You are on page 1of 15

Environmental Pollution 281 (2021) 116950

Contents lists available at ScienceDirect

Environmental Pollution
journal homepage: www.elsevier.com/locate/envpol

Review

Occurrence, influencing factors, toxicity, regulations, and abatement


approaches for disinfection by-products in chlorinated drinking
water: A comprehensive review
Sundas Kali a, Marina Khan a, Muhammad Sheraz Ghaffar a, Sajida Rasheed b,
Amir Waseem c, Muhammad Mazhar Iqbal d, e, Muhammad Bilal khan Niazi f,
Mazhar Iqbal Zafar a, *
a
Department of Environmental Sciences, Faculty of Biological Sciences, Quaid-i-Azam University, Islamabad, 45320, Pakistan
b
Department of Biotechnology, Faculty of Sciences, University of Kotli, Azad Jamu Kashmir, Pakistan
c
Department of Chemistry, Quaid-i-Azam University, Islamabad, 45320, Pakistan
d
Laboratory of Analytical Chemistry and Applied Eco-chemistry, Department of Applied Analytical and Physical Chemistry, Ghent University, Ghent,
Belgium
e
Soil and Water Testing Laboratory, Department of Agriculture, Chiniot, Government of Punjab, Pakistan
f
Department of Chemical Engineering, National University of Sciences and Technology, Islamabad, Pakistan

a r t i c l e i n f o a b s t r a c t

Article history: Disinfection is considered as a vital step to ensure the supply of clean and safe drinking water. Various
Received 24 December 2020 approaches are adopted for this purpose; however, chlorination is highly preferred all over the world.
Received in revised form This method is opted owing to its several advantages. However, it leads to the formation of certain by-
9 March 2021
products. These chlorination disinfection by-products (DBPs) are genotoxic, carcinogenic and mutagenic.
Accepted 12 March 2021
Still chlorination is being practiced worldwide. Present review gives insights into the occurrence, toxicity
Available online 17 March 2021
and factors affecting the formation of regulated (THMs, HAAs) and emerging DBPs (N-DBPs, HKs, HAs and
aromatic DBPs) found in drinking water. Furthermore, remediation techniques used to control DBPs have
Keywords:
Chlorination
also been summarized here. Key findings are: (i) concentration of regulated DBPs surpassed the
Drinking water permissible limit in most of the regions, (ii) high chlorine dose, high NOM, more reaction time (up to 3 h)
Disinfection by-products and high temperature (up to 30  C) enhance the formation of THMs and HAAs, (iii) high pH favors the
Occurrence formation of THMs while low pH is suitable of the formation of HAAs, (iv) high NOM, low temperature,
Influencing factors low chlorine dose and moderate pH favors the formation of unstable DBPs (N-DBPs, HKs and HAs), (v)
Toxicity DBPs are toxic not only for humans but for aquatic fauna as well, (vi) membrane technologies, enhanced
Abatement approaches coagulation and AOPs remove NOM, (vii) adsorption, air stripping and other physical and chemical
methods are post-formation approaches (viii) step-wise chlorination is assumed to be an efficient
method to reduce DBPs formation without any treatment. Toxicity data revealed that N-DBPs are found
to be more toxic than C-DBPs and aromatic DBPs than aliphatic DBPs. In majority of the studies, merely
THMs and HAAs have been studied and USEPA has regulated just these two groups. Future studies should
focus on emerging DBPs and provide information regarding their regulation.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction

Disinfection is a crucial step for the treatment of water for public


supply (How et al., 2017). Several factors like population explosion,
* Corresponding author. climate change, industrialization and agriculture have posed
E-mail addresses: sundaschoudhary93@gmail.com (S. Kali), considerable stress on water resources (Du et al., 2017; Sun et al.,
marinakhanktk1996@gmail.com (M. Khan), sherazkhan5505@gmail.com
2016). Such activities have increased the demand of fresh water
(M.S. Ghaffar), rsajida142@gmail.com (S. Rasheed), amir@qau.edu.pk
(A. Waseem), mazhar1621@gmail.com (M.M. Iqbal), M.b.k.niazi@scme.nust.edu.pk and discharge of wastewater into the environment hence polluting
(M. Bilal khan Niazi), mzafar@qau.edu.pk (M.I. Zafar). the clean water resources (Da Costa et al., 2014; Sun et al., 2016).

https://doi.org/10.1016/j.envpol.2021.116950
0269-7491/© 2021 Elsevier Ltd. All rights reserved.
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

List of abbreviations HOBr Hypobroumous acid


HOI Hypoiodous acid
DBPs Disinfection by-products GI Galvanized iron
THMs Trihalomethanes SS Stainless steel
HAAs Haloacetic acids GS Galvanized steel
N-DBPs Nitrogenous disinfection by-products HDPE High density polyethylene
HANs Haloacetonitriles PVC Polyvinyl chloride
HKs Haloketones PP Polypropylene
HAs Haloaldehydes PE Polyethylene
C-DBPs Carbonaceous disinfection byproducts DI Ductile iron
HPANs Halophenylacetonitriles CHO Chinese hamster ovary
HNPs Halonitrophenols AOPs Advanced oxidation processes
HPs Halophenols GAC Granular activated carbon
HBADs Halobenzaldehyde USEPA US Environmental Protection Agency
HBACs Halohydroxybenzoic acids WHO World Health Organization
HBQs Halobenzoquinones IARC International Agency for Research on Cancer
NOM Natural organic matter GDP Gross domestic production
HOCl Hypochlorus acid NTU Nephelometric turbidity unit
OCl Hypochlorite ion

