Next Article in Journal
Comprehensive Performance Assessment for Sponge City Construction: A Case Study
Next Article in Special Issue
Novel Oxidation Strategies for the In Situ Remediation of Chlorinated Solvents from Groundwater—A Bench-Scale Study
Previous Article in Journal
A Case Study for Stability Analysis of Toppling Slope under the Combined Action of Large Suspension Bridge Loads and Hydrodynamic Forces in a Large Reservoir Area
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Abatement of Nitrophenol in Aqueous Solution by HOCl and UV/HOCl Processes: Kinetics, Mechanisms, and Formation of Chlorinated Nitrogenous Byproducts

1
College of Resources and Environmental Sciences, Nanjing Agricultural University, Nanjing 210095, China
2
Univ Lyon, Université Claude Bernard Lyon 1, CNRS, IRCELYON, F-69626 Villeurbanne, France
*
Author to whom correspondence should be addressed.
Water 2023, 15(23), 4038; https://doi.org/10.3390/w15234038
Submission received: 11 October 2023 / Revised: 12 November 2023 / Accepted: 14 November 2023 / Published: 21 November 2023
(This article belongs to the Special Issue New Technologies for Soil and Groundwater Remediation)

Abstract

:
HOCl and UV activated HOCl (UV/HOCl) have been applied for water disinfection and abatement of organic contaminants. However, the production of toxic byproducts in the HOCl and UV/HOCl treatment should be scrutinized. This contribution comparatively investigated the elimination of 4-nitrophenol and the generation of chlorinated byproducts in HOCl and UV/HOCl treatment processes. 61.4% of 4-nitrophenol was removed by UV/HOCl in 5 min with HOCl dose of 60 μM, significantly higher than that by UV (3.3%) or HOCl alone (32.0%). Radical quenching test showed that HO and Cl played important roles in UV/HOCl process. 2-Chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol were generated consecutively in HOCl process; but their formation was less in the UV/HOCl process. Trichloronitromethane (TCNM) was only found in the UV/HOCl process, and its production increased with increasing HOCl dosage. Besides chlorinated products hydroxylated and dinitrated products were also identified in the UV/HOCl process. Transformation pathways involving electrophilic substitution, hydroxylation, and nitration were proposed for 4-nitrophenol transformation in the UV/HOCl process. Wastewater matrix could significantly promote the transformation of 4-nitrophenol to 2-chloro-4-nitrophenol in UV/HOCl process. Results of this study are helpful to advance the understanding of the transformation of nitrophenolic compounds and assess the formation potential of chlorinated byproducts in HOCl and UV/HOCl disinfection processes.

1. Introduction

Chlorination is extensively applied for disinfection in drinking water treatment worldwide [1,2]. However, chlorine alone does not effectively inactivate chlorine-resistant pathogens, such as Giardia and Cryptosporidium [3,4,5]. Ultraviolet (UV) light has a potential to modify the nucleic structure of microorganisms, thus reducing the risk of water-borne pathogens [6,7]. The combination of UV and chlorine (UV/HOCl) has attracted extensive attention due to its efficient disinfection performance and propensity to decompose organic contaminants [8,9].
The photolysis of hypochlorous acid/hypochlorite ion (HOCl/OCl) can directly generate reactive oxidizing species including chlorine (Cl) and hydroxyl radicals (HO) (Equation (1)) [10]. As a non-selective oxidant, HO can react with various contaminants at nearly diffusion-controlled rates (108–1010 M−1 s−1) [11]. In contrast, Cl prefers to interact with substances containing aromatic rings and electron-rich moieties selectively, such as phenols, benzoic acids, toluenes, and anilines, with second rate constants of 1.2–2.5 × 1010 M−1 s−1 [12,13]. HO and Cl also react with Cl or HOCl/OCl to generate other secondary radicals (i.e., ClO and Cl2•−) (Equations (2)–(6)) [14,15,16]. These reactive species make UV/HOCl a promising strategy for the elimination of a wide spectrum of micropollutants [17,18]. It should be noted that, the quantum yields for HOCl/OCl photolysis are greater than 1.2 by UV254 irradiation (pH 4–10) due to the chain reactions [10]. The quantum yields are also highly dependent on the irradiation wavelength, solution pH, and concentration of HOCl/OCl [19,20,21].
HOCl/OCl + hv → HO/O•− + Cl
Cl + Cl ↔ Cl2•− k+ = 6.5 × 109 M−1s−1, k = 1.1 × 105 s−1
HOCl + HO → ClO + H2O k = 2.0 × 109 M−1s−1
OCl + HO → ClO + OH k = 8.8 × 109 M−1s−1
HOCl + Cl → ClO + H+ + Cl k = 3.0 × 109 M−1s−1
OCl + Cl → ClO + Cl k = 8.2 × 109 M−1s−1
The UV/hydrogen peroxide (UV/H2O2) process is the conventional UV-based advanced oxidation process (AOP) [18]. However, under UV254 irradiation, the molar absorption coefficient (ε) and quantum yield (Φ) of H2O2 ( ε H 2 O 2 = 19.6 M−1 cm−1, Φ H 2 O 2 = 1.0) are relatively lower than HOCl/OCl ( ε HOCl = 59 M−1 cm−1, ε OCl = 66 M−1 cm−1, Φ HOCl / OCl greater than 1.2) [18,22]. In the UV/H2O2 process, only 5–10% of the dosed H2O2 is consumed and the H2O2 residual is required to be removed by further process in drink water treatment, which causes an increase in cost [18,23]. The UV/persulfate (UV/PS) process is also an alternative AOP, which generates sulfate radicals (SO4•−) and HO [18,23]. However, UV/PS is sensitive to pH, chloride, and inorganic carbon [23]. This sensitivity can affect its treatment efficiency depending on the reactivity of radicals with individual contaminants [23]. Therefore, compared with these two UV-based AOPs, UV/HOCl is a more cost-effective process [18].
HOCl can transform numerous organic and inorganic micropollutants typically encountered in drinking water and wastewater (e.g., ferrous, arsenic, nitrite, phenols, pesticides, drugs, etc.) [24,25]. Phenols are toxic organic compounds widely used as raw materials for synthesis of pesticides, dyes, drugs, plastics, and antioxidants [1]. Since HOCl is not as strong an oxidant to mineralize organic micropollutants, potentially harmful chlorinated byproducts may be formed by electrophilic substitution [25,26,27]. Electrophilic substitution is one of the most important mechanisms accounting for the transformation of phenols during chlorination [28]. As a consequence of the ortho/para orientation of the hydroxyl group, the chlorination of phenol occurs through consecutive substitution at the 2, 4, and 6 positions, resulting in the generation of mono-, di-, and tri-chlorophenols [25]. However, during UV/HOCl treatment, the transformation rate of phenol is faster than chlorination alone, which is ascribed to the joint action of reactive radicals (e.g., HO and Cl) and HOCl [29]. In addition to electrophilic substitution by HOCl, phenol also undergoes fast reaction with reactive chlorine species (e.g., Cl and Cl2•−) [13,30]. Cl can extract hydrogen atom from phenol to form phenoxyl radical, which subsequently undergoes Cl addition generating chlorophenols [13]. Also, Cl2•− reacts moderately fast with phenol by single electron transfer or hydrogen abstraction to generate phenoxyl radical [30]. Chlorophenols are toxic and difficult to biodegrade and are lethal to a variety of organisms [31]. These chemicals have been listed as priority organic contaminants [32], and their use and production are restricted [31]. Therefore, the formation of chlorophenols in the processes of HOCl and UV/HOCl is a matter of serious concern.
The disinfection byproducts (DBPs) generated by the interaction between disinfectants and organic contaminants, or natural organic matter (NOM) have been a major concern of water safety [33,34,35]. Compared with the commonly studied DBPs (e.g., trihalomethanes, THMs, and haloacetic acids, HAAs), the unregulated DBPs, in particular those nitrogenous disinfection byproducts (N-DBPs), have been less studied although they are of higher carcinogenicity and mutagenicity [36,37]. As a significant group of N-DBPs, halonitromethanes (HNMs) have attracted great consideration because of their high genotoxicity and cytotoxicity [38,39]. Among the studied HNMs, trichloronitromethane (TCNM) comprises more than 50% proportion of total HNMs formed during drinking water treatment [40,41]. Therefore, TCNM has been recognized as a surrogate of HNMs to study the formation potential and regulation strategy of HNMs [41]. The amount of TCNM in disinfected water was reported to be 0.2–0.5 μg L−1 [42]. However, reports on the potential and mechanism of TCNM formation in UV/HOCl processes are limited [43]. In addition, halonitrophenols are a newly identified group of nitrogen-containing aromatic halogenated DBPs [16,44,45]. Toxicological studies have indicated that aromatic halogenated DBPs generally induced higher developmental toxicity and growth inhibition than the commonly known aliphatic halogenated DBPs [44,45]. A study on the constitution of the DBP mixture from chlorination of bromide-rich raw water indicated that aromatic fractions accounted for 49–67%, which dominated the developmental toxicity of chlorinated water samples [44]. Because studies indicated that the one- and two-carbon-atom DBPs of current interest accounted for only ~16% of disinfected water cytotoxicity, there is a need to identify toxicity drivers within the poorly characterized higher-molecular-weight (more than two carbon atoms) DBP fraction, e.g., halonitrophenols [46]. However, reports on the formation potential and mechanisms of halonitrophenols and TCNM in UV/HOCl processes are limited [43]. In consequence, it is still essential to investigate the formation potential of halonitrophenols and TCNM during the degradation of substituted phenols in the UV/HOCl process in order to assure the safety of drinking water.
4-Nitrophenol is a nitro-substituted phenol with high toxicity and poor biodegradability [47]. It is extensively found in the industrial effluent of pesticides and synthetic dyes [47]. 4-Nitrophenol has endocrine disrupting effect, and its biological toxicity is greater than that of 2- and 3-nitrophenol [48,49]. In consequence, 4-nitrophenol can cause serious harmful effect on human health and ecosystem. This chemical is also ranked as a priority pollutant by the U.S. EPA with a range from 1 to 20 ppb for its maximum allowable concentrations [50]. As a result, it is advisable to develop efficacious water treatment methods to remove 4-nitrophenol from aqueous solution. When the water containing 4-nitrophenol is disinfected by HOCl and UV/HOCl processes, hydroxylated and chlorinated intermediates are likely formed under the electrophilic attack of HOCl, HO, and Cl [28,51,52]. Further oxidation of these intermediates may give rise to N-DBPs through ring-opening. However, the transformation mechanism of 4-nitrophenol in the HOCl and UV/HOCl processes, and the formation potential of N-DBPs have been barely reported, so further investigations are essentially needed. In real water, the efficiency of HOCl and UV/HOCl processes is influenced by water matrix, e.g., dissolved organic matter (DOM), alkalinity, halides, and ammonia [18,53]. DOM significantly restrains the degradation of some micropollutants by UV/HOCl process through scavenging radicals (e.g., HO and Cl), consuming the oxidant HOCl/OCl, and filtering UV light [18,53]. Carbonate radicals (CO3•−) can be produced from the scavenging of HO and Cl by HCO3 or CO32− [18]. Due to the scavenging effect of (bi)carbonate and overall lower reactivity of CO3•− with micropollutants, the presence of HCO3 and CO32− is usually disadvantageous for micropollutant destruction [54]. In addition, halides (e.g., Cl and Br) affect UV/HOCl process negatively or positively depending on the structures of targeted pollutants [18,55]. Therefore, effects of wastewater matrix on the transformation of 4-nitrophenol during the HOCl and UV/HOCl processes also warrant investigation.
The goals of this work are: (1) to investigate the transformation kinetics of 4-nitrophenol by HOCl and UV/HOCl processes and identify chlorinated intermediates; (2) to evaluate the formation potential of TCNM in HOCl and UV/HOCl processes; (3) to propose the degradation mechanisms and pathways of 4-nitrophenol in HOCl and UV/HOCl processes; (4) to investigate effects of wastewater matrix on the transformation of 4-nitrophenol during the HOCl and UV/HOCl processes.

