCAcAm by VUV/sulfite process
H⋅ and SO⋅−3 ), which are effective for the dehalogenation of compounds
containing C-X (X = F, Cl, Br, I) bonds (Li et al., 2012; Fischer and
3.1.1. Degradation of TCAcAm using different reductants under VUV (or
Warneck, 1996; Yang et al., 2013).
UV) irradiation
Different reductants can significantly affect the removal of pollutants e−aq + H2 O→H⋅ + OH− k = 1.9 × 101 M− 1
s− 1
(3)
by ARPs (Milh et al., 2021). Thus, the degradation of TCAcAm using
different reductants (sulfite, sulfide, and ferrous iron) under VUV (or e−aq + HO⋅→ OH− k = 3.0 × 1010 M− 1
s− 1
(4)
UV) irradiation were investigated (Fig. 2).
Fig. 2 demonstrates that both reductants and VUV (or UV) were e−aq + H⋅→ H2 + OH− k = 2.5 × 1010 M− 1
s− 1
(5)
uncapable of abating TCAcAm effectively. Sulfite, sulfide and ferrous
iron only accounted for 4.9 %, 1.6 % and 0.8 % of TCAcAm removal, As shown in Fig. S2, the degradation of TCAcAm during VUV, UV/
respectively. Meanwhile, UV and VUV could achieve less than 15.0 % sulfite and VUV/sulfite processes followed pseudo-first-order kinetics
degradation of TCAcAm. However, when different reductants irradiated well. The first-order rate constant (kobs, min− 1) in VUV/sulfite process
under VUV (or UV), TCAcAm removal rates reached up to 85.6 % (or was 0.24416 min− 1, which was 14.3 and 1.3 folds greater than that in
78.0 %) in VUV (or UV)/sulfite under the same condition. VUV and UV/sulfite processes, respectively. The kobs during VUV irra­
The difference in degradation efficiency is mainly due to the gener­ diation was only 0.01705 min− 1, which indicated that the generated
ated reactive species with different reductive capabilities and the dif­ reactive species in VUV/sulfite process played a crucial role in the
ference of optimal photolysis wavelength of reductants Yu et al., 2018; TCAcAm degradation. Accordingly, VUV/sulfite process is expected to

3
H. Huang et al. Water Research 225 (2022) 119122

be an alternative technology for the elimination of TCAcAm. ( )2−


NO−3 + e−aq → NO⋅3 k = 9.7 × 109 M− 1
s− 1
(6)
3.1.2. Contribution of reactive species and VUV, UV to TCAcAm
( )2−
degradation NO−2 + e−aq → NO⋅2 k = 4.1 × 109 M− 1
s− 1
(7)
In order to explore the reaction mechanism, EPR experiments were
performed to detect the generated reactive species (RS), such as e−aq, H⋅, NO−3 + H⋅→(NO3 H⋅)− k = 1.4 × 106 M− 1
s− 1
(8)
HO⋅ and SO⋅-3 in VUV/sulfite process, and the results were shown in
Fig. 3a. NO−2 + H⋅→NO⋅ + OH− k = 7.1 × 108 M− 1
s− 1
(9)
The weak DMPO adduct of HO⋅ (DMPO-HO⋅) was detected during
VUV irradiation under anaerobic conditions, while the e−aq was not (CH3 )3 COH + HO⋅→H2 O + ⋅CH2 (CH3 )2 COH
(10)
captured by DMPO, which may be due to the low quantum yield of e−aq in k = 6.0 × 108 M− 1 s− 1
VUV process or its short lifetime. Fortunately, the signals of e−aq and SO⋅−3 [ ]⋅−
as indicated by DMPO-H and DMPO-SO⋅−3 were obtained in VUV irra­ (CH3 )3 COH + e−aq → (CH3 )3 COH
(11)
diation of N2-purged 10 mM sulfite solution at pH 9.2. And the DMPO-H k = 4.0 × 105 M− 1
s− 1

signal disappeared when 3 mM NaNO2 was introduced into the sulfite


solution. (CH3 )3 COH + H⋅→H2 + ⋅CH2 (CH3 )2 COH
(12)
Meanwhile, quenching experiments were conducted to determine k = 1.7 × 105 M− 1 s− 1
the contribution of these RS. Nitrate (NO−3 ) and nitrite (NO−2 ) can react
Fig. 3c shows the TCAcAm degradation by VUV/sulfite process with
rapidly with e−aq Eqs. (6),((7)), while NO−3 reacts much more slowly with
different scavengers. The addition of 1 mM TBA could significantly
H⋅ than NO−2 Eqs. (8)-((9)) Li et al., 2012; Buxton et al., 1988a, b), so
enhance the degradation of TCAcAm from 85.6 % to 94.8 %. To further
NO−3 and NO−2 were selected as quenchers of e−aq and H⋅. T-butanol (TBA)
demonstrate the role of HO⋅, the UV/H2O2 process was adopted to
can quench HO⋅ with a rate constant of 6.0 × 108 M− 1 s− 1 (Eq. (10)), but
degrade TCAcAm. As shown in Fig. S5, the concentration of TCAcAm
it has a weak effect on e−aq and H⋅ (Eqs. (11),((12)) (Buxton et al., 1988b,
remained almost unchanged in the presence or absence of 1 mM TBA,
1988a; Wu et al., 2020). Thus, it is crucial to determinate the suitable
which revealed that the contribution of HO⋅ to TCAcAm degradation
scavenger dosage in VUV/sulfite process. The calculation details about
could be ignored. However, the improvement of TCAcAm decomposi­
TBA dosage were shown in Text S5 according to the previous literature
tion in the presence of TBA during VUV/sulfite process may be attrib­
(Wu et al., 2020) and the dosages of NO−3 and NO−2 were determined by
uted to the fact that TBA can prevent e−aq from being consumed by HO⋅
chemical modeling software Kintecus 6.80 (Fig. 3b). The results indi­
(Eq. (4)). The introduction of NO−3 or NO−2 (300 μM) could decrease the
cated that 1 mM TBA could consume nearly 100 % HO⋅, but only 1.3 %
removal of TCAcAm significantly, and NO−2 exerted a greater negative
e−aq. 300 μM NO−2 could consume 99.9 % e−aq and ~100 % H⋅. On the
effect than that of NO−3 , which suggested that e−aq and/or H⋅ played an
contrary, 300 μM NO−3 could deplete ~100 % e−aq, while only 8.4 % H⋅.