According to World Health Organization (WHO), unsafe drinking maintaining a free chlorine residual of 0.2 mg L1 in stored
water is responsible for nearly 80% of human diseases in developing household water treatment by chlorination (WHO, 2017).
countries (Sulehria et al., 2013). In Pakistan, approximately 80% of Gaseous chlorine, sodium hypochlorite and calcium hypochlo-
whole population has to rely on poor quality drinking water due to rite are used for chlorination purposes (see Table 1). Hydrolysis of
the scarcity of safe drinking water resources (Daud et al., 2017). these disinfectants produces hypochlorous acid (HOCl) and hypo-
Moreover, 30% of all diseases and 40% of all deaths in Pakistan chlorite ion (OCl) which are defined as free chlorine (Yang et al.,
occurred on account of unsafe drinking water (Nabeela et al., 2014). 2011). In addition to these, monochloramine (NH2Cl) and dichlor-
Literature reported that about 0.6e1.44% of country’s Gross Do- amine (NHCl2) are also found which is called combined chlorine.
mestic Product (GDP) loss is associated with water borne diseases However, several factors like pH, temperature etc. affect the
(Khwaja and Aslam, 2018; Tahir et al., 2010). Pakistan ranked 80th dominant form of chlorine which ultimately describes the disin-
from 122 countries with respect to drinking water quality (Hisam fection process (Voukkali and Zorpas, 2015). The concentrations of
et al., 2014) as its 70% of surface and subsurface water is unsafe hypochlorous acid and the hypochlorite ion are approximately
for drinking (Abbas et al., 2014; Baig et al., 2011). Therefore, equal at pH 7.5 and 25  C (De la Guardia and Garrigues, 2015).
reclaimed water is a useful alternative for socio-economic benefits. Advantage of using chlorine is that it removes almost all kinds of
Due to the presence of pathogens, reclaimed water needs to be pathogens, additionally, residual chlorine also prevents microbial
disinfected before its use for drinking purpose, agriculture, ground recontamination in water supply systems (Hassan et al., 2010).
water recharge or discharge is water bodies (Du et al., 2017; Li et al., Nevertheless, the emergence of carcinogenic, mutagenic and gen-
2013). otoxic disinfectant by products (DBPs) along with residual chlorine
Various methods are employed for water disinfection against is the major drawback of this technique (How et al., 2017; Da Costa
various pathogens; bacteria, viruses, protozoa and fungi (Ding et al., et al., 2014). When chlorine reacts with organic matter content,
2019; Da Costa et al., 2014) such as; chlorine (Scarlet et al., 2016; bromine and other contaminants present in water, it oxidizes them
Pardakhti et al., 2011), chloramine (How et al., 2017; Moradi et al., and produces DBPs (Abbas et al., 2014). Generally, surface waters
2017), chlorine dioxide (Van Haute et al., 2017; Scarlet et al., contain elevated concentrations of DBPs than ground water
2016), peracetic acid (Da Costa et al., 2014), ozone (Ding et al., because of the greater concentrations of organic material (Abbas
2019; Zheng et al., 2017) and ultra-violet (UV) radiations (Zheng et al., 2014; Nieuwenhuijsen et al., 2000). Researchers have
et al., 2017; Scarlet et al., 2016). blown the whistle on the presence of DBPs in disinfected drinking
Chlorination is a traditional and highly preferred method due to water. Currently, more than 700 DBPs have been reported in dis-
its simple operation, effective oxidizing potential, and above all, infected water (Plewa et al., 2017)and many of them are reported as
cost effectiveness (Chowdhury et al., 2012; Pardakhti et al., 2011; carcinogenic, mutagenic, teratogenic, cytotoxic, and genotoxic,
Karim et al., 2011). It is the only trusted mechanism that developing (Benson et al., 2017). Well known classes of aliphatic DBPs include;
countries can count to disinfect water (Mahmood, 2015). According trihalomethanes (THMs), haloacetic acids (HAAs), haloketones
to the WHO, residual chlorine levels should be kept equal or higher (HKs), haloaldehydes (HAs), and nitrogenous DBPs (N-DBPs) such
than 0.5 mg L1 after contact time of 30 min at pH below 8 and as haloacetonitriles (HANs), halonitromethanes (HNMs) and hal-
temperature 20ᵒC (Mahmood, 2015). It is critical to maintain oacetamides (HAcAms) (Benson et al., 2017; Bond et al., 2011).
appropriate dosing amount of chlorine in order to provide enough However, aromatic DBPs have also become an area of interest in the
residual chlorine during storage and utilization. Suggestions are to field of DBPs research which include phenolic and heterocyclic
dose clear water (<10 nephelometric turbidity units [NTU]) with aromatic DBPs (Fig. 1).
free chlorine at approximately 2 mg L1 and twice (4 mg L1) to The objective of the present review is to analyze the outcomes of
turbid water (>10 NTU). Even though these free chlorine dosages various studies and summarize the occurrence, influencing factors,
may leash to chlorine residues that goes beyond the recommended toxicity, regulations and remediation techniques for important
chlorine residual for water which is centrally dosed at the point of halogenated aliphatic and aromatic DBPs formed by the chlorina-
delivery, 0.2_0.5 mg L1, these doses are considered appropriate for tion of surface and ground water. Moreover, the concentrations of

2
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Table 1
Disinfectant and their by-products formation in drinking Water (WHO, 2017).

Disinfectant DBPs

Organohalogen Products Inorganic Non-halogenated


Products Products

Chlorine (Sodium hypochlorite, Calcium THMs, HAAs, HANs, HKs, HAs, N-chloramines, Chlorophenols, Chloropicrin, Chloral hydrates, Chlorate Aldehydes,
hypochlorite) Halo furanones, Bromohydrins Carboxylic acids,
Benzene,
Alkanoic acid,
Cyano alkanoic acid
Chlorine dioxide Chlorite, Unknown
Chlorate

3. Occurrence of DBPs in drinking water

Highly volatile THMs is the major group of DBPs followed by


non-volatile halo acetic acids (HAAs), occurring generally at around
half the concentrations of THMs (Nieuwenhuijsen et al., 2000).
Both THMs and HAAs constitute about 25% of the total DBPs (Yang
et al., 2011). The chlorine atom in hypochlorous acid act as an
electrophile, that causes electrophilic substitution (i.e. halogen
substitution) for organic compounds (Jiang et al., 2020). Further,
aromatic DBPs are also formed by the reaction of disinfectant with
organic matter. They are highly unstable compounds which
degrade into THMs and HAAs (Hu et al., 2019). Similarly, N-DBPs,
HKs and HAs also decompose into relatively stable DBPs. World-
wide reported concentrations of DBPs in drinking water has been
presented in Table 2.