2. Materials and Methods

2.1. Reagents and Materials

Chemicals, suppliers, and purities are provided in Text S1 of Supplementary Materials.

2.2. Chlorination Experiments

50 mL aqueous solution involving 10 μM 4-nitrophenol was buffered to neutral condition with 10 mM phosphate buffer. Note that the initial concentration of 4-nitrophenol (10 μM) was higher than the environmental concentrations, just for the purpose to amplify the reactions for better detection of transformation products. 10 μM HOCl was spiked into the reaction vials to initiate the chlorination reaction. The vials were retained in the dark at 20 °C. A control experiment without HOCl was carried out concurrently. 1 mL sample was collected from each vial at pre-set time intervals and quenched promptly with 20 μL 1 M Na2S2O3. Samples were kept at 4 °C in the dark and analyzed within 24 h.

2.3. UV/HOCl Experiments

The UV/HOCl experiments were conducted in a BL-GHX-V photoreactor (Shanghai Bilon Instrument Co., Ltd., Shanghai, China) coupled with an ozone-free, low-pressure mercury lamp (15 W) emitting predominantly 254 nm monochromatic light (UV254). The photoreactor was turned on for 15 min to achieve lamp emission stabilization in advance. The average UV fluence rate was measured to be 8.7 × 10−7 Einstein L−1 s−1 by a chemical actinometer [56]. Experimental conditions of the UV/HOCl treatment were the same as chlorination except that reaction solutions were transferred into quartz vials and exposed to UV254 irradiation. The effect of HOCl dosage (ranged from 40 to 100 μM) on the elimination of 4-nitrophenol during the UV/HOCl process was investigated. A control experiment without HOCl (direct photolysis) was carried out concurrently. To assess the contribution of the free radicals to 4-nitrophenol elimination by UV/HOCl, an experiment was carried out by spiking 5 mM TBA as radical quencher under identical operating conditions to those in the UV/HOCl treatment.

2.4. HPLC Analyses

The concentrations of 4-nitrophenol and its chlorinated byproducts in the samples were determined using a Hitachi L-2000 high performance liquid chromatograph (HPLC) with an L-2455 diode array detector (DAD). Separation was achieved by an Agilent Zorbax Eclipse Plus C18 reverse phase column (5 μm, 4.6 mm × 250 mm). The injection volume was 10 μL. The elution was isocratic with a mixture of 65% methanol (with 0.1% formic acid) and 35% water (with 0.1% formic acid) at a flow rate of 1 mL min−1. The detection wavelength of 4-nitrophenol, 2-chloro-4-nitrophenol, and 2,6-dichloro-4-nitrophenol was 312 nm, and the corresponding retention times were 4.7, 7.0, 10.5 min, respectively.

2.5. Quantification of TCNM

The formation potential of TCNM in the HOCl and UV/HOCl processes was tested in a set of vials. The reaction solution was 50 mL, containing 10 μM 4-nitrophenol and 60 μM HOCl. Other reaction conditions were identical to those as mentioned above. At pre-set time points, reaction solutions were quenched with 1 mL 1 mM Na2SO3. The quenched reaction solutions were refrigerated at 4 °C until further treatment and analysis.
The concentration of TCNM was measured according to the U.S. EPA standard method 551.1 [57]. TCNM was extracted by methyl tert-butyl ether (MtBE) and analyzed using an Agilent 7890 gas chromatograph (GC) equipped with an electron capture detector (ECD). A DB5-fused silica capillary column (1.5 μm, 0.53 mm × 30 m) was used for separation. The analytical parameters were as follows: inlet temperature of 220 °C; detector temperature of 280 °C; detector current of 1 nA. The heating procedure was described as follows: the initial temperature of 35 °C kept for 6 min; raising from 35 to 100 °C at a speed of 10 °C min−1 and kept 100 °C for 2 min; rising from 100 to 200 °C at a speed of 20 °C min−1 and kept 200 °C for 2 min, with a total of 27.5 min.

2.6. Identification of Intermediate Products

To identify intermediates, two aqueous solutions (50 mL, pH 7.0) containing 10 μM 4-nitrophenol and 60 μM HOCl were irradiated for 30 min and then concentrated by solid phase extraction (SPE) using hydrophilic-lipophilic balance (HLB) cartridges. Details of the procedures of SPE are provided in Text S2 of Supplementary Materials.
The SPE-concentrated samples were examined by a high-resolution mass spectrometer (HRMS, Triple TOF 5600+ MS/MS, AB Sciex, Boston, MA, USA) coupled with an HPLC (Shimadzu, Kyoto, Japan) to determine intermediate products. Analytical separation was accomplished on a Kinetex LC C18 column (2.6 μm, 100 Å, 100 mm × 2.1 mm, i.d.). The sample injection volume was 20 µL. The elution flow rate was 0.3 mL min−1 with a mixture of eluent A (H2O plus 0.1% formic acid) and eluent B (MeOH plus 0.1% formic acid), employing a linear gradient as follows: 10% B, 0–1 min; 10% to 90% B, 1–13 min, and 90% kept for 5 min; 90% to 10%, 18–18.1 min, and 10% kept for 1.9 min. Mass analyses were carried out under negative electrospray ionization (ESI(–)) in the TOF MS full scan (m/z 50–1000) mode. Operating parameters of ESI were set as: ion spray voltage floating of −4.5 kV; source temperature of 550 °C; gas I and gas II of 65 psi; curtain gas of 35 psi.