Fig. 3. (a) EPR spectra of DMPO adducts of RS produced by VUV or VUV/sulfite, (b) Simulation the concentration of RS consumed by different scavengers, (c)
TCAcAm degradation by VUV/sulfite process with different scavengers, and (d) The contribution of RS, VUV and UV in the VUV/sulfite process. Experimental
conditions except (a): [TCAcAm]0 = 200 μg/L, [sulfite]0 = 1 mM, [DO] 0 < 1.0 mg/L, initial pH = 7.00 ± 0.05.

4
H. Huang et al. Water Research 225 (2022) 119122

important role in TCAcAm degradation. real-time concentration and the initial concentration of TCM (μg/L),
To further distinguish the contributions of e−aq, H⋅, UV, VUV and other kTCAcAm− e−aq and kTCM− e−aq are the second-order rate constant of e−aq with
RS, their relative contributions (R) were calculated via the pseudo-first- TCAcAm and TCM (M− 1 s− 1), respectively. As shown in Fig. S6, the
order rate constant (Table S2) in VUV/sulfite process using the following kTCAcAm− e−aq was measured as 2.41 × 1010 M− 1 s− 1. This result was lower
equations: than that the second-order rate constant of DCAN with e−aq (3.76 × 1010
( ) kobsVUV/sulfite − kobsNO− M− 1 s− 1) using the same method (Wu et al., 2020). The difference may
R e−aq = 3
× 100% (13) be attributed to the fact that TCAcAm has one more C-Cl bond in its
kobsVUV/sulfite
structure than DCAN even though they are both N-DBPs.
kobsNO−3 − kobsNO−2
(H⋅) = × 100% (14)
kobsVUV/sulfite 3.2. Influence of main reaction parameters in VUV/sulfite process

kobsUV 3.2.1. Influence of initial pH


R(UV) = × 100% (15)
kobsVUV/sulfite pH is considered as a key parameter for sulfite activation, which can
( ) affect the interaction between sulfite and target pollutions, as well as the
R(VUV and other RS) = 1 − R e−aq − R(H⋅) − R(UV) (16) yields of reactive species (Wu et al., 2021). As shown in Fig. 4, the effect
of initial pH on TCAcAm degradation was investigated.
The results were presented in Fig. 3d. R (e−aq), R (H⋅), R (UV) and R It can be seen from Fig. 4a that the degradation of TCAcAm increased
(VUV and other RS) were determined to be 83.5 %, 8.8 %, 4.7 % and 3.0 from 11.9 % to 100 % when the initial pH increased from 5.8 to 9.2. It is
%, respectively. The difference of R may be related to the redox potential well known that the existence forms of sulfite species (e.g., SO2− 3 , HSO3 )

of the reactive species, which are ranked in order of their reducing ca­ is highly pH-dependent (pKa2 = 7.2) Banayan Esfahani and Mohseni,
pacity: e−aq (E = -2.9 V) > H⋅ (E = -2.3 V) > SO⋅-3 (E = 0.8 V) (Buxton et al., 2022; Eldridge et al., 2016; Tartar and Garretson, 1941). SO2- 3 is the
1988b). Compared with the degradation of monochloroacetic acid in dominant sulfite species at pH > 7.2, while HSO−3 occupies a predomi­
VUV/sulfite process by Zhang et al., similar R (H⋅) was calculated in the nant position in pH range of 2.0-7.0. SO2− 3 can be easily irradiated to
same way, while the R(e−aq) was higher than that of them (61.1 %) generate e−aq (Eq. (18)) under alkaline conditions because it has a higher
(Zhang et al., 2022), which might be due to the fact that HO⋅ did not molar absorption coefficient and quantum yield than HSO−3 (Eq. (19))
contribute to the degradation of TCAcAm in this study. The low value of (Wu et al., 2021). Meanwhile, the reactions of e−aq with H+ and HSO−3
R (UV) can be ascribed to the relatively weak molar absorbance coeffi­ (Eqs. (20)-((21)) result in the decrease of e−aq yield in VUV/sulfite process
cient of TCAcAm at 254 nm (ε254 = 27.64 M− 1 cm− 1 in this study). These under acidic conditions (Xiao et al., 2017). pH can affect the generation
results revealed that e−aq was the dominant contributor to TCAcAm of reactive species and thus affect the decomposition of TCAcAm. To
decomposition. further explain the yields of reactive species, the yields of e−aq and H⋅ as a
function of initial pH were simulated by Kintecus 6.80 (Fig. 4b). As the
3.1.3. Second-order rate constant between TCAcAm and e−aq pH increased from 1.0 to 12.0, the concentration of e−aq increased from
Competitive kinetic method was employed to determine the second- 1.46 × 10− 18 to 2.54 × 10− 12 M and then kept stability under alkaline
order rate constant between e−aq and TCAcAm for further understanding conditions (pH 9.0-12.0). However, the concentration of H⋅ presented a
the reaction mechanism. Trichloromethane (TCM), which has a known downward tendency totally, which decreased from 4.65 × 10− 10 to 3.00
second-order rate constant with e−aq (3.0 × 1010 M− 1 s− 1) (Wu et al., × 10− 10 M slightly at first (pH 1.0-6.0), and then dropped sharply to
2020), was chosen as the probe compound for its similar structure with 5.15 × 10− 15 M (pH 12.0). Noteworthy, taking pH 7.0 as the cut-off
TCAcAm. Thus, the second-order rate constant of TCAcAm with e−aq point, e−aq was the dominant reactive species at pH > 7.0, while the
during UV/sulfite process (a process that created a good environment for high concentration of H⋅ was observed at pH < 7.0. Meanwhile, the
e−aq) (Yu et al., 2018; Li et al., 2012) could be calculated using the degradation efficiency of TCAcAm at pH 5.8 was only 11.9 %, which
following equation: indirectly indicated that the contribution of H⋅ was much lower than
that of e−aq. However, further increasing pH from 9.2 to 11.0 only
[TCAcAm]t kTCAcAm− e−aq [TCM]t
ln = ln (17) enhanced the decomposition of TCAcAm slightly, which may be attrib­
[TCAcAm]0 kTCM− e−aq [TCM]0
uted to the fact that the quantum yield of e−aq reaches the peak value of
2.54 × 10− 12 M (Fig. 4b) in VUV/sulfite process. Interestingly, the
Where [TCAcAm]t and [TCAcAm]0 are the real-time concentration and
removal of TCAcAm also achieved 85.6 % under neutral conditions, so
the initial concentration of TCAcAm (μg/L), [TCM]t and [TCM]0 are the
the initial pH of the experiment was set at 7.0 to approximate the actual