Fig. 1. Classification of DBPs reviewed in present study (shaded area shows that few
3.1. Aliphatic DBPs
classes of C-DBPs are regulated while rest of them are unregulated).
Among aliphatic DBPs, THMs, HAAs, N-DBPs, HKs and HAs are
commonly found in chlorinated drinking water (Table 3).
these DBPs in various regions of the world have also been critically
analyzed with respect to the international regulations. 3.1.1. Trihalomethanes
THMs is the most prevalent group among DBPs. They are formed
2. Methodology by the reaction of chloramine and chlorine with the inorganic
species and organic matter in water (Benson et al., 2017). These are
A comprehensive literature review was performed for infor- skin permeable and highly volatile compounds. THM4 includes four
mation regarding DBPs and findings have been summarized here. compounds (i) chloroform (CF) (ii) bromoform (BF) (iii) bromodi-
Databases like Science Direct and Google Scholar were used for chloromethane (BDCM) and (iv) dibromochloromethane (DBCM)
literature hunting. Articles were searched using key words like (see Table 3), which have been regulated by EU (100mgL1), US
“disinfection methods”, “chlorination”, “drinking water distribu- (80mgL1), and some other countries (Villanueva et al., 2016; Abbas
tion systems”, “disinfection by-products in chlorinated drinking et al., 2014). They possess different physiochemical and toxicolog-
water”, “natural organic matter” “regulations of DBPs”, “factors ical properties. Among these, chloroform is highly volatile, while
affecting the formation of DBPs”, “trihalomethanes”, “haloacetic rest of the brominated THMs are more lipophilic (Villanueva et al.,
acids”, “haloacetonitriles”, “haloketones”, “haloaldehydes”, “hal- 2016). They are not mutagenic but have carcinogenic and other
onitromethanes”, “halocacetamides”, “aromatic DBPs”, “heterocy- severe toxic effects in humans (Villanueva et al., 2015; Abbas et al.,
clic DBPs”, “phenyl DBPs”, N-DBPs”, “C-DBPs”, “remediation of 2014). THMs are produced during the hydrolysis reaction of un-
aliphatic and aromatic DBPs”, “control of DBPs”, “pipe material and stable DBPs (Hong et al., 2013).
DBPs formation”, “toxicity of DBPs”, “health impacts”, “DBPs in
municipal drinking water” etc. Gathered research articles were 3.1.2. Haloacetic acids
screened out and 86 studies which reported: (i) various disinfection HAAs include monochloroacetic acid (MCAA), dichloroacetic
techniques, (ii) chlorination conditions, (iii) the levels of DBPs in acid (DCAA), trichloroacetic acid (TCAA), monobromoacetic acid
drinking water, (iv) toxicity of DBPs, (v) effect of NOM and other (MBAA), dibromoacetic acid (DBAA), tribromoacetic acid (TBAA),
factors on DBPs formation, and (vi) abatement technologies for bromochloroacetic acid (BCAA), bromodichloroacetic acid (BDCAA)
NOM and DBPs removal, were selected. Further, websites of United and dibromochloroacetic acid (DBCAA) (Tian et al., 2017). Among
States Environmental Protection Agency (US-EPA), WHO and these, first five (HAA5) are more common and produced by the
Council of European Union and International Agency for Research chlorination of drinking water (Benson et al., 2017; Richardson and
on Cancer (IARC) were consulted for DBPs related information. Postigo, 2011). These compounds are known to cause bladder

3
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Table 2
Levels of DBPs reported in drinking water of various countries.

Country Source Water THMs (4) HAAs (5) HANs(4) HKs HAs HNMs HAcAms HBADs HBACs HNPs CPANs References
Name
mg L1 ng L1

Hungary Treatment plants 0.138e458 0.16e136 Srivastav et al. (2020)


South China DWTP 2.8 1e9.8 2.9 Zhang et al. (2020)
e6.1 e8.3
Croatia DWDS 0.7e32.8 LOQ e 17.2 Kurajica et al. (2020)
Iran Surface water 92.9 ± 43.7a 70.6 ± 26.5a Dobaradaran et al. (2020)
Spain WTP 15.9e55.2 N.D. e 0.98 Dominguez-Tello et al.
3.5 e1.02 (2020)
Hungary DWTPs 14.2e143 5.2e129 <1.0e17 n et al. (2019)
Stefa
China DWTPs 0.6 Zhang et al. (2019)
e155
China WTPs 3.26e69.28 3.48e82.18 0.83 0.33 Yu et al. (2019)
e26.85 e14.28
India Sea water 38.3a 9.8 Padhi et al. (2019)
River Water 18.8a 12.8
Reservior 21.5a 20.6
Thailand WDNS 584a 30b Ratpukdi et al. (2019)
China DWTPs 170 Zhang et al. (2018)
e530
China Surface water 35e228.4 188e394.5 2.98 1.03 Hao et al. (2017)
e85.5 e17.9
China Surface water 1.12e8.20 Niu et al. (2017)
Saudi Arabia Desalinated Water 0.1e33.6 Chowdhury and Chowdhury
Blended Water 2.1e52.4 (2017)
Bahrain Desalinated Water 0.27e6.4
Kuwait Desalinated Water 7.11
e104.89
Qatar Desalinated Water 4.02e83.31
UAE Desalinated Water 7.0e15.0
China Yuqiao reservoir 79.4e136.9 133.2 8.2e9.3 Zhai et al. (2017)
e250.9
Nigeria Raw water 0 Benson et al. (2017)
Primary water 21.92
e999.64
Secondary water 26.89
e960.68
China Water distribution 10e30 N.D. e N.D. e 0.1 Huang et al. (2017)
systems 6.4 5.5 e11.4
Qatar Reservoirs 0e85 Al-Otoum et al. (2016)
Desalination plant 0e84
Consumer taps 0e89
China DWTPs 100 Mao et al. (2016)
e150
Cyprus Surface water 21.23e57.3 Ioannou et al. (2016)
Surface water 8.72e86.2
a a
United Disinfected water 2.8* 0.2 1.3 Bond et al. (2015)
Kingdom Final water 2.6a 0.2a 1a
Distribution samples 2.8a 0.2a 1.4a
Pakistan Surface water 24.42 Abbas et al., 2015
e595.86
Groundwater 189.97
e431.26
a
Average value; bmaximum value
N.D. (not defined); DWTP: Drinking water treatment plant; WTP: Wastewater treatment plant.
DWDS: Drinking water distribution system; WDNS: Water distribution network system.