2.7. Evaluate the Influence of Wastewater

Wastewater was also used to evaluate the influences of water constituents on the transformation of 4-nitrophenol during the HOCl and UV/HOCl processes. Wastewater was obtained from a secondary sedimentation tank of a municipal sewage treatment plant (Nanjing city, China). The water sample was filtered through 0.22 μm cellulose acetate membrane before using. Water quality parameters including TOC, Abs254, nitrate, nitrite, pH, and typical inorganic ions are shown in Table S1 of Supplementary Materials. Similar chlorination and UV/HOCl experiments were conducted in wastewater as mentioned above, except that no phosphate buffer was used.

3. Results and Discussions

3.1. Degradation Kinetics

The degradation kinetics of 4-nitrophenol by UV irradiation, chlorination alone, and UV/HOCl are illustrated in Figure 1a. No loss of 4-nitrophenol was observed under dark without HOCl. Direct photolysis of 4-nitrophenol was negligible (~3.3%) within 5 min upon UV254 irradiation, which is in agreement with the report of Zhao et al. [47]. The photostability of 4-nitrophenol may be ascribed to the low quantum yield of the excited state of 4-nitrophenol, which probably stems from the stabilization effect of the −NO2 group as a strong para electron-acceptor substituent [47,58]. Chlorination resulted in approximately 32.0% removal of 4-nitrophenol over a course of 5 min. Notably, the elimination of 4-nitrophenol during UV/HOCl process was appreciably enhanced. A degradation efficiency of 61.4% was accomplished within 5 min by UV/HOCl. The enhanced removal efficiency of 4-nitrophenol in UV/HOCl process indicates that, except for HOCl, there are other reactive species responsible for the elimination of 4-nitrophenol. HOCl and OCl can produce reactive free radicals (e.g., HO and Cl) under UV irradiation, which were likely involved in the elimination of 4-nitrophenol. In order to determine whether free radicals contributed to the elimination of 4-nitrophenol, TBA was spiked into the UV/HOCl system as a radical quencher for HO and Cl [59]. TBA reacts quickly with HO and Cl with the second-order rate constants of 1.9 × 109 and 6.0 × 108 M−1 s−1, respectively [60]. As shown in Figure 1a, the presence of 5 mM TBA appreciably depressed the elimination of 4-nitrophenol, and the removal efficiency decreased from 61.4% to 33.6%. Therefore, it can be inferred that HO and Cl significantly contributed to the removal of 4-nitrophenol in the UV/HOCl process. The fact that 4-nitrophenol could still degrade in the UV/HOCl process with 5 mM TBA indicates that HOCl per se also played a significant role.
The elimination of 4-nitrophenol by UV/HOCl could be well fitted by pseudo-first-order kinetics (Equation (7)). As HOCl dosage increased from 40 to 100 μM, the observed pseudo-first-order rate constant of 4-nitrophenol (kobs) increased linearly ( R 2 = 0.99 ) from 0.137   ±   0.016 to 0.416   ±   0.073 min−1 (Figure 1b). This result shows that the elimination of 4-nitrophenol by UV/HOCl was positively affected by the amount of HOCl. With the increase of HOCl dose, the concentration of reactive radicals produced by HOCl/OCl photolysis increased, which accelerated the degradation of 4-nitrophenol. Similar result was observed by Guo et al. in the decomposition of chlortoluron by UV/HOCl process [59].
ln [ 4 - Nitrophenol ] t [ 4 - Nitrophenol ] 0 = k obs t

3.2. Formation of Chlorinated Nitrophenols

Electron-rich phenols undergo electrophilic substitution with HOCl [25]. The generation of chlorinated products should be scrutinized because they are generally more toxic than their parent compounds [61,62]. In this study, two halogenated intermediates, 2-chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol, were identified by HPLC during the chlorination of 4-nitrophenol. 2-Chloro-4-nitrophenol was observed to be the main byproduct, and its concentration gradually increased to 5.3 μM with prolonged reaction time within 15 min (Figure 2a). The formation of 2,6-dichloro-4-nitrophenol was observed after 5 min, consistent with the speculation that dichlorophenols are derived from further chlorination of the monochlorophenols [25]. After reaction for 15 min, the concentration of 2,6-dichloro-4-nitrophenol accumulated to 0.63 μM. The mass balance calculation showed that the sum of these two halogenated products accounted for ~92% of 4-nitrophenol, indicating that chlorination is the primary transformation mechanism of 4-nitrophenol in the presence of HOCl alone.
Interestingly, when UV/HOCl was applied for treatment of 4-nitrophenol, only 2-chloro-4-nitrophenol was identified by HPLC (Figure 2b). At HOCl dosage of 60 μM, the concentration of 2-chloro-4-nitrophenol reached the maximum (1.15 μM) within 2 min, afterwards it gradually decreased to 0.35 μM at 15 min. Obviously, the evolution pattern of 2-chloro-4-nitrophenol in UV/HOCl system was disparate to that in HOCl system. The hump shape of 2-chloro-4-nitrophenol profile in UV/HOCl process indicated that reactive radicals (e.g., HO and Cl) could promote further transformation of 2-chloro-4-nitrophenol. This hypothesis was further confirmed by radical quenching experiments. As shown in Figure 2b, when 5 mM TBA was spiked into UV/HOCl system, the concentration of 2-chloro-4-nitrophenol increased continuously as the irradiation time was extended, reaching 2.8 μM at 15 min. This observation confirms again that 2-chloro-4-nitrophenol is predominantly generated by electrophilic substitution between HOCl and 4-nitrophenol (Equation (8)), whereas further transformation of 2-chloro-4-nitrophenol requires the contribution of reactive radicals (Equation (9)). Notably, the amount of 2-chloro-4-nitrophenol produced in the UV/HOCl process with 5 mM TBA was significantly lower than that in the direct chlorination process (2.8 vs. 5.3 μM). This may be due to the fact that a fraction of HOCl was photochemically decomposed under UV254 radiation, leading to a decrease in the concentration of free chlorine capable of halogenating 4-nitrophenol.
4-Nitrophenol + HOCl     2-Chloro-4-nitrophenol
2-Chloro-4-nitrophenol + HO / Cl   Products
In addition, the concentration of HOCl remarkably affected the generation of 2-chloro-4-nitrophenol in UV/HOCl process (Figure 3). As the dosage of HOCl rose from 40 to 100 μM, the maximum amount of 2-chloro-4-nitrophenol elevated from 0.91 to 1.82 μM. It should be noted that the increase of HOCl dosage will also increase the concentration of reactive radicals during UV/HOCl process. Therefore, although the concentration of 2-chloro-4-nitrophenol raised rapidly at the inception phase of the reaction, it decreased rapidly thereafter due to the oxidation by more reactive radicals. For instance, when the dosage of HOCl was 80 μM, the resulting 2-chloro-4-nitrophenol was almost completely converted within 15 min. HO and Cl are highly oxidative species (with redox potentials of 2.8 and 2.4 V, respectively) that are capable of oxidizing many organic pollutants. Therefore, we speculate that 2-chloro-4-nitrophenol generated in the UV/HOCl process may undergo further oxidation (e.g., ring-opening cleavage) under the attack of HO and Cl, resulting in the production of low-molecular-weight organic compounds such as N-DBPs.
Based on Equations (8) and (9), the generation and transformation of 2-chloro-4-nitrophenol by UV/HOCl can be depicted as Equation (10).
d 2-Chloro-4-nitrophenol dt = k 1 4-Nitrophenol k 2 2-Chloro-4-nitrophenol
where [4-Nitrophenol] and [2-Chloro-4-nitrophenol] are the concentrations of 4-nitrophenol and 2-chloro-4-nitrophenol, respectively; k 1 and k 2 are the pseudo-first-order rate constants for the generation and transformation of 2-chloro-4-nitrophenol, respectively. Because the degradation of 4-nitrophenol obeys pseudo-first-order kinetics (as Equation (7)), the concentration of 4-nitrophenol can be calculated as Equation (11). Therefore, Equation (10) can be modified as Equation (12). With rearrangement of Equation (12), the generation and transformation of 2-chloro-4-nitrophenol in UV/HOCl process can be described by a consecutive kinetic model (Equation (13)).
4-Nitrophenol = 4-Nitrophenol 0 e k 1 t
d 2-Chloro-4-nitrophenol dt = k 1 4-Nitrophenol 0 e k 1 t k 2 2-Chloro-4-nitrophenol
2-Chloro-4-nitrophenol = k 1 4-Nitrophenol 0 k 2 k 1 e k 1 t e k 2 t
The kinetic model was used for fitting experimental data by least-square nonlinear regression to determine the k 1 and k 2 values of 2-chloro-4-nitrophenol under different HOCl dosages (Table 1). Results suggest that 2-chloro-4-nitrophenol had higher generation and lower transformation rate constants in the presence of low concentration of HOCl. In contrast, lower generation and higher transformation rate constants of 2-chloro-4-nitrophenol were observed with high concentration of HOCl. These results suggest that higher dosage of HOCl could result in higher concentrations of radicals, which was favorable for further transformation of 2-chloro-4-nitrophenol.