Fig. 4. (a) Degradation of TCAcAm at different initial pH and (b) The yields of e−aq and H⋅ as a function of initial pH in VUV/sulfite process. Experimental and
simulation conditions: [TCAcAm]0 = 200 μg/L, [sulfite]0 = 1 mM, [DO]0 < 1.0 mg/L and 0 mg/L, simulation time = 8 min.

5
H. Huang et al. Water Research 225 (2022) 119122

aquatic environment. sulfite results in the generation of more e−aq and SO⋅−3 . However, the
presence of high concentrations of e−aq and SO⋅−3 in the reaction solution
SO2−3 +hv→SO⋅−3 +e−aq (18)
may compete with TCAcAm for e−aq (Eqs. (22)-((23)) Fischer and War­
neck, 1996). In addition, e−aq also may be consumed by sulfite (Eqs.
HSO−3 + hv→SO⋅−3 + H k = 2.0 × 107 M− 1
s− 1
(19) (24)-((25)) (Zhang et al., 2022).

H+ + e−aq →H⋅ k = 2.3 × 1010 M− 1


s− 1
(20) e−aq +SO⋅−3 →SO2−3 (22)

HSO−3 + e−aq → H⋅ + SO2−3 k = 2.0 × 107 M− 1


s− 1
(21) e−aq + e−aq →H2 + 2OH− k = 5.5 × 109 M− 1
s− 1
(23)

3.2.2. Influence of sulfite dosage e−aq + SO2−3 →Product k < 1.3 × 106 M− 1
s− 1
(24)
The concentration of sulfite is crucial to the reductive degradation of
contaminants during sulfite activation (Wu et al., 2021). Therefore, the e−aq + HSO−3 → H⋅ + SO2−3 k = 2.0 × 107 M− 1
s− 1
(25)
effect of sulfite dosage on TCAcAm degradation and the yields of e−aq and
Fig. 5c indicates that the e−aq yield increased from 0 to 2.58 × 10− 12 M
H⋅ were investigated in VUV/sulfite process (Fig. 5).
as sulfite dosage increased from 0 to 4.0 mM, this further explains that
Compared with VUV (13.5 %) solely, the introduction of sulfite
the yield of e−aq was controlled by sulfite dosage. The conclusion is in
(0.25, 0.5, 1.0, 2.0, 4.0 mM) could enlarge the degradation of TCAcAm
accordance with the results of Song et al. in which the relative quasi-
(52.7 %, 63.5 %, 85.6 %, 85.9 %, 96.0 %, respectively) in VUV/sulfite
stationary concentration of e−aq was increased with increasing SO2− 3
process Fig. 5a). Obviously, the TCAcAm decomposition efficiency
concentration (Song et al., 2013). Noteworthy, the Kintecus 6.80
significantly increased from 52.7 % to 85.6 % with increasing the sulfite
simulation results also demonstrated the existence of two phases (fast
dosage from 0.25 mM to 1.0 mM. However, when the sulfite dosage
and slow), which was consistent with the experimental results (Fig. 5b).
further increased to 2.0 mM and 4.0 mM, the TCAcAm decomposition
This also explains that sulfite dosage was the determining factor for
efficiency was only enhanced by 0.3 % and 10.4 %, respectively,
controlling e−aq generation. Unlike the e−aq yield, the concentration of H⋅
compared to that of 1.0 mM sulfite. The kobs increased from 0.09426
firstly increased to 1.02 × 10− 10 M (0-0.75 mM) and then decreased
min− 1 to 0.40232 min− 1 when the sulfite dosage increased from 0.25
sharply to 3.51 × 10− 17 M (0.75-4.0 mM) with increasing sulfite dosage
mM to 4.0 mM. Fig. 5b shows that the TCAcAm decomposition can be
from 0 to 4.0 mM (Fig. 5d), which may be closely related to the depletion
divided into two stages: rapid growth and slow growth, which consistent
of H⋅ by the high concentration of e−aq (0.75-4 mM) (Eq. (5)).
with the study by Zhang et al. (Zhang et al., 2022). The phenomenon
may be related to the yield of e−aq, which was controlled by sulfite
3.2.3. Influence of dissolved oxygen
dosage. At rapid growth stage, the generated e−aq is largely consumed to
DO may exert a significant effect on the degradation of contaminants
degrade TCAcAm. While at slow growth stage, the presence of more

Fig. 5. Influence of sulfite dosage on the (a) degradation and (b) kobs of TCAcAm in VUV/sulfite process, and (c, d) The yields of e−aq and H⋅ in VUV/sulfite process.
Experimental and simulation conditions: [TCAcAm]0 = 200 μg/L, [DO] 0 < 1.0 and 0 mg/L, initial pH = 7.00 ± 0.05, simulation time = 8 min.