cancer and other adverse reproductive anomalies. Therefore, US only 4 compounds are found in drinking water which include
EPA has set the maximum contamination level of HAA5 at 60 mgL1 dichloroacetonitrile (DCAN), trichloroacetonitrile (TCAN), bromo-
(Tian et al., 2017). Natural organic matter is responsible for 80% of chloroaceontrile (BCAN) and dibromoacetonitrile (DBAN) (Hang
HAAs formation, while rest of HAAs are formed by the reaction of et al., 2016). However, little is known about the presence of HAs in
intermediate DBPs with chlorine (Bougeard et al., 2010). drinking water (Mao et al., 2016; Jeong et al., 2015).
These intermediate DBPs are usually present in lower amounts
than THMs and HAAs. These compounds are usually produced
3.1.3. Other aliphatic DBPs instantly during water disinfection but are decomposed rapidly
HANs, HNMs, HacAms, HKs and HAs are unstable DBPs. They are during reactions with residual disinfectants or hydrolysis reactions
formed in almost all chlorinated water, but for short time. They are n et al., 2019). Presence of ammonia in water favors the for-
(Stefa
also known as intermediate DBPs as they get hydrolyzed at higher mation of N-DBPs (Ding et al., 2018). Nevertheless, HKs and HAs are
pH and temperature (Hong et al., 2013; Fang et al., 2010). Details of formed by the reaction of chlorine with NOM. It has been reported
these compounds have been presented in Table 3. Among HANs,
4
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Table 3
Toxicity summary of aliphatic DBPs.

Group Name Compound names *LC50 mol General Toxicity References


L1
Regulated Aliphatic DBPs
THMs (4) Chloroform 9.62  103 Carcinogenic, Cytotoxic, Genotoxic, Reproductive anomalies, Birth Ewaid et al. (2018); Stalter et al.
Chlorodibromomethane 5.36  103 defects, Kidney and liver damage, Nervous system damage (2020); Wagner and Plewa (2017)
Bromodichloromethane 1.15  102
Bromoform 3.96  103
HAA (5) Monochloroacetic Acid 8.10  104 Carcinogenic, Reproductive anomalies, Growth retardation, Genotoxic, Mompremier et al. (2019); Stalter
Dichloroacetic Acid 7.3  103 Cytotoxic, Spleen, liver and kidney damage et al. (2020); Plewa et al. (2010)
Trichloroacetic Acid 2.4  103
Monobromoacetic acid 9.60  106
Dibromoacetic acid 5.9  104
Unregulated Aliphatic DBPs
HANs Chloroacetonitrile 6.83  105 Carcinogenic, Mutagenic, Clastogenic, Liver damage Postigo et al., 2018; Zhang et al.
Dichloroacetonitrile 5.73  105 (2017); Muellner et al. (2007)
Trichloroacetonitrile 1.60  104
Bromochloracetonitrile 8.46  106
Bromoacetonitrile 3.21  106
Dibromoacetonitrile 2.85  106
Iodoacetonitrile 3.3.  106
HNMs Bromonitromethane 7.06  10⁻6 DNA damage, cytotoxic, mutagenic, oxidative stress Dominguez-Tello et al. (2020); Yin
Chloronitromethane 5.29  10⁻4 et al. (2018); Plewa et al. (2004)
Dibromochloronitromethane 6088  10⁻6
Dichloronitromethane 3.73  10⁻4
Dibromonitromethane 6.09  10⁻6
Bromochloronitromethane 4.05  10⁻5
Bromodichloronitromethane 1.32  10⁻5
Trichloronitromethane 5.36  10⁻4
Tribromonitromethane 8057  10⁻6
HAcAms Chloroacetamide 1.48  10⁻4 Cytotoxic, Genotoxic, developmental toxicity Chu et al. (2012)
Dichloroacetamide 1.92  10⁻3 Plewa et al. (2008); Ding et al. (2020)
Trichloroacetamide 2.05  10⁻3
Bromoacetamide 1.89  10⁻6
Dibromoacetamide 1.22  10⁻5
Tribromoacetamide 3.14  10⁻6
Bromochloroacetamide 1.71  10⁻6
Bromodichloroacetamide 8.68  10⁻6
Dibromochloroacetamide 4.75  10⁻6
Iodoacetamide 1.42  10⁻6
Diiodoacetamide 6.78  10⁻7
Chloroiodoacetamide 5.97  10⁻6
Bromoiodoacetamide 3.81  10⁻6
HKs 1,1-dichloro-propanone e Mutagenic, Oxidative stress Li et al. (2020); Stalter et al. (2020)
1,1,1-trichloro-propanone e
HAs Chloroacetaldehyde 3.51  10⁻6 Mutagenic, Nephrotoxic, Cytotoxic, Chaves et al. (2019); Jeong (2014)
Bromoacetaldehyde 1.728  10⁻5
Dichloroacetaldehyde 2.925  10⁻5
Bromochloroacetaldehyde 5.34  10⁻6
Dibromoacetaldehyde 4.7  10⁻6
Chloral hydrate 1.163  10⁻3
Dibromochloroacetaldehyde 5.15  10⁻6
Bromodichloroacetaldehyde 2.035  10⁻5
Iodoacetaldehyde 6  10⁻6
Tribromoacetaldehyde 3.58  10⁻6