3.3. Formation Potential of TCNM

By GC analysis, we quantified the formation potential of TCNM during the UV/HOCl process, as shown in Figure 4a. When the HOCl dosage was set at 60 μM, the formation of TCNM was not observed within the first 5 min. After 5 min of reaction, TCNM concentration increased remarkably, accompanied by a steady decline of 2-chloro-4-nitrophenol concentration (see Figure 2b). This phenomenon indicates that the generation of TCNM is a result of further transformation of 2-chloro-4-nitrophenol. Studies have shown that nitrophenols (e.g., 3-nitrophenol) can be good precursors of TCNM during chlorine disinfection [63]. These precursors undergo chlorination and subsequent ring-opening reactions, resulting in the generation of TCNM and other N-DBPs. With the reaction time reaching 30 min, the yield of TCNM gradually increased to 5.8 nM. Therefore, the reactive radicals generated by photolysis of HOCl play a crucial role in opening of the aromatic rings of chlorinated intermediates to generate TCNM. Note that no production of TCNM was observed in the presence of HOCl alone, possibly owing to the relatively weak oxidation ability of HOCl (E0 = 1.48 V) [64]. In addition, due to the electron-withdrawing effect of Cl substituent [65], the reactivity of 2-chloro-4-nitrophenol with HOCl is less as compared to the parent 4-nitrophenol. Therefore, it is not easy for HOCl alone to oxidatively cleave the benzene ring as compared to more reactive radicals. The lack of TCNM is also consistent with the fact that electrophilic substitution was the predominant mechanism accounting for 4-nitrophenol transformation in the presence of HOCl alone (Figure 2a).
It is also found that the formation potential of TCNM during the UV/HOCl process correlated well with the concentration of HOCl. As shown in Figure 4b, with the increase of HOCl dosage from 40 to 100 μM, the yield of TCNM increased by 5.7-fold (from 2.4 to 13.6 nM) after 30 min reaction. The increase of HOCl concentration led to the generation of 2-chloro-4-nitrophenol in the early stage, and also accelerated the degradation of 2-chloro-4-nitrophenol due to the formation of more reactive radicals. Therefore, the formation potential of TCNM increased significantly with the increase of HOCl dosage. Due to the high genotoxicity and cytotoxicity of TCNM [38,39], its formation needs to be controlled and regulated during the UV/HOCl process for removing 4-nitrophenol from wastewater.

3.4. Intermediate Products

The maximum concentration of 2-chloro-4-nitrophenol accounted for only 30% of transformed 4-nitrophenol in UV/HOCl system. Therefore, other intermediate products might also be generated. The reaction solutions of 4-nitrophenol treated by UV/HOCl for 30 min were concentrated by SPE and then examined by HRMS for identification of those intermediates escaped from HPLC analyses. The structural elucidation of intermediates was according to: (1) the measured mass with error less than 10 ppm of the theoretical accurate mass based on the elemental composition; (2) the fragmentation pattern of the molecular ions; (3) the isotopic distribution endowed by chlorine atoms (i.e., isotopic abundance ratio of 35Cl and 37Cl). Based on the above tactics, we preliminarily identified 5 intermediates of 4-nitrophenol (see Figure S1 and Table S2 of Supplementary Materials). Among them, 2-chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol are chlorinated products of 4-nitrophenol (note that 2,6-dichloro-4-nitrophenol was produced at a low concentration during the UV/HOCl process and could not be directly identified and quantified by HPLC). In addition, two nitrated products (i.e., 2,4-dinitrophenol, 2-chloro-4,6-dinitrophenol) and one hydroxylated product (e.g., 4-nitrocatechol) were identified. Based on the identified products and radical chemistry, the transformation pathways of 4-nitrophenol in the UV/HOCl process are described in Scheme 1. Detailed mechanisms accounting for the transformation of 4-nitrophenol by UV/HOCl process are illustrated as follows.

3.5. Transformation Mechanisms and Pathways

It is well known that HOCl reacts with electron-rich phenols to form chlorinated products via electrophilic substitution [25]. Substitution reactions usually proceed in the ortho or para position of the phenolic hydroxyl substituent due to electronic resonance effects [25]. Since the para position of 4-nitrophenol is occupied by the nitro group, one of its ortho position should be substituted by Cl to form 2-chloro-4-nitrophenol. Because 2-chloro-4-nitrophenol also has an unsubstituted ortho position, it is likely subjected to HOCl electrophilic attack again and generates 2,6-dichloro-4-nitrophenol sequentially. Meanwhile, a fraction of HOCl is decomposed into HO and Cl under UV254 irradiation. Under the attack of reactive radicals, ring-opening reactions of 2-chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol occur, resulting in the generation of TCNM (see Scheme 1).
In addition, 4-nitrophenol undergoes electrophilic substitution by HO attack, leading to the generation of hydroxylated products. HO attacks the ortho position of 4-nitrophenol to produce 4-nitrocatechol [51]. Under further attack of HO, the C-N bond of 4-nitrocatechol is broken, producing 1,2,4-trihydroxybenzene and nitrogen dioxide radical (NO2). A similar phenomenon of NO2 production has previously been reported in the oxidation of nitroaromatic compounds by HO attack [66]. These hydroxylated products undergo ring-opening reactions under further attack of HO to form oxygen-containing aliphatic compounds, which are eventually mineralized into H2O and CO2 [67]. When HO directly attacks the nitro position, 4-nitrophenol is converted to hydroquinone [51]. Further reaction of hydroquinone with HO also leads to the formation of 1,2,4-trihydroxybenzene, which is eventually mineralized.
In many advanced oxidation processes, NO2 is regarded as a potent nitrating agent that leads to the generation of nitrated byproducts [68,69,70]. NO2 can be formed by many pathways, including the photolysis of NO3, decomposition of peroxynitrous acid (HOONO), oxidation of NO2, etc. [68,71]. In this study, NO2 may be derived either from the denitration of 4-nitrocatechol or 4-nitrophenol under the attack of HO/Cl, or through HO/Cl oxidation of NO2 converted from nitro substituent. As an electrophilic nitrating reagent, NO2 can react with electron-rich compounds to generate nitrated byproducts through hydrogen atom abstraction, electrophilic addition or substitution [71,72]. For instance, 4-nitrophenol can undergo electrophilic substitution reaction with NO2 to form 2,4-dinitrophenol [73]. Similarly, 2-chloro-4-nitrophenol can be converted to 2-chloro-4,6-dinitrophenol under electrophilic attack of NO2. All these nitrated byproducts can undergo further ring-opening and oxidation reactions under the attack of HO/Cl, which finally forms TCNM and other small molecular compounds.