6
H. Huang et al. Water Research 225 (2022) 119122

in VUV/sulfite process (Cao et al., 2021). The influence of DO on TCA­ induce the generation of HO⋅, 1O2, H2O2 and other reactive oxygen
cAm decay was shown in Fig. 6. species through a series of chain reactions during VUV or UV irradiation
Compared to that of 8.5 mg/L (without N2 purge), the removal ef­ (Dalrymple et al., 2010; Lester et al., 2013), and these generated reactive
ficiency of TCAcAm decreased from 85.6 % to 21.7 % when the DO oxygen species will quench e−aq at different reaction rates. In addition,
concentration was lower than 1 mg/L Fig. 6a). DO is an excellent HA can also act as the optical filter to absorb UV light (Wang et al.,
scavenger of e−aq and H⋅, which can react with e−aq and H⋅ with rate 2012). Therefore, shading effect and/or quenching of e−aq are caused by
constants of 1.9 × 1010 M− 1 s− 1 and 1.2 × 1010 M− 1 s− 1, respectively HA, which may be account for inhibition effect on TCAcAm degradation
(Eqs. (26)-((27)) (Wu et al., 2021). DO can not only convert e−aq to the in VUV/sulfite process.
rather weakly reactive species O⋅−2 and HO⋅2, but also reduces the
available steady state concentrations of e−aq and H⋅ that for TCAcAm
degradation. 3.3. Intermediate products and mechanisms of TCAcAm decomposition
Fig. 6b predicts the yields of e−aq and H⋅ as a function of DO in VUV/ by VUV/sulfite process
sulfite process, which also supports the view that DO can reduce the
concentrations of reactive species. The yields of e−aq and H⋅ were grad­ Fig. 8 shows the concentration variations of TCAcAm, Cl− , dichlor­
ually decreased as DO increased in VUV/sulfite process. When DO oacetamide (DCAcAm), trichloroacetic acid (TCAA) and dichloroacetic
concentration was set at 8.0 mg/L, the concentrations of e−aq and H⋅ were acid (DCAA) during the TCAcAm degradation in VUV/sulfite process.
found to be 7.04 × 10− 16 M and 1.39 × 10− 18 M, respectively, which As shown in Fig. 8a, the theory and actual dechloriantion of TCAcAm
were 3.267 × 103 and 3.2 × 104 times lower than that of DO 0 mg/L. were studied. TCAcAm theory dechloriantion (TCl, TCl = 3 × [TCA­
cAm]t, where [TCAcAm]t is the real-time degradation concentration)
e−aq + O2 → O⋅−2 k = 1.9 × 1010 M− 1
s− 1
(26) efficiency was always higher than actual dechloriantion (ACl) efficiency,
which suggested that a certain of Cl-containing products were likely to
H⋅ + O2 →HO⋅2 k = 1.2 × 1010 M− 1
s− 1
(27) generated in TCAcAm degradation process. To confirm this hypothesis,
three intermediate products, namely DCAcAm, TCAA and DCAA were
3.2.4. Influence of water matrix detected by GC-ECD according to certified reference substances
Water is a mixture which consists of many components, such as (Fig. 8b), respectively, and the concentrations of intermediate products
chloride (Cl− ), bicarbonate (HCO−3 ) and dissolved organic matters were also quantified. The concentrations of DCAcAm and TCAA both
(DOM) (Wu et al., 2021). It is reported that the presence of Cl− , HCO−3 , firstly increased (0-20.26 μg/L and 0-4.84 μg/L) with reaction time (0-2
and DOM may pose an impact on the performance of AOPs/APRs min and 0-3 min), and then decreased to 0.61 μg/L and 2.21 μg/L at the
(Nawaz et al., 2017; Cao et al., 2021; Gu et al., 2017; Zhang et al., 2022), end of reaction (8 min), whereas the DCAA concentration was only
Therefore, the influence of Cl− , HCO−3 and DOM on TCAcAm degrada­ detected at the last sampling point.
tion in VUV/sulfite process were investigated (Fig. 7). The results revealed that reductive dechlorination and hydrolysis
A 15.5 % decline of TCAcAm degradation was observed in Fig. 7a at a coexisted during the degradation of TCAcAm in VUV/sulfite process,
Cl− concentration of 16 mM when compared with the absence of Cl− . which was consistent with results from previous researches (Chu et al.,
And the corresponding kobs decreased from 0.23932 min− 1 to 0.14818 2016; Chu et al., 2009). However, a significant difference existed be­
min− 1 when the concentration of Cl− increased from 2 mM to 16 mM tween the concentrations sum of intermediate products (the remain
(Fig. 7b). Cl− has a high molar absorption coefficient of 3500 M− 1 cm− 1 TCAcAm, formed DCAcAm, TCAA and DCAA) and the initial TCAcAm
at 185 nm (Zhang et al., 2022), which can compete with sulfite for VUV concentration in which the concentrations sum of intermediate products
light source. Furthermore, the reaction between Cl− and e−aq may be was much lower than the initial TCAcAm concentration (Fig. 8b).
enhanced at high Cl− concentration, and thus leading to a decrease in Moreover, the undetectable of monochloroacetamide (MCAcAm) and
TCAcAm degradation (Nawaz et al., 2017). The degradation of TCAcAm MCAA showed that the dechlorination of TCAcAm or TCAA was
was also inhibited from 82.8 % to 70.4 % with increasing HCO−3 con­ incomplete, and other products, such as DCAcAm sulfonate, might be
centration from 0.5 mM to 4 mM (Fig. 7c and d). The inhibitory effect of generated during VUV/sulfite process (Ding et al., 2018).
HCO−3 on TCAcAm decay can be attributed to the fact that e−aq is In view of the above results, three possible degradation pathways of
consumed by HCO−3 at a relatively slow second-order rate constant of < TCAcAm by VUV/sulfite process were proposed in Fig. 8c. (i) Most of
106 M− 1 s− 1 (Buxton et al., 1988b). TCAcAm degradation comes from the attack of C-Cl bond by e−aq and H⋅
Fig. 7e-f show that the decomposition of TCAcAm decreased from to form DCAcAm and Cl− . (ii) Some TCAcAm is reduced to DCAcAm
85.6 % to 32.1 % and kobs decreased from 0.24416 min− 1 to 0.04805 sulfonate and Cl− by SO⋅−3 , and then DCAcAm sulfonate quite slowly
min− 1 as HA concentration increased from 0 mg/L to 8 mg/L. HA can hydrolyses to form DCAcAm and SO2− 4 (Yiin et al., 1987; Ding et al.,