*LC50 values included that induced a cell density of 50% as compared to the concurrent negative controls by Chinese hamster ovary cells bioassay with the exposure time of 72
hours.
THMs: trihalomethanes; HAAs: Haloacetic acids; HANs: Haloacetonitriles; HNMs:Halonitromethanes.
HAcAms: Haloacetamides; HKs:Haloketones; HAs:Haloaldehydes.

that HANs and HAs are more toxic than THMs and HAAs (Mao et al., Zhang, 2016). Therefore, focus of the research has been diverted
2016; Jeong et al., 2015; Muellner et al., 2007). Although HAcAms to investigate the unknown classes of emerging DBPs including
are also found in low concentration, yet their toxicity is 142, 2 and aromatic DBPs. Humic fraction of NOM contains number of aro-
1.4 times higher than HAAs (5), HANs and HNMs, respectively (Ding matic compounds which are responsible for the formation of aro-
et al., 2018). matic DBPs (Jiang et al., 2020). Halofuranones is the very first group
of aromatic DBPs. Now, more than 100 aromatic DBPs have been
found in disinfected water (Liu et al., 2020a).
3.2. Aromatic DBPs
Aromatic DBPs can be classified as phenyl and heterocyclic DBPs.
Phenyl DBPs contain eight subgroups including halopeptides, hal-
Although regulated DBPs are found in higher concentration, still
ophenols (HPs), halobenzoquinones (HBQs), halophenylacetonitrile
their contribution towards overall toxicity of chlorinated water is
(HPANs), halohydroxybenzaldehydes (HBADs), halonitrophenols
small (Ding et al., 2018). Less than 40% of total organic carbon is
(HNPs), halobenzenesulfonic acids, halopyridines and
responsible for the formation of well-known DBPs (Yang and
5
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

halohydroxybenzoic acids (HBACs). Whereas, heterocyclic DBPs THMs and HAAs up to contact time of 2 h. Chlorine firstly reacts
include halofuranones and halopyrroles etc. (Table 4) (Liu et al., with the active group quickly, leading to the rapid production of
2019, 2020a; Yang et al., 2019; Zhang et al., 2018; Wang et al., DBPs in the beginning. As the chlorination continued, availability of
2016; Zhai et al., 2014). Heterocyclic highly DBPs are unstable and both reactive groups and free chlorine declined which reduced the
get decompose under alkaline conditions and in the presence of formation of DBPs (Hong et al., 2012). Another study reported the
excess chlorine. Whereas, the stability of phenyl DBPs vary from increasing formation potential of HAAs in first 3-h of chlorination.
one another depending upon the presence of functional groups. Increased contact time decreases the level of residual chlorine
Nitrogen containing functional groups are relatively more stable, which also reduces HAAs formation (Mompremier et al., 2019).
while brominated phenyl DBPs get decompose in the presence of However, longer reaction time results in the formation of some
chlorine (Hu et al., 2018; 2019; Nihemaiti et al., 2017). The data on temporary DBPs (HANs, HKs and HAs) which shows increasing
the stability of phenyl DBPs is scarce. followed by decreasing pattern with prolonged reaction time
Aromatic DBPs are found in very low concentration i.e. ng L1 to n et al., 2019). They undergo hydrolysis in the existence of free
(Stefa
mg L1 (Yang et al., 2019). Due to less availability of standards and chlorine (Sun et al., 2013). Additionally, phenyl N-DBPs are rela-
high analysis requirements, aromatic DBPs are rarely analyzed tively stable, while rest of the aromatic DBPs get decompose. Yet,
quantitatively (Jiang et al., 2019). Among phenyl DBPs, HPANs, the detailed data on stability of aromatic DBPs is limited (Liu et al.,
HBACs and HPs are found in higher amount due to the presence of 2020a).
carboxylic group in NOM. Among heterocyclic DBPs, only halofur-
anones have been studied and found in relatively higher concen- 4.3. Temperature
tration than phenyl DBPs (Liu et al., 2020a). It has been reported
that aromatic DBPs have dozen to hundred times more toxicity than Temperature affects reaction kinetics. For chlorination, increase
aliphatic DBPs (Liu et al., 2019). It has also been suggested that in temperature leads to the formation of more THMs because
phenyl N- DBPs are more toxic than phenyl carbonaceous (C-DBPs). warmer environment increases the reaction rate. Additionally,
This could be associated with the hydrophilic nature of C-DBPs more NOM is available in summer as compared to the colder
(Zhang et al., 2018; 2020). months (Chowdhury et al., 2011). However, very high temperature
can lead to the volatilization of THMs except chloroform, which is
4. Factors affecting the formation of DBPs stable than other THMs, while temperature dependency of HAAs
formation is variable (Chowdhury et al., 2011). The yield of HAAs
The formation of various DBPs is dependent on number of firstly increases as temperature elevated from 10 to 20  C, but
influential factors: (i) chorine dose (ii) temperature (iii) pH (iv) declined as the temperature continuously raised up to 30  C. The
contact time (v) organic matter content (vi) constituents of water, reduced production of HAAs at 30  C chiefly resulted from the
and (vii) type of the pipe material (Fig. 2) (Voukkali et al., 2015; decomposition of unstable HAAs like trihaloacetic acid (THAA)
Abbas et al., 2015). According to WHO, the residual chlorine level (Hong et al., 2013). Whereas, increase in temperature from 17 to
must be equal or higher than 0.5 mg L1 after contact time of 37  C also lead to the decline in HAAs formation (Mompremier
30 min at pH below 8 and temperature 20ᵒC (Mahmood, 2015). et al., 2019). This decline at temperature below 20  C could be
associated with the combined effect of various pipe materials used
4.1. Chorine dose in the study. Gan et al. (2013) reported the seasonal dependency of
THMs and HAAs formation which is favored by winter.
High disinfectant dose leads to the formation of more hypo- Generally, an elevation of temperature accelerates the reaction
chlorous acid which react with NOM, as a result, more DBPs are rate. If the products are comparatively stable, their production rises
formed (Stef an et al., 2019). An experiment was conducted by Gao with the temperature. However, the temperature upsurge may also
et al. (2013) in which chlorine decay was tested and concentrations improve the decomposition rate of THAAs (trihalogenated acetic
of DBPs including HAAs and THMs were measured. Results revealed acid) as per they are thermally unstable DBPs (Chen et al., 2002;
that higher the chlorine dose is responsible for higher oxidation of Levesque et al., 2006). Further, high temperature also enhances the
organic matter in water samples. Furthermore, the concentrations decomposition of HANs, HK and HAs (Stef an et al., 2019). It has been
of DBPs increased with chlorine dose and reaction time (Gao et al., reported that HKs production increases up to 10e20  C and then
2013). The formation of THMs and HAAs generally increases with start to decrease with further increase in temperature (Hong et al.,
disinfectant dose and residual chlorine (Zhang et al., 2017; Hong 2013). Therefore, the concentrations of the unstable DBPs as the
et al., 2012; Mompremier et al., 2019; Gao et al., 2013). Intermediate function of temperature depend on the relative quantities of their
DBPs also increase with increasing dose, but up to a certain limit formation and decomposition (Hong et al., 2012).
depending upon the stability of intermediated DBPs at given dose,
pH and contact time, afterwards they react with free chlorine to 4.4. pH
produce THMs, HAAs and other DBPs (Hong et al., 2012). For
instance, the degradation rate of DCAN is observed to be higher High pH water produces more THMs, whereas more HAAs are
than its formation at chlorine dose of 10.2 mg L1 (Fang et al., 2010). formed at low pH values (Tian et al., 2017). Moreover, hydrolysis of
Similarly, heterocyclic DBPs also get decompose into THMs and HANs, HKs and HAs is also favored by higher pH values (Stef an et al.,
HAAs in the presence of higher amount of residual chlorine (Hu 2019). The effect of pH on DBPs formation is dissimilar for indi-
et al., 2109). vidual compounds. In the study of chlorination, THMs level elevate
significantly as pH increases. This result is predictable as the alka-
4.2. Reaction time line conditions can assist the hydrolysis reactions of various inter-
mediate DBPs to produce THMs (Xie, 2004). Yet for HAAs, the yields
Reaction time impacts the formation of various categories of were approximately stable as pH rose from 6 to 8 and dropped at
DBPs in a multiple way. High contact time produces more DBPs higher pH (Fang et al., 2010; Hong et al., 2012). Further, pH as low as
because of the reaction with residual chlorine. Reaction time affects 4 also increased the formation of HAAs (Fang et al., 2010). Acidic
the formation of different categories of DBPs in a different way. conditions (at pH 6) favored the formation of N-DBPs (HANs), while
Hong et al. (2012) reported a significant increase in the formation of alkaline conditions up to pH 9 favors the formation of stable THMs
6
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Table 4
Toxicity summary of aromatic DBPs.
a a
Group Name Compound names LC50 Compound names LC50 General Toxicity References
mol L1 mol L1