3.6. Effects of Wastewater Matrix

As expected, the wastewater matrix slightly inhibited the transformation of 4-nitrophenol by HOCl process as compared to Milli-Q water (Figure 5a). This is likely due to the consumption of HOCl by reducing species (e.g., Fe2+, S2−) and effluent organic matter (EfOM) as they are ubiquitously present in wastewater.
Like in Milli-Q water, the elimination efficiency of 4-nitrophenol in wastewater by UV/HOCl process was significantly higher than that by HOCl process. However, the conversion of 4-nitrophenol by UV/HOCl process was increased in wastewater as compared to Milli-Q water. This unusual phenomenon may be associated with the indirect photolysis of 4-nitrophenol caused by some wastewater components (e.g., EfOM and NO3) under UV254 irradiation. These components have photosensitizing activity and can generate transient reactive species upon radiation, such as triplet excited state (3EfOM*) and HO [74,75]. Note that EfOM may quench a fraction of HOCl and HO/Cl, thus decreasing the removal efficiency of target compounds [76,77]. However, we cannot rule out the possibility that the transient reactive species mentioned above (e.g., 3EfOM*) may also promote the decay of organic compounds.
In both wastewater and Milli-Q water, the generation of 2-chloro-4-nitrophenol in HOCl treatment showed a continuously increasing trend (Figure 5b), suggesting that HOCl alone could not convert 2-chloro-4-nitrophenol efficiently. The concentration of 2-chloro-4-nitrophenol in wastewater was 4.63 μM in 15 min, slightly lower than that in Milli-Q water (5.31 μM). In UV/HOCl process, however, the concentration of 2-chloro-4-nitrophenol in both water matrices showed a first increasing and then decreasing pattern. This suggest that reactive radicals such as HO and Cl are important to further convert 2-chloro-4-nitrophenol. Most noticeably, the maximum concentration of 2-chloro-4-nitrophenol in wastewater was 4.80 μM, much higher than that in Milli-Q water (1.15 μM). This observation might be attributed to the light screening effect of wastewater constituents that partially inhibited the photolysis of HOCl [10,78]. The reduced decomposition of HOCl was conducive for chlorination of 4-nitrophenol to generate 2-chloro-4-nitrophenol. Overall, the effects of wastewater matrix appear to be complex and further investigation is still necessary.

4. Conclusions

Results showed that the elimination efficiency of 4-nitrophenol by UV/HOCl was significantly higher than that by chlorination alone. Radical quenching experiments confirmed the important roles of HO and Cl in the removal of 4-nitrophenol during UV/HOCl process. Electrophilic substitution is the predominant reaction mechanism for 4-nitrophenol transformation in HOCl process, which quantitatively results in the formation of 2-chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol. In UV/HOCl process, the chlorinated product 2-chloro-4-nitrophenol can undergo further transformation upon the attack of HO/Cl. Other intermediates, including 4-nitrocatechol, 2,4-dinitrophenol, and 2-chloro-4,6-dinitrophenol were identified by HRMS. Among them, chlorinated nitrophenolic derivatives can undergo further ring-opening and oxidation reactions under the attack of reactive radicals, generating TCNM and other small molecular organic compounds. The formation of chlorinated nitrophenols and TCNM in HOCl and UV/HOCl disinfection processes should be scrutinized because of their carcinogenicity and mutagenicity. The wastewater matrix could significantly promote the transformation of 4-nitrophenol to 2-chloro-4-nitrophenol in UV/HOCl process, which might be related with the formation of transient reactive species and light screening effect of wastewater constituents.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/w15234038/s1, Figure S1: Detailed mass spectra of intermediate products generated during the UV/HOCl treatment of 4-nitrophenol; Table S1: Characteristics of wastewater used in this work; Table S2: Mass spectra data and proposed molecular structures for the intermediate products generated during the UV/HOCl treatment of 4-nitrophenol. Reference [79] is cited in Supplementary Materials file.