Fig. 6. (a) Influence of DO on the degradation of TCAcAm and (b) The yields of e−aq and H⋅ as a function of DO in VUV/sulfite process. Experimental and simulation
conditions: [TCAcAm]0 = 200 μg/L, [sulfite]0 = 1 mM, initial pH = 7.00 ± 0.05, simulation time = 8 min.

7
H. Huang et al. Water Research 225 (2022) 119122

Fig. 7. Influence of water matrix (Cl− , HCO−3 , HA) on the degradation of TCAcAm in VUV/sulfite process. Experimental conditions: [TCAcAm]0 = 200 μg/L,
[sulfite]0 = 1 mM, [DO] 0 < 1.0 mg/L, initial pH = 7.00 ± 0.05.

2018; Haag and Mill, 2010). (iii) A small part of TCAcAm is converted to combined toxicity values are 233, 168, 833 and 274 M− 1, respectively
TCAA because of its own hydrolysis reaction (Chu et al., 2009). Then, (Table S3) (Plewa et al., 2010; Plewa et al., 2008). To evaluate the toxic
TCAA can be further reduced by e−aq and H⋅ to form DCAA and Cl− . effect during TCAcAm decomposition by VUV/sulfite process, the
combined toxicity risk value was calculated by Eq. (28) (Table S4) ac­
cording to the previous research (Chu et al., 2016).
3.4. Combined toxicity risk during TCAcAm decomposition by VUV/ ∑[ ]
sulfite process CTRVt = CTVProducts × CProducts, t (28)

Intermediate products, including DCAcAm, TCAA, DCAA, were Where CTRVt is the combined toxicity risk value at t reaction time (min),
detected during the degradation of TCAcAm by VUV/sulfite process. As CTVProducts is the combined toxicity value (M− 1) of each detected in­
reported, TCAcAm, DCAcAm, TCAA and DCAA are a series of DBPs with termediate product, CProducts, t is the intermediate product concentration
cytotoxicity and genotoxicity in drinking water (Table S3), and their (nM) at t reaction time (min).

8
H. Huang et al. Water Research 225 (2022) 119122

Fig. 8. (a) Theory and actual dechloriantion (TCl and ACl) in TCAcAm degradation by VUV/sulfite process, (b) The formation of intermediate products in TCAcAm
degradation process, and (c) Possible pathways of TCAcAm degradation by VUV/sulfite process. TCl = 3 × [TCAcAm]t. Experimental conditions: [sulfite]0 = 1 mM,
[DO] 0 < 1.0 mg/L, initial pH = 7.00 ± 0.05.

Although the combined toxicity values of TCAA (833 M− 1) and 4. Conclusions


DCAA (274 M− 1) were higher than TCAcAm (233 M− 1), the combined
toxicity risk value was still gradually decreased during the TCAcAm The VUV/sulfite process presented the best performance for TCA­
decomposition by VUV/sulfite process (Fig. S7). This may be attributed CAm removal. e−aq, as one of the most reactive species, played an
to the fact that the generated intermediate products were further important role in the TCAcAm decomposition. The second-order rate
degraded by VUV/sulfite process, and the degradation rate was higher constant of e−aq reacting with TCAcAm was determined to be 2.41 × 1010
than the production rate. The results supported that VUV/sulfite process M− 1 s− 1. DO posed a negative effect on the degradation of TCAcAm,
could not only effectively degrade TCAcAm, but also show favorable which can be due to the fact that DO can significantly consume e−aq. With
toxicity inhibition. the increase of pH and sulfite dosage, the degradation efficiency of
TCAcAm increased sharply, which may be attributed to the fact that e−aq
is more stable under alkaline conditions and increasing sulfite dosage