Phenyl DBPs
HPs 2,4-dichlorophenol 2-chloro-6-iodo-4- Cytotoxic, developmental Zhang et al. (2020); Liu et al. (2019);
methyl-phenol anomalies, endocrine Yang et al. (2019); Liberatore et al.
2-bromophenol 2,6-dibromo-4- 1.33  10⁻4 disruptive and growth (2017); Wagner and Plewa (2017)
chlorophenol inhibition
4-bromophenol 2,6-dichloro-4-iodophenol
2,4,6-trichlorophenol 2.44  10⁻4 2,4,6-tribromophenol 9.99  10⁻5
2-iodophenol 6.01  10⁻4 2,4,6-tribromoresorcinol
4-iodophenol 2.16  10⁻4 2,4-diiodophenol
4-iodo-2-methylphenol 1.63  104 2,6-diiodophenol
2-iodo-24-methylphenol 4,6-diiodo-2,3-
dimethyphenol
2,6-dichloro-4 1.75  10⁻4 2-chloro-4,6-diiodophenol
bromophenol
2-iodo-4,5-dimethylphenol 2,4,6-triiodophenol 7.01  10⁻5
4-iodo-2,5-dimethylphenol 2,4,6-triiodo-1,3,5-
benzenetriol
4-hydroxy-3-iodophenyl 4.08  104 4-hydroxy-3,5- 3.32  105
diiodophenyl
4-iodo-2,6-dimethylphenol 2,4-dibromophenol
HBQs 2,5-dibromohydroquinone 2,3,6-trichloro-1,4- Developmental Yang et al. (2019); Wang et al.
benzoquinone malformations, cytotoxic zka
(2018); Sun et al. (2019); Procha
2,6-dichloro-p- 1.2  105 2,6-Dibromo-p- 1.7  105 et al. (2015)
benzoquinone benzoquinone
2,6-dichloro-3-methyl-1,4- 2,6-dibromo-1,4-
benzoquinone benzoquinone
HBACs 5-chlorosalicylic acid 3,5-dibromo-4- 1.96  103 Growth inhibition, Zhang et al. (2020); Wagner and
hydroxybenzoic acid cytotoxic Plewa (2017); Zhai and Zhang
3,5-dichlorosalicylic acid 4.77  104 3,5-dibromosalicylic acid 2.21  104 (2011)
3,5-dichloro-4- 2.16  103 4-hydroxy-5-iodo-1,3-
hydroxybenzoic acid benzenedicarboxylic acid
5-bromosalicylic acid 2,6-dichloro-4-iodo-3,5-
dihydroxy-benzoic acid
5-bromo-3-methyl- 4-hydroxy-3,5- 2.88  104
salicylic acid diiodobenzoic acid
3-bromo-5-chloro-4- 3,5-diiodosalicylic acid
hydroxybenzoic acid
3-bromo-5-chlorosalicylic 3.35  104 3,5-diiodosalicylic acid
acid
3-iodo-4-hydroxybenzoic 1.5  105 5,6-diiodo-3-ethyl-
acid salicylic acid
3-iodo-4-hydroxy-5-
methylbenzoic acid
HBADs 3,5-dichloro-4-hydroxy- 3.20  104 3,5-dibromo-4- 1.30  104 Cytotoxic, developmental Zhang et al. (2020); Yang et al.
benzaldehyde hydroxybenzaldehyde effects (2019); Zhai and Zhang (2011)
3-bromo-5-chloro-4- 2.04  104 4-hydroxy-3,5-
hydroxy-benzaldehyde diiodobenzaldehyde
4-hydroxy-3- 2-hydroxy-3,5-
iodobenzaldehyde diiodobenzaldehyde
2,6-dibromo-4-
hydroxybenzaldehyde
HNPs 2,6-dichloro-4-nitrophenol 6.73  105 2,4-dibromo-6- Cytotoxic, developmental Zhang et al. (2020); Selvam et al.
nitrobenzenediol defects (2018); Wagner and Plewa (2017);
2-bromo-4-nitrophenol 2,4-diiodo-6-nitrophenol Zhai et al. (2014)
5
2-bromo-6-chloro-4- 8.01  10 2,6-diiodo-4-nitrophenol 3.35  105
nitrophenol
2,6-dibromo-4-nitro- 5.07  105
phenol
HPANs Phenylacetonitrile 2,5- Cytotoxic, developmental Liu et al. (2020a); Zhang et al.
dichlorophenylacetonitrile effects (2018)
4
2-chlorophenylacetonitrile 1.48  10 2,6-
dichlorophenylacetonitrile
3-chlorophenylacetonitrile 2,5-
dichlorophenylacetonitrile
4-chlorophenylacetonitrile 3,4- 8.3  105
dichlorophenylacetonitrile
2,4-
dichlorophenylacetonitrile
Halodipeptidesb NeCl-Phe-Gly NeBreNeCl-Tyr-Ala Tian et al. (2020); Huang et al.
NeCl-Tyr-Gly 3-I-Tyr-Gly (2019)
NeCl-Tyr-Ala 3-I-Tyr-Ala 5.0  103
N,N-di-Cl-Phe-Gly N,N-di-Br-Tyr-Gly
(continued on next page)