Author Contributions

X.L., Investigation and Writing-original draft. Y.C., Methodology, Formal analysis, Visualization. J.L., Supervision. J.-M.C., Writing-Renew and Editing. J.C., Writing-Renew and Editing. C.J., Writing-Renew and Editing. Y.J., Conceptualization, Funding acquisition and Writing-Renew and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Grant No. 22076080).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Ge, F.; Zhu, L.Z.; Chen, H.R. Effects of pH on the chlorination process of phenols in drinking water. J. Hazard. Mater. 2006, 133, 99–105. [Google Scholar] [CrossRef] [PubMed]
  2. Abou Mehrez, O.; Dossier-Berne, F.; Legube, B. Chlorination and chloramination of aminophenols in aqueous solution: Oxidant demand and by-product formation. Environ. Technol. 2015, 36, 310–316. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, X.J.; Hu, X.X.; Wang, H.B.; Hu, C. Synergistic effect of the sequential use of UV irradiation and chlorine to disinfect reclaimed water. Water Res. 2012, 46, 1225–1232. [Google Scholar] [CrossRef]
  4. Zhou, S.Q.; Wu, Y.T.; Zhu, S.M.; Sun, J.L.; Bu, L.J.; Dionysiou, D.D. Nitrogen conversion from ammonia to trichloronitromethane: Potential risk during UV/chlorine process. Water Res. 2020, 172, 115508. [Google Scholar] [CrossRef]
  5. Deng, L.; Luo, W.; Chi, X.; Huang, T.T.; Wen, L.J.; Dong, H.Y.; Wu, M.X.; Hu, J. Formation of halonitromethanes from methylamine in the presence of bromide during UV/Cl2 disinfection. J. Environ. Sci. 2022, 117, 28–36. [Google Scholar] [CrossRef]
  6. Song, K.; Mohseni, M.; Taghipour, F. Application of ultraviolet light-emitting diodes (UV-LEDs) for water disinfection: A review. Water Res. 2016, 94, 341–349. [Google Scholar] [CrossRef]
  7. Li, G.Q.; Huo, Z.Y.; Wu, Q.Y.; Lu, Y.; Hu, H.Y. Synergistic effect of combined UV-LED and chlorine treatment on Bacillus subtilis spore inactivation. Sci. Total Environ. 2018, 639, 1233–1240. [Google Scholar] [CrossRef]
  8. Remucal, C.K.; Manley, D. Emerging investigators series: The efficacy of chlorine photolysis as an advanced oxidation process for drinking water treatment. Environ. Sci. Water Res. Technol. 2016, 2, 565–579. [Google Scholar] [CrossRef]
  9. Cai, W.W.; Peng, T.; Yang, B.; Xu, C.; Liu, Y.S.; Zhao, J.L.; Gu, F.L.; Ying, G.G. Kinetics and mechanism of reactive radical mediated fluconazole degradation by the UV/chlorine process: Experimental and theoretical studies. Chem. Eng. J. 2020, 402, 126224. [Google Scholar] [CrossRef]
  10. Watts, M.J.; Linden, K.G. Chlorine photolysis and subsequent OH radical production during UV treatment of chlorinated water. Water Res. 2007, 41, 2871–2878. [Google Scholar] [CrossRef]
  11. Wu, Z.H.; Fang, J.Y.; Xiang, Y.Y.; Shang, C.; Li, X.C.; Meng, F.G.; Yang, X. Roles of reactive chlorine species in trimethoprim degradation in the UV/chlorine process: Kinetics and transformation pathways. Water Res. 2016, 104, 272–282. [Google Scholar] [CrossRef]
  12. Mártire, D.O.; Rosso, J.A.; Bertolotti, S.; Le Roux, G.C.; Braun, A.M.; Gonzalez, M.C. Kinetic study of the reactions of chlorine atoms and Cl2•− radical anions in aqueous solutions. II. Toluene, benzoic acid, and chlorobenzene. J. Phys. Chem. A 2001, 105, 5385–5392. [Google Scholar] [CrossRef]
  13. Lei, Y.; Cheng, S.S.; Luo, N.; Yang, X.; An, T.C. Rate Constants and Mechanisms of the Reactions of Cl and Cl•− with Trace Organic Contaminants. Environ. Sci. Technol. 2019, 53, 11170–11182. [Google Scholar] [CrossRef]
  14. Fang, J.Y.; Fu, Y.; Shang, C. The Roles of Reactive Species in Micropollutant Degradation in the UV/Free Chlorine System. Environ. Sci. Technol. 2014, 48, 1859–1868. [Google Scholar] [CrossRef] [PubMed]
  15. Xiong, R.H.; Lu, Z.J.; Tang, Q.; Huang, X.L.; Ruan, H.Z.; Jiang, W.; Chen, Y.Q.; Liu, Z.Z.; Kang, J.X.; Liu, D.Q. UV-LED/chlorine degradation of propranolol in water: Degradation pathway and product toxicity. Chemosphere 2020, 248, 125957. [Google Scholar] [CrossRef]
  16. Liu, H.Y.; Hou, Z.C.; Li, Y.J.; Lei, Y.J.; Xu, Z.H.; Gu, J.J.; Tian, S.L. Modeling degradation kinetics of gemfibrozil and naproxen in the UV/chlorine system: Roles of reactive species and effects of water matrix. Water Res. 2021, 202, 117445. [Google Scholar] [CrossRef] [PubMed]
  17. Kong, X.J.; Jiang, J.; Ma, J.; Yang, Y.; Liu, W.L.; Liu, Y.L. Degradation of atrazine by UV/chlorine: Efficiency, influencing factors, and products. Water Res. 2016, 90, 15–23. [Google Scholar] [CrossRef]
  18. Yeom, Y.J.; Han, J.R.; Zhang, X.R.; Shang, C.; Zhang, T.; Li, X.Y.; Duan, X.D.; Dionysiou, D.D. A review on the degradation efficiency, DBP formation, and toxicity variation in the UV/chlorine treatment of micropollutants. Chem. Eng. J. 2021, 424, 130053. [Google Scholar] [CrossRef]
  19. Yin, R.; Ling, L.; Shang, C. Wavelength-dependent chlorine photolysis and subsequent radical production using UV-LEDs as light sources. Water Res. 2018, 142, 452–458. [Google Scholar] [CrossRef]
  20. Feng, Y.G.; Smith, D.W.; Bolton, J.R. Photolysis of aqueous free chlorine species (NOCl and OCl) with 254 nm ultraviolet light. J. Environ. Eng. Sci. 2007, 6, 277–284. [Google Scholar] [CrossRef]
  21. Cooper, W.J.; Jones, A.C.; Whitehead, R.F.; Zika, R.G. Sunlight-induced photochemical decay of oxidants in natural waters: Implications in ballast water treatment. Environ. Sci. Technol. 2007, 41, 3728–3733. [Google Scholar] [CrossRef] [PubMed]
  22. Miklos, D.B.; Remy, C.; Jekel, M.; Linden, K.G.; Drewes, J.E.; Hübner, U. Evaluation of advanced oxidation processes for water and wastewater treatment—A critical review. Water Res. 2018, 139, 118–131. [Google Scholar] [CrossRef] [PubMed]
  23. Li, W.; Jain, T.; Ishida, K.; Liu, H.Z. A mechanistic understanding of the degradation of trace organic contaminants by UV/hydrogen peroxide, UV/persulfate and UV/free chlorine for water reuse. Environ. Sci. Water Res. Technol. 2017, 3, 128–138. [Google Scholar] [CrossRef]
  24. Magara, Y.; Aizawa, T.; Matumoto, N.; Souna, F. Degradation of pesticides by chlorination during water-purification. Water Sci. Technol. 1994, 30, 119–128. [Google Scholar] [CrossRef]
  25. Deborde, M.; von Gunten, U. Reactions of chlorine with inorganic and organic compounds during water treatment-Kinetics and mechanisms: A critical review. Water Res. 2008, 42, 13–51. [Google Scholar] [CrossRef] [PubMed]
  26. Duirk, S.E.; Bridenstine, D.R.; Leslie, D.C. Reaction of benzophenone UV filters in the presence of aqueous chlorine: Kinetics and chloroform formation. Water Res. 2013, 47, 579–587. [Google Scholar] [CrossRef] [PubMed]
  27. Ma, L.Y.; Li, J.; Xu, L. Aqueous chlorination of fenamic acids: Kinetic study, transformation products identification and toxicity prediction. Chemosphere 2017, 175, 114–122. [Google Scholar] [CrossRef] [PubMed]
  28. Criquet, J.; Rodriguez, E.M.; Allard, S.; Wellauer, S.; Salhi, E.; Joll, C.A.; von Gunten, U. Reaction of bromine and chlorine with phenolic compounds and natural organic matter extracts—Electrophilic aromatic substitution and oxidation. Water Res. 2015, 85, 476–486. [Google Scholar] [CrossRef]
  29. Wang, D.; Hua, Z.C.; Cui, Y.L.; Dong, Z.J.; Li, C.; Fang, J.Y. Probing into the mechanisms of disinfection by-product formation from natural organic matter and model compounds after UV/chlorine treatment. Environ. Sci. Water Res. Technol. 2023, 9, 1587–1598. [Google Scholar] [CrossRef]
  30. Vione, D.; Maurino, V.; Minero, C.; Calza, P.; Pelizzetti, E. Phenol chlorination and photochlorination in the presence of chloride ions in homogeneous aqueous solution. Environ. Sci. Technol. 2005, 39, 5066–5075. [Google Scholar] [CrossRef]
  31. Olaniran, A.O.; Igbinosa, E.O. Chlorophenols and other related derivatives of environmental concern: Properties, distribution and microbial degradation processes. Chemosphere 2011, 83, 1297–1306. [Google Scholar] [CrossRef]
  32. Wennrich, L.; Popp, P.; Möder, M. Determination of chlorophenols in soils using accelerated solvent extraction combined with solid-phase microextraction. Anal. Chem. 2000, 72, 546–551. [Google Scholar] [CrossRef] [PubMed]
  33. Bellar, T.A.; Lichtenberg, J.J.; Kroner, R.C. The Occurrence of Organohalides in Chlorinated Drinking Waters. J. Am. Water Work. Assoc. 1974, 66, 703–706. [Google Scholar] [CrossRef]
  34. Singer, P.C. Control of Disinfection by-Products in Drinking-Water. J. Environ. Eng. 1994, 120, 727–744. [Google Scholar] [CrossRef]
  35. Postigo, C.; Richardson, S.D. Transformation of pharmaceuticals during oxidation/disinfection processes in drinking water treatment. J. Hazard. Mater. 2014, 279, 461–475. [Google Scholar] [CrossRef]