9
H. Huang et al. Water Research 225 (2022) 119122

results in the rising of the e−aq concentration. The presence of HA obvi­ Buxton, G.V., Greenstock, C.L., Helman, W.P., Ross, A.B., 1988b. Critical Review of rate
constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals
ously hindered TCAcAm degradation, whereas Cl− and HCO−3 exhibited
(⋅OH/⋅O− ) in aqueous solution. J. Phys. Chem. Ref. Data.
slight inhibition. Intermediate products, including DCAcAm, TCAA, Cao, Y., Qiu, W., Li, J., Jiang, J., Pang, S., 2021. Review on UV/sulfite process for water
DCAA and Cl− , were detected, indicating that reductive dechlorination and wastewater treatments in the presence or absence of O2. Sci. Total Environ. 765,
and hydrolysis coexisted during the degradation of TCAcAm in VUV/ 142762.
Chen, L., Deng, Y., Dong, S., Wang, H., Li, P., Zhang, H., Chu, W., 2021. The occurrence
sulfite process. VUV/sulfite process could not only decompose TCAcAm and control of waterborne viruses in drinking water treatment: a review.
but also remove the formation of intermediate products and inhibit Chemosphere 281, 130728.
consequence on toxicity during the degradation of TCAcAm, supporting Chen, S., Chu, W., Wei, H., Zhao, H., Xu, B., Gao, N., Yin, D., 2018. Reductive
dechlorination of haloacetamides in drinking water by Cu/Fe bimetal. Sep. Purif.
its effective applicability in the drinking water treatment. Therefore, Technol. 203, 226–232.
VUV/sulfite process is expected to be an alternative technology for the Chu, W., Gao, N., Deng, Y., 2009. Stability of newfound nitrogenous disinfection by-
elimination of TCAcAm in drinking water. products haloacetamides in drinking water. Chin. J. Org. Chem. 29 (10), 1569–1574.
Chu, W., Li, D., Gao, N., Templeton, M.R., Tan, C., Gao, Y., 2015. The control of emerging
haloacetamide DBP precursors with UV/persulfate treatment. Water Res. 72,
CRediT authorship contribution statement 340–348.
Chu, W., Li, X., Bond, T., Gao, N., Bin, X., Wang, Q., Ding, S., 2016. Copper increases
reductive dehalogenation of haloacetamides by zero-valent iron in drinking water:
Huahan Huang: Conceptualization, Data curation, Formal analysis, reduction efficiency and integrated toxicity risk. Water Res. 107, 141–150.
Investigation, Methodology, Writing – original draft. Xinrui Liang: Chuang, Y., Parker, K.M., Mitch, W.A., 2016. Development of predictive models for the
Conceptualization, Data curation, Formal analysis. Qingsong Li: degradation of halogenated disinfection byproducts during the UV/H2O2 advanced
oxidation process. Environ. Sci. Technol. 50 (20), 11209–11217.
Funding acquisition, Investigation, Methodology, Project administra­ Dalrymple, R.M., Carfagno, A.K., Sharpless, C.M., 2010. Correlations between dissolved
tion, Writing – review & editing. Jing Deng: Supervision, Validation, organic matter optical properties and quantum yields of singlet oxygen and
Visualization, Writing – review & editing. Jing Zou: Data curation, hydrogen peroxide. Environ. Sci. Technol. 44 (15), 5824–5829.
Deng, Y., Zhang, Y., Zhang, R., Wu, B., Ding, L., Xu, K., Ren, H., 2014. Mice in vivo
Writing – review & editing. Xia Li: Validation, Data curation. Xiaoyan toxicity studies for monohaloacetamides emerging disinfection byproducts based on
Ma: Writing – review & editing, Validation. Guoxin Li: Writing – review metabolomic methods. Environ. Sci. Technol. 48 (14), 8212–8218.
& editing, Validation. Guoyuan Chen: Writing – review & editing, Ding, S., Wang, F., Chu, W., Cao, Z., Pan, Y., Gao, N., 2018. Rapid degradation of
brominated and iodinated haloacetamides with sulfite in drinking water:
Validation. degradation kinetics and mechanisms. Water Res. 143, 325–333.
Du, Y., Zhao, L., Ban, J., Zhu, J., Wang, S., Zhu, X., Zhang, Y., Huang, Z., Li, T., 2021.
Cumulative health risk assessment of disinfection by-products in drinking water by
Declaration of Competing Interest different disinfection methods in typical regions of China. Sci. Total Environ. 770,
144662.
Eldridge, D.L., Guo, W., Farquhar, J., 2016. Theoretical estimates of equilibrium sulfur
The authors declare that they have no known competing financial isotope effects in aqueous sulfur systems: highlighting the role of isomers in the
interests or personal relationships that could have appeared to influence sulfite and sulfoxylate systems. Geochim. Cosmochim. Ac. 195, 171–200.
the work reported in this paper. Fang, X., Zeng, Z., Li, Q., Liu, Y., Chu, W., Maiyalagan, T., Mao, S., 2021. Ultrasensitive
detection of disinfection byproduct trichloroacetamide in drinking water with Ag
nanoprism@MoS2 heterostructure-based electrochemical sensor. Sens. Actuators B
Data availability 332, 129526.