7
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Table 4 (continued )
a a
Group Name Compound names LC50 Compound names LC50 General Toxicity References
mol L1 mol L1

N,N-di-Cl-Tyr-Gly N,N-di-Br-Tyr-Ala
NeBr-Tyr-Gly NeBr-3-I-Tyr-Gly
N,N-di-Cl-Tyr-Ala NeBr-3-I-Tyr-Ala
NeBr-Tyr-Ala 3,5-di-Cl-Tyr 4.2  103
NeBreNeCl-Tyr-Gly 3,5-di-Br-Tyr 1.9  103
3-Cl-Tyr 1.15  102 3,5-di-I-Tyr 5.5  104
3-Br-Tyr 1.16  102 3,5-di-I-Tyr-Gly
3-I-Tyr 1.06  103 3,5-di-I-Tyr-Ala 7.8  104
Halobenzenesulfonic 2-iodo-3,4,5-trihydroxy- 2,6-dichloro-3,5- Liu et al. (2020a)
acids benzenesulfonic acid dihydroxy-4-
dodecylbenzenesulfonic
acid
2-chloro-6-iodo-3,4,5- 2,6-diiodo-3,4,5-
trihydroxybenzenesulfonic trihydroxy-
acid benzenesulfonic acid
3,5-dichloro-4- 2,6-dibromo-3,5-
dodecylbenzenesulfonic dihydroxy-4-
acid dodecylbenzenesulfonic
acid
2-bromo-6-iodo-3,4,5-
trihydroxy-
benzenesulfonic
Acid
Other phenyls DBPs 4-bromo-2-methoxy-5- 4-hydroxy-3-iodo-5- Liu et al. (2020a)
methylphenol nitro-benzaldehyde
2,6-dibromo-4- 2-hydroxy-3,5-diiodo-4-
methoxyphenol formyl-benzoic acid
3-formyl-4-hydroxy-5- 2,4,6-triiodo-5-methoxy-
iodo-benzoic acid 1,3-benzenediol
Heterocyclic DBPs
Halopyrroles 3-chloro-2,5-furandione 2,3,5-tribromopyrrole 6.1  105 Developmental toxicity Liu et al. (2020a)
2,3,4,5-tetrabromopyrrol 2,3,4-tribromo-5-
chloropyrrol
2-hydroxy-3,5,6- 2,3,4-tribromo-5-
trichloropyridine iodopyrrol
Halofuranones chloro-4-methyl-5- trichloro-4-methyl-5- 2.75  104 Carcinogenic, genotoxic, s and
Diana et al. (2019); Corte
hydroxy-2(5H)-furanones hydroxy-2(5H)-furanones mutagenic developmental Marcos (2018); Wang et al. (2013)
chloro-5-methyl-5- trichloro-5-methyl-5- malformations
hydroxy-2-furanones hydroxy-2-furanones
dichloro-4-methyl-5- tertchloro-5-methyl-5-
hydroxy-2(5H)-furanones hydroxy-2-furanones
trichloro-5-methyl-5- tertchloro-5-methyl-5-
hydroxy-2-furanones hydroxy-2-furanones
a
LC50 values obtained by Chinese hamster ovary cells bioassay with the exposure time of 72 hours. b48 hours exposure duration.
HP: Halophenols; HBQs: Halobenzoquinone; HBACs: Halohydroxybenzoic acids; HBADs: Halohydroxybenzaldehydes; HNPs: Halonitrophenols; HPANs: (Halo)
phenylacetonitriles