  36. Ma, X.Y.; Deng, J.; Feng, J.; Shanaiah, N.; Smiley, E.; Dietrich, A.M. Identification and characterization of phenylacetonitrile as a nitrogenous disinfection byproduct derived from chlorination of phenylalanine in drinking water. Water Res. 2016, 102, 202–210. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, Y.T.; Qu, D.X.; Bu, L.J.; Zhu, S.M.; Zhou, S.Q. Enhanced trichloronitromethane formation during chlorine-UV treatment of nitrite-containing water by organic amines. Sci. Total Environ. 2022, 853, 158304. [Google Scholar] [CrossRef]
  38. Plewa, M.J.; Wagner, E.D.; Jazwierska, P.; Richardson, S.D.; Chen, P.H.; McKague, A.B. Halonitromethane drinking water disinfection byproducts: Chemical characterization and mammalian cell cytotoxicity and genotoxicity. Environ. Sci. Technol. 2004, 38, 62–68. [Google Scholar] [CrossRef]
  39. Yin, J.B.; Wu, B.; Zhang, X.X.; Xian, Q.M. Comparative toxicity of chloro- and bromo-nitromethanes in mice based on a metabolomic method. Chemosphere 2017, 185, 20–28. [Google Scholar] [CrossRef]
  40. Chen, H.F.; Yin, J.B.; Zhu, M.J.; Cao, C.; Gong, T.T.; Xian, Q.M. Cold on-column injection coupled with gas chromatography/mass spectrometry for determining halonitromethanes in drinking water. Anal. Methods 2016, 8, 362–370. [Google Scholar] [CrossRef]
  41. Wang, J.J.; Li, Z.G.; Hu, S.Y.; Ma, J.; Gong, T.T.; Xian, Q.M. Formation and influence factors of halonitromethanes in chlorination of nitro-aromatic compounds. Chemosphere 2021, 278, 130497. [Google Scholar] [CrossRef] [PubMed]
  42. Deng, L.; Huang, C.H.; Wang, Y.L. Effects of Combined UV and Chlorine Treatment on the Formation of Trichloronitromethane from Amine Precursors. Environ. Sci. Technol. 2014, 48, 2697–2705. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, X.; Shen, Q.Q.; Guo, W.H.; Peng, J.F.; Liang, Y.M. Precursors and nitrogen origins of trichloronitromethane and dichloroacetonitrile during chlorination/chloramination. Chemosphere 2012, 88, 25–32. [Google Scholar] [CrossRef] [PubMed]
  44. Han, J.R.; Zhang, X.R.; Jiang, J.Y.; Li, W.X. How Much of the Total Organic Halogen and Developmental Toxicity of Chlorinated Drinking Water Might Be Attributed to Aromatic Halogenated DBPs? Environ. Sci. Technol. 2021, 55, 5906–5916. [Google Scholar] [CrossRef] [PubMed]
  45. Yang, M.T.; Zhang, X.R. Comparative Developmental Toxicity of New Aromatic Halogenated DBPs in a Chlorinated Saline Sewage Effluent to the Marine Polychaete Platynereis dumerilii. Environ. Sci. Technol. 2013, 47, 10868–10876. [Google Scholar] [CrossRef] [PubMed]
  46. Mitch, W.A.; Richardson, S.D.; Zhang, X.R.; Gonsior, M. High-molecular-weight by-products of chlorine disinfection. Nat. Water 2023, 1, 336–347. [Google Scholar] [CrossRef]
  47. Ye, J.; Singh, A.; Ward, O.P. Biodegradation of nitroaromatics and other nitrogen-containing xenobiotics. World J. Microbiol. Biotechnol. 2004, 20, 117–135. [Google Scholar] [CrossRef]
  48. Podeh, M.R.H.; Bhattacharya, S.K.; Qu, M.B. Effects of nitrophenols on acetate utilizing methanogenic systems. Water Res. 1995, 29, 391–399. [Google Scholar] [CrossRef]
  49. Ahmed, E.; Nagaoka, K.; Fayez, M.; Abdel-Daim, M.M.; Samir, H.; Watanabe, G. Suppressive effects of long-term exposure to P-nitrophenol on gonadal development, hormonal profile with disruption of tissue integrity, and activation of caspase-3 in male Japanese quail (Coturnix japonica). Environ. Sci. Pollut. Res. 2015, 22, 10930–10942. [Google Scholar] [CrossRef]
  50. Di Paola, A.; Augugliaro, V.; Palmisano, L.; Pantaleo, G.; Savinov, E. Heterogeneous photocatalytic degradation of nitrophenols. J. Photochem. Photobiol. A 2003, 155, 207–214. [Google Scholar] [CrossRef]
  51. Daneshvar, N.; Behnajady, M.A.; Asghar, Y.Z. Photooxidative degradation of 4-nitrophenol (4-NP) in UV/H2O2 process: Influence of operational parameters and reaction mechanism. J. Hazard. Mater. 2007, 139, 275–279. [Google Scholar] [CrossRef]
  52. Chen, C.Y.; Du, Y.; Zhou, Y.J.; Wu, Q.Y.; Zheng, S.S.; Fang, J.Y. Formation of nitro(so) and chlorinated products and toxicity alteration during the UV/monochloramine treatment of phenol. Water Res. 2021, 194, 116914. [Google Scholar] [CrossRef] [PubMed]
  53. Wu, Z.H.; Guo, K.H.; Fang, J.Y.; Yang, X.Q.; Xiao, H.; Hou, S.D.; Kong, X.J.; Shang, C.; Yang, X.; Meng, F.A.; et al. Factors affecting the roles of reactive species in the degradation of micropollutants by the UV/chlorine process. Water Res. 2017, 126, 351–360. [Google Scholar] [CrossRef] [PubMed]
  54. Xiang, Y.Y.; Fang, J.Y.; Shang, C. Kinetics and pathways of ibuprofen degradation by the UV/chlorine advanced oxidation process. Water Res. 2016, 90, 301–308. [Google Scholar] [CrossRef] [PubMed]
  55. Gao, Y.Q.; Gao, N.Y.; Chen, J.X.; Zhang, J.; Yin, D.Q. Oxidation of β-blocker atenolol by a combination of UV light and chlorine: Kinetics, degradation pathways and toxicity assessment. Sep. Purif. Technol. 2020, 231, 115927. [Google Scholar] [CrossRef]
  56. Canonica, S.; Meunier, L.; Von Gunten, U. Phototransformation of selected pharmaceuticals during UV treatment of drinking water. Water Res. 2008, 42, 121–128. [Google Scholar] [CrossRef]
  57. USEPA. Standard Methods for the Examination of Water and Wastewater, 20th ed.; American Public Health Association: Washington, DC, USA, 1998.
  58. Bordwell, F.G.; Cheng, J.P. Substituent effects on the stabilities of phenoxyl radicals and the acidities of phenoxyl radical cations. J. Am. Chem. Soc. 1991, 113, 1736–1743. [Google Scholar] [CrossRef]
  59. Guo, K.H.; Wu, Z.H.; Shang, C.; Yao, B.; Hou, S.D.; Yang, X.; Song, W.H.; Fang, J.Y. Radical Chemistry and Structural Relationships of PPCP Degradation by UV/Chlorine Treatment in Simulated Drinking Water. Environ. Sci. Technol. 2017, 51, 10431–10439. [Google Scholar] [CrossRef]
  60. Gilbert, B.C.; Stell, J.K.; Peet, W.J.; Radford, K.J. Generation and Reactions of the Chlorine Atom in Aqueous Solution. J. Chem. Soc. Faraday Trans. 1 1988, 84, 3319–3330. [Google Scholar] [CrossRef]
  61. Negreira, N.; Regueiro, J.; de Alda, M.L.; Barceló, D. Transformation of tamoxifen and its major metabolites during water chlorination: Identification and in silico toxicity assessment of their disinfection byproducts. Water Res. 2015, 85, 199–207. [Google Scholar] [CrossRef]
  62. Nika, M.C.; Bletsou, A.A.; Koumaki, E.; Noutsopoulos, C.; Mamais, D.; Stasinakis, A.S.; Thomaidis, N.S. Chlorination of benzothiazoles and benzotriazoles and transformation products identification by LC-HR-MS/MS. J. Hazard. Mater. 2017, 323, 400–413. [Google Scholar] [CrossRef] [PubMed]
  63. Merlet, N.; Thibaud, H.; Dore, M. Chloropicrin formation during oxidative treatments in the preparation of drinking-water. Sci. Total Environ. 1985, 47, 223–228. [Google Scholar] [CrossRef]
  64. Yuan, R.X.; Ramjaun, S.N.; Wang, Z.H.; Liu, J.S. Effects of chloride ion on degradation of Acid Orange 7 by sulfate radical-based advanced oxidation process: Implications for formation of chlorinated aromatic compounds. J. Hazard. Mater. 2011, 196, 173–179. [Google Scholar] [CrossRef] [PubMed]
  65. Li, J.; Jiang, J.; Manasfi, T.; von Gunten, U. Chlorination and bromination of olefins: Kinetic and mechanistic aspects. Water Res. 2020, 187, 116424. [Google Scholar] [CrossRef] [PubMed]
  66. Zhao, X.L.; Li, Y.Y.; Lu, J.H.; Zhou, L.; Chovelon, J.M.; Zhou, Q.S.; Ji, Y.F. UV/H2O2 oxidation of chloronitrobenzenes in waters revisited: Hydroxyl radical induced self-nitration. J. Photochem. Photobiol. A 2021, 410, 113162. [Google Scholar] [CrossRef]
  67. Brillas, E.; Garcia-Segura, S. Benchmarking recent advances and innovative technology approaches of Fenton, photo-Fenton, electro-Fenton, and related processes: A review on the relevance of phenol as model molecule. Sep. Purif. Technol. 2020, 237, 116337. [Google Scholar] [CrossRef]
  68. Dzengel, J.; Theurich, J.; Bahnemann, D.W. Formation of nitroaromatic compounds in advanced oxidation processes: Photolysis versus photocatalysis. Environ. Sci. Technol. 1999, 33, 294–300. [Google Scholar] [CrossRef]
  69. Wu, Y.T.; Bu, L.J.; Duan, X.D.; Zhu, S.M.; Kong, M.H.; Zhu, N.Y.; Zhou, S.Q. Mini review on the roles of nitrate/nitrite in advanced oxidation processes: Radicals transformation and products formation. J. Clean. Prod. 2020, 273, 123065. [Google Scholar] [CrossRef]
  70. Zhao, X.L.; Zhang, T.; Lu, J.H.; Zhou, L.; Chovelon, J.M.; Ji, Y.F. Formation of chloronitrophenols upon sulfate radical-based oxidation of 2-chlorophenol in the presence of nitrite. Environ. Pollut. 2020, 261, 114242. [Google Scholar] [CrossRef]