Fischer, M., Warneck, P., 1996. Photodecomposition and photooxidation of hydrogen
sulfite in aqueous solution. J. Phys. Chem. 100 (37), 15111–15117 (1952).
Data will be made available on request. Gonzalez, M.G., Oliveros, E., Wörner, M., Braun, A.M., 2004. Vacuum-ultraviolet
photolysis of aqueous reaction systems. J. Photochem. Photobiol. C 5 (3), 225–246.
Gu, Y., Liu, T., Wang, H., Han, H., Dong, W., 2017. Hydrated electron based
decomposition of perfluorooctane sulfonate (PFOS) in the VUV/sulfite system. Sci.
Acknowledgments Total Environ. 607-608, 541–548.
Haag, W.R., Mill, T., 2010. Some reactions of naturally occurring nucleophiles with
This study was supported in part by the National Natural Science haloalkanes in water. Environ. Toxicol. Chem. 7 (11), 917–924.
Ianni, J.C., 2021. Kintecus, Windows Version 6.80 (www.kintecus.com), p. Chemical
Foundation of China (Grant No. 51878582, 51978618), the Guiding modeling software for regression and optimization of chemical kinetics simulation of
Project of Fujian Province of China (No. 2021Y0041), the Natural Sci­ combustion, nuclear, enzyme, catalyst reactor (CSTR) and other processes and
ence Foundation of Fujian Province of China (No. 2020J01256), New reactions.
Jo, C.H., Dietrich, A.M., Tanko, J.M., 2011. Simultaneous degradation of disinfection
Century Excellent Talents in Fujian Province University (JA14227), byproducts and earthy-musty odorants by the UV/H2O2 advanced oxidation process.
Natural Science Foundation of Zhejiang Province (LY21E080018) and Water Res. 45 (8), 2507–2516.
Foundation of Key Laboratory of Yangtze River Water Environment and Kiattisaksiri, P., Khan, E., Punyapalakul, P., Ratpukdi, T., 2016. Photodegradation of
haloacetonitriles in water by vacuum ultraviolet irradiation: mechanisms and
Ministry of Education (Tongji University), China (YRWEF201901). intermediate formation. Water Res. 98, 160–167.
Krasner, S.W., Weinberg, H.S., Richardson, S.D., Pastor, S.J., Chinn, R., Sclimenti, M.J.,
Onstad, G.D., Thruston, A.D., 2006. Occurrence of a new generation of disinfection
Supplementary materials
byproducts. Environ. Sci. Technol. 40 (23), 7175–7185.
Lester, Y., Sharpless, C., Mamane, H., Linden, K., 2013. Production of photo-oxidants by
Supplementary material associated with this article can be found, in dissolved organic matter during UV water treatment. Environ. Sci. Technol. 47.
the online version, at doi:10.1016/j.watres.2022.119122. Li, B., Ma, X., Li, Q., Chen, W., Deng, J., Li, G., Chen, G., Liao, W., 2020. Factor affecting
the role of radicals contribution at different wavelengths, degradation pathways and
toxicity during UV-LED/chlorine process. Chem. Eng. J. 392, 124552.
References Li, M., Shi, Y., Sun, S., Qian, Y., An, D., 2022. Degradation pathways and kinetics of
chloroacetonitriles by UV/persulfate in the presence of bromide. Sci. Total Environ.
834, 155373.
Airey, P.L., Dainton, F.S., 1966. The photochemistry of aqueous solutions of Fe (II). I.
Li, X., Ma, J., Liu, G., Fang, J., Yue, S., Guan, Y., Chen, L., Liu, X., 2012. Efficient
Photoelectron detachment from ferrous and ferrocyanide ions. Proc. R. Soc. Lond.
reductive dechlorination of monochloroacetic acid by sulfite/UV process. Environ.
291 (1426), 340–352.
Sci. Technol. 46 (13), 7342.
Banayan Esfahani, E., Mohseni, M., 2022. Fluence-based photo-reductive decomposition
Liang, S., Xu, S., Wang, C., Ling, J., Xian, Z., Wu, H., Tian, H., Zhou, S., Gu, C., 2021.
of PFAS using vacuum UV (VUV) irradiation: effects of key parameters and
Enhanced alteration of poly(vinyl chloride) microplastics by hydrated electrons
decomposition mechanism. J. Environ. Chem. Eng. 10 (1), 107050.
derived from indole-3-acetic acid assisted by a common cationic surfactant. Water
Bao, L., Maruya, K.A., Snyder, S.A., Zeng, E.Y., 2012. China’s water pollution by
Res. 191, 116797.
persistent organic pollutants. Environ. Pollut. 163, 100–108.
Ling, L., Sun, J., Fang, J., Shang, C., 2016. Kinetics and mechanisms of degradation of
Bond, T., Huang, J., Templeton, M.R., Graham, N., 2011. Occurrence and control of
chloroacetonitriles by the UV/H2O2 process. Water Res. 99, 209–215.
nitrogenous disinfection by-products in drinking water – a review. Water Res. 45
Liu, X., Vellanki, B.P., Batchelor, B., Abdel-Wahab, A., 2014. Degradation of 1,2-
(15), 4341–4354.
dichloroethane with advanced reduction processes (ARPs): effects of process
Buxton, G.V., Greenstock, C.L., Helman, W.P., Ross, A.B., 1988a. Critical Review of rate
variables and mechanisms. Chem. Eng. J. 237, 300–307.
constants for reactions of hydrated electrons Chemical Kinetic Data Base for
Combustion Chemistry. Part 3: Propane. J. Phys. Chem. Ref. Data 17, 536–635.

10
H. Huang et al. Water Research 225 (2022) 119122

Liu, X., Yoon, S., Batchelor, B., Abdel-Wahab, A., 2013. Photochemical degradation of Song, Z., Tang, H., Wang, N., Zhu, L., 2013. Reductive defluorination of
vinyl chloride with an Advanced Reduction Process (ARP) – effects of reagents and perfluorooctanoic acid by hydrated electrons in a sulfite-mediated UV
pH. Chem. Eng. J. 215-216, 868–875. photochemical system. J. Hazard. Mater. 262, 332–338.
Luo, J., Hu, J., Wei, X., Fu, L., Li, L., 2015. Dehalogenation of persistent halogenated Tang, Y., Long, X., Wu, M., Yang, S., Gao, N., Xu, B., Dutta, S., 2020. Bibliometric review
organic compounds: a review of computational studies and quantitative of research trends on disinfection by-products in drinking water during 1975–2018.
structure–property relationships. Chemosphere 131, 17–33. Sep. Purif. Technol. 241, 116741.
Melsheimer, J., Schlögl, R., 1997. Identification of reaction products of mild oxidation of Wang, C., Zhu, L., Wei, M., Chen, P., Shan, G., 2012. Photolytic reaction mechanism and
H2S in solution and in solid state by UV-VIS spectroscopy. Fresenius. J. Anal. Chem. impacts of coexisting substances on photodegradation of bisphenol A by Bi2WO6 in
357 (4), 397–400. water. Water Res. 46 (3), 845–853.
Milh, H., Yu, X., Cabooter, D., Dewil, R., 2021. Degradation of ciprofloxacin using UV- Wang, K., Guo, J., Yang, M., Junji, H., Deng, R., 2009. Decomposition of two haloacetic
based advanced removal processes: comparison of persulfate-based advanced acids in water using UV radiation, ozone and advanced oxidation processes.
oxidation and sulfite-based advanced reduction processes. Sci. Total Environ. 764, J. Hazard. Mater. 162 (2), 1243–1248.
144510. Wu, S., Shen, L., Lin, Y., Yin, K., Yang, C., 2021. Sulfite-based advanced oxidation and
Moussavi, G., Rezaei, M., 2017. Exploring the advanced oxidation/reduction processes in reduction processes for water treatment. Chem. Eng. J. 414, 128872.
the VUV photoreactor for dechlorination and mineralization of trichloroacetic acid: Wu, Z., Shang, C., Wang, D., Zheng, S., Wang, Y., Fang, J., 2020. Rapid degradation of
parametric experiments, degradation pathway and bioassessment. Chem. Eng. J. dichloroacetonitrile by hydrated electron (e–aq) produced in vacuum ultraviolet
328, 331–342. photolysis. Chemosphere 256, 126994.
Nawaz, S., Shah, N.S., Khan, J.A., Sayed, M., Al-Muhtaseb, A.H., Andersen, H.R., Xiao, Q., Wang, T., Yu, S., Yi, P., Li, L., 2017. Influence of UV lamp, sulfur (IV)
Muhammad, N., Murtaza, B., Khan, H.M., 2017. Removal efficiency and economic concentration, and pH on bromate degradation in UV/sulfite systems: mechanisms
cost comparison of hydrated electron-mediated reductive pathways for treatment of and applications. Water Res. 111, 288–296.
bromate. Chem. Eng. J. 320, 523–531. Yang, L., Li, M., Li, W., Qiang, Z., 2018. A green method to determine VUV (185 nm)
Nihemaiti, M., Le Roux, J., Hoppe-Jones, C., Reckhow, D.A., Croué, J., 2017. Formation fluence rate based on hydrogen peroxide production in aqueous solution.
of haloacetonitriles, haloacetamides, and nitrogenous heterocyclic byproducts by Photochem. Photobiol. 94.
chloramination of phenolic compounds. Environ. Sci. Technol. 51 (1), 655–663. Yang, S., Sun, J., Hu, Y., Cheng, J., Liang, X., 2013. Effect of vacuum ultraviolet on
Pan, H., Huang, Y., Li, J., Li, B., Yang, Y., Chen, B., Zhu, R., 2022. Coexisting oxidation ultrasonic defluorination of aqueous perfluorooctanesulfonate. Chem. Eng. J. 234,
and reduction of chloroacetaldehydes in water by UV/VUV irradiation. Water Res. 106–114.
214, 118192. Yazdanbakhsh, A., Eslami, A., Mahdipour, F., Ghanbari, F., Ghasemi, S.M., Atamaleki, A.,
Plewa, M.J., Muellner, M.G., Richardson, S.D., Fasano, F., Buettner, K.M., Woo, Y., Maleksari, H.S., Lin, K.A., 2021. Dye degradation in aqueous solution by dithionite/
McKague, A.B., Wagner, E.D., 2008. Occurrence, synthesis, and mammalian cell UV-C advanced reduction process (ARP): kinetic study, dechlorination, degradation
cytotoxicity and genotoxicity of haloacetamides: an emerging class of nitrogenous pathway and mechanism. J. Photochem. Photobiol. A 407, 112995.
drinking water disinfection byproducts. Environ. Sci. Technol. 42 (3), 955–961. Yiin, B.S., Walker, D.M., Margerum, D.W., 1987. Nonmetal redox kinetics: general-acid-
Plewa, M.J., Simmons, J.E., Richardson, S.D., Wagner, E.D., 2010. Mammalian cell assisted reactions of chloramine with sulfite and hydrogen sulfite. Inorg. Chem. 26
cytotoxicity and genotoxicity of the haloacetic acids, a major class of drinking water (21), 3435–3441.
disinfection by-products. Environ. Mol. Mutagen. 51 (8-9), 871–878. Yu, K., Li, X., Chen, L., Fang, J., Chen, H., Li, Q., Chi, N., Ma, J., 2018. Mechanism and
Rao, D., Dong, H., Lian, L., Sun, Y., Zhang, X., Dong, L., Zhou, G., Guan, X., 2021. New efficiency of contaminant reduction by hydrated electron in the sulfite/iodide/UV
mechanistic insights into the transformation of reactive oxidizing species in an process. Water Res. 129, 357–364.
ultraviolet/sulfite system under aerobic conditions: modeling and the impact of Mn Yu, X., Cabooter, D., Dewil, R., 2018. Effects of process variables and kinetics on the
(II). ACS ES&T Water 1 (8), 1785–1795. degradation of 2,4-dichlorophenol using advanced reduction processes (ARP).
Richardson, S., PLEWA, M., WAGNER, E., SCHOENY, R., DEMARINI, D., 2007. J. Hazard. Mater. 357, 81–88.
Occurrence, genotoxicity, and carcinogenicity of regulated and emerging Zhang, D., Wang, F., Duan, Y., Chen, S., Zhang, A., Chu, W., 2020. Removal of
disinfection by-products in drinking water: a review and roadmap for research. Mut. trihalomethanes and haloacetamides from drinking water during tea brewing:
Res./Rev. Mut. Res. 636 (1-3), 178–242. removal mechanism and kinetic analysis. Water Res. 184, 116148.
Sedlak, D.L., Gunten, U.V., 2011. The chlorine dilemma. Science 331 (6013), 42–43. Zhang, J., Zhang, H., Liu, X., Cui, F., Zhao, Z., 2022. Efficient reductive and oxidative
Shah, A.D., Mitch, W.A., 2012. Halonitroalkanes, halonitriles, haloamides, and n- decomposition of haloacetic acids by the vacuum-ultraviolet/sulfite system. Water
nitrosamines: a critical review of nitrogenous disinfection byproduct formation Res. 210, 117974.
pathways. Environ. Sci. Technol. 46 (1), 119–131. Zhang, Y., Zhao, F., Deng, Y., Zhao, Y., Ren, H., 2015. Metagenomic and metabolomic
analysis of the toxic effects of trichloroacetamide-induced gut microbiome and urine
metabolome perturbations in mice. J. Proteome Res. 14 (4), 1752–1761.

11

You might also like