(chloroform) (Ding et al., 2018). Further, when pH reaches 7.6, hy- understand the proper mechanism and impact of various ions and
pochlorite become dominant species which has less disinfectant metals on DBPs formation (Zhao et al., 2016). Molecular weight,
potential, therefore, less DBPs are formed (Sun et al., 2013; Tian structure and hydrophobicity of NOM also affect the formation of
et al., 2017). DBPs. The hydrophobic organic matter usually consists of high
Alkaline pH (>8) leads to the hydrolysis of N-DBPs and HKs molecular weight organic materials like humic substances,
(Stef
an et al., 2019). The yields of HKs at pH 6 and 7 are reported to whereas, hydrophilic organic matter is composed of low molecular
be higher than pH 8. As HKs and HAs are unstable, they can easily weight organic contents i.e. amino acids, carboxylic acids, carbo-
decompose under the higher pH conditions and convert into more hydrates etc. (Korotta-Gamage and Sathasivan, 2017). Literature
stable DBPs i.e., THMs and HAAs (Hong et al., 2013). Moreover, reported that low molecular weight NOM promote the formation of
alkaline conditions also lead to the decomposition of halofuranones brominated DBPs (e.g. DBCM and BDCM) (Kristiana et al., 2017;
(Nihemaiti et al., 2017). Gough et al., 2012). During chlorination, bromide and iodide gets
oxidized into HOBr and HOI. Hydrophilic/low MW organic fractions
are more reactive towards HOBr and HOI and form brominated and
4.5. Organic matter content iodinated DBPs. Besides, hydrophobic/high molecular weight NOM
fractions enhance the formation of THMs and HANs (Li et al., 2017;
DBPs are formed by the reaction of disinfectants with NOM or Lin et al., 2014). It has been observed that hydrophobic NOM has
the other constituents of water. Therefore, NOM plays vital role in two times higher THMs formation potential than hydrophilic frac-
the formation of DBPs. NOM is usually higher in surface water as tions because THMs formation occur due to the reaction of HOCl/
compared to the ground water, therefore, more DBPs are formed in OCl with the humic like substances (Korotta-Gamage and
surface water as compared to the groundwater (Benson et al., 2017; Sathasivan, 2017). However, no effect of hydrophobicity or
Abbas et al., 2015). The complex nature of NOM makes it difficult to
8
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Navalon et al., 2009) depending upon the type of NOM (Ta et al.,
2020; Zhao et al., 2016). Similarly, NHþ 4 consumes disinfectant,
hence, reduce the formation of THMs (Ta et al., 2020; Sun et al.,
2013). It is suggested that NHþ 
4 reacts with ClO and converted into
NH2Cl. As a result, availability of disinfectant decreases which
eventually leads to the decreased THMs formation (Ta et al., 2020).
However, NHþ 4 favors the formation of N-DBPs (Zhang et al., 2019;
Voukkali and Zorpas, 2015).
Presence of multiple metal ions showed analogous but not
synergistic effect on the formation of DBPs, specifically THMs and
HAAs (Liu et al., 2012). Influence of each metal ion on DBPs for-
mation is highly dependent upon the structure and molecular
weight of the precursors (Zhao et al., 2016). Presence of As3þ re-
duces THMs formation as it consumes the disinfectant (Ta et al.,
2020). During the chlorination of tannic acid containing water in
the presence of Mg2þ, Fe2þ, Mn2þ, Cu2þ and Al3þ inhibit the for-
mation of THMs with the order of Mg2þ < Cu2þ < Fe2þ
< Mn2þ < Al3þ and promote the formation of HAAs with the order
of Mg2þ < Mn2þ < Cu2þ < Al3þ < Fe2þ (Liu et al., 2012). The reaction
mechanism of each metal is different. Fe3þ, Cu2þ, and Al3þ accel-
erates THMs formation as they make bond with humic acid and Cl
or Br present in water. An intermediate product (e.g. metald Cl
Fig. 2. Factors affecting the formation of DBPs. dhumic acid) is formed which weakens the original bonds of hu-
mic acid and they get easily attacked by the disinfectant (Ta et al.,
2020). The catalytic effect of iron and copper depends upon the
hydrophilicity has been observed on other classes like HKs and HAs molecular structure and molecular weight of NOM, respectively

(Li et al., 2017; Swietlik and Sikorska, 2004). Hydrophobic and (Zhao et al., 2016). Presence of Mn2þ with co-existing algal blooms
HMW NOM contains more phenolic groups, hence, increase the in water can enhance the formation potential of HAAs (Tada et al.,
formation of aromatic DBPs (Diemert et al., 2013). 2020).

4.6. Constituents of water 4.7. Type of the pipe material

In drinking water treatment facilities, the high reactivity of Chlorine reacts differently with pipe material and formation of
bromine with phenol-like organic structures led to the production DBPs varies with the type pipe material used in the water distri-
of brominated DBPs during chlorination (Dong et al., 2017). A study bution networks (Mompremier et al., 2019). Generally, pipe mate-
exhibited that the existence of Br and I boost the production of rials used in water supply systems include three generic types:
halogenated DBPs, however the impact of Br is dominants and cementitious, plastic, and metallic. Among them, high density
forms more toxic DBPs (Dong et al., 2017). In bromine containing polyethylene (HDPE), polypropylene (PP), polyvinyl chloride (PVC),
water, Br gets oxidized by HOCl to HOBr which is a strong ductile iron pipe (DI), stainless steel pipe (SS), galvanized steel
oxidizing agent and incorporate bromine in DBPs (Chowdhury, pipes (GS) and polyethylene (PE) pipe are widely applied
2013). High concentration of brominated furanones and bromi- (Mompremier et al., 2019; Zhang et al., 2017). In recent years,
nated benzoquinones have been reported in bromine rich water plastic pipes have become the dominant material for various
(Pan and Zhang, 2013). Similarly, presence of iodine produces HOI/ diverse types of supply networks owing to their high flexibility high
OI which also reacts with NOM producing I-DBPs (Pantelaki and corrosion resistance, ease of installment, jointing properties, low
Voutsa, 2018). Nevertheless, both processes are pH dependent. cost, low weight and durability. These pipe materials have great
Literature reported that bromine has more reactivity for THMs and effect on production of DBPs (Poleneni and Innis, 2019).
HAAs formation as compared to chlorine. Bond et al. (2012) re- Literature suggested that the formation of THMs under the three
ported that only 5e10% hypochlorous acid became incorporated different pipe materials followed the order as: PE > DI > SS (Zhang
into THMs, however, 