  71. Augusto, O.; Bonini, M.G.; Amanso, A.M.; Linares, E.; Santos, C.C.X.; De Menezes, S.L. Nitrogen dioxide and carbonate radical anion: Two emerging radicals in biology. Free Radic. Biol. Med. 2002, 32, 841–859. [Google Scholar] [CrossRef]
  72. Ji, Y.F.; Shi, Y.Y.; Yang, Y.; Yang, P.Z.; Wang, L.; Lu, J.H.; Li, J.H.; Zhou, L.; Ferronato, C.; Chovelon, J.M. Rethinking sulfate radical-based oxidation of nitrophenols: Formation of toxic polynitrophenols, nitrated biphenyls and diphenyl ethers. J. Hazard. Mater. 2019, 361, 152–161. [Google Scholar] [CrossRef]
  73. Vione, D.; Maurino, V.; Minero, C.; Pelizzetti, E. Aqueous atmospheric chemistry: Formation of 2,4-dinitrophenol upon nitration of 2-nitrophenol and 4-nitrophenol in solution. Environ. Sci. Technol. 2005, 39, 7921–7931. [Google Scholar] [CrossRef] [PubMed]
  74. Keen, O.S.; Love, N.G.; Linden, K.G. The role of effluent nitrate in trace organic chemical oxidation during UV disinfection. Water Res. 2012, 46, 5224–5234. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, Y.F.; Roddick, F.A.; Fan, L.H. Direct and indirect photolysis of seven micropollutants in secondary effluent from a wastewater lagoon. Chemosphere 2017, 185, 297–308. [Google Scholar] [CrossRef]
  76. Guo, K.H.; Wu, Z.H.; Yan, S.W.; Yao, B.; Song, W.H.; Hua, Z.C.; Zhang, X.W.; Kong, X.J.; Li, X.C.; Fang, J.Y. Comparison of the UV/chlorine and UV/H2O2 processes in the degradation of PPCPs in simulated drinking water and wastewater: Kinetics, radical mechanism and energy requirements. Water Res. 2018, 147, 184–194. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, Y.R.; dos Santos, M.M.; Ding, X.X.; Labanowski, J.; Gombert, B.; Snyder, S.A.; Croué, J.P. Impact of EfOM in the elimination of PPCPs by UV/chlorine: Radical chemistry and toxicity bioassays. Water Res. 2021, 204, 117634. [Google Scholar] [CrossRef]
  78. Wang, W.L.; Wu, Q.Y.; Huang, N.; Wang, T.; Hu, H.Y. Synergistic effect between UV and chlorine (UV/chlorine) on the degradation of carbamazepine: Influence factors and radical species. Water Res. 2016, 98, 190–198. [Google Scholar] [CrossRef]
  79. APHA/AWWA/WEF. Standard Methods for the Examination of Water and Wastewater, 21st ed.; American Public Health Association/American Water Works Association/Water Environment Federation: Washington, DC, USA, 2005. [Google Scholar]
Figure 1. (a) The degradation of 4-nitrophenol by UV, HOCl, and UV/HOCl processes. Note that 5 mM TBA was added to UV/HOCl system as a radical quencher. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. (b) The effect of HOCl dosage on the degradation of 4-nitrophenol by UV/HOCl process. The concentrations of HOCl increased from 40 to 100 μM and the other conditions were the same as above. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations.
Figure 1. (a) The degradation of 4-nitrophenol by UV, HOCl, and UV/HOCl processes. Note that 5 mM TBA was added to UV/HOCl system as a radical quencher. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. (b) The effect of HOCl dosage on the degradation of 4-nitrophenol by UV/HOCl process. The concentrations of HOCl increased from 40 to 100 μM and the other conditions were the same as above. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations.
Water 15 04038 g001
Figure 2. (a) Formation of chlorinated nitrophenol intermediates (including 2-chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol) in the chlorination of 4-nitrophenol; (b) Formation of 2-chloro-4-nitrophenol in the UV/HOCl process. Note that 5 mM TBA was spiked into the reaction solution as a radical quencher. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations.
Figure 2. (a) Formation of chlorinated nitrophenol intermediates (including 2-chloro-4-nitrophenol and 2,6-dichloro-4-nitrophenol) in the chlorination of 4-nitrophenol; (b) Formation of 2-chloro-4-nitrophenol in the UV/HOCl process. Note that 5 mM TBA was spiked into the reaction solution as a radical quencher. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations.
Water 15 04038 g002
Figure 3. Effect of HOCl dosage on the formation of 2-chloro-4-nitrophenol during the oxidation of 4-nitrophenol by UV/HOCl process. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 40−100 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations. Symbols represent experimental data, and lines are kinetic model fitting with Equation (13).
Figure 3. Effect of HOCl dosage on the formation of 2-chloro-4-nitrophenol during the oxidation of 4-nitrophenol by UV/HOCl process. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 40−100 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations. Symbols represent experimental data, and lines are kinetic model fitting with Equation (13).
Water 15 04038 g003
Figure 4. (a) Formation of TCNM in the HOCl and UV/HOCl processes. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. (b) Effect of HOCl dosage on the formation of TCNM during the oxidation of 4-nitrophenol by UV/HOCl. The concentration of HOCl increased from 40 to 100 μM and the other conditions were the same as above.
Figure 4. (a) Formation of TCNM in the HOCl and UV/HOCl processes. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL, solution pH was adjusted by phosphate buffer. (b) Effect of HOCl dosage on the formation of TCNM during the oxidation of 4-nitrophenol by UV/HOCl. The concentration of HOCl increased from 40 to 100 μM and the other conditions were the same as above.
Water 15 04038 g004
Scheme 1. Transformation pathways of 4-nitrophenol by UV/HOCl process. Note that the molecular structures in the square brackets (R1, 1,2,4-trihydroxybenzene and R2, hydroquinone) were the reactive intermediates not identified in the experiments but were proposed to be involved in the transformation of 4-nitrophenol. P1-P5 were the intermediate products of 4-nitrophenol identified by HPLC and/or HRMS analyses. Nomenclature: P1, 2-chloro-4-nitrophenol; P2, 2,6-dichloro-4-nitrophenol; P3, 2,4-dinitrophenol; P4, 2-chloro-4,6-dinitrophenol; and P5, 4-nitrocatechol.
Scheme 1. Transformation pathways of 4-nitrophenol by UV/HOCl process. Note that the molecular structures in the square brackets (R1, 1,2,4-trihydroxybenzene and R2, hydroquinone) were the reactive intermediates not identified in the experiments but were proposed to be involved in the transformation of 4-nitrophenol. P1-P5 were the intermediate products of 4-nitrophenol identified by HPLC and/or HRMS analyses. Nomenclature: P1, 2-chloro-4-nitrophenol; P2, 2,6-dichloro-4-nitrophenol; P3, 2,4-dinitrophenol; P4, 2-chloro-4,6-dinitrophenol; and P5, 4-nitrocatechol.
Water 15 04038 sch001
Figure 5. (a) Degradation of 4-nitrophenol by HOCl and UV/HOCl processes in wastewater and Milli-Q water. (b) Effects of wastewater matrix on the formation of 2-chloro-4-nitrophenol by HOCl and UV/HOCl processes. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations.
Figure 5. (a) Degradation of 4-nitrophenol by HOCl and UV/HOCl processes in wastewater and Milli-Q water. (b) Effects of wastewater matrix on the formation of 2-chloro-4-nitrophenol by HOCl and UV/HOCl processes. Experimental conditions: [4-Nitrophenol]0 = 10 μM, [HOCl]0 = 60 μM, solution volume = 50 mL. The experimental data is the average of the two parallel experiments and the error bars (when not visible, smaller than the symbols) represent the standard deviations.
Water 15 04038 g005
Table 1. k1 and k2 values obtained from the experimental data fitted by the kinetic model at different HOCl dosage.
Table 1. k1 and k2 values obtained from the experimental data fitted by the kinetic model at different HOCl dosage.
Concentration of HOCl (μM)k1 (min−1)k2 (min−1)
401.8010.024
601.8720.102
800.8680.271
1001.4090.201
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, X.; Cai, Y.; Lu, J.; Chovelon, J.-M.; Chen, J.; Jiang, C.; Ji, Y. Abatement of Nitrophenol in Aqueous Solution by HOCl and UV/HOCl Processes: Kinetics, Mechanisms, and Formation of Chlorinated Nitrogenous Byproducts. Water 2023, 15, 4038. https://doi.org/10.3390/w15234038

AMA Style

Li X, Cai Y, Lu J, Chovelon J-M, Chen J, Jiang C, Ji Y. Abatement of Nitrophenol in Aqueous Solution by HOCl and UV/HOCl Processes: Kinetics, Mechanisms, and Formation of Chlorinated Nitrogenous Byproducts. Water. 2023; 15(23):4038. https://doi.org/10.3390/w15234038

Chicago/Turabian Style

Li, Xiaoci, Yan Cai, Junhe Lu, Jean-Marc Chovelon, Jing Chen, Canlan Jiang, and Yuefei Ji. 2023. "Abatement of Nitrophenol in Aqueous Solution by HOCl and UV/HOCl Processes: Kinetics, Mechanisms, and Formation of Chlorinated Nitrogenous Byproducts" Water 15, no. 23: 4038. https://doi.org/10.3390/w15234038

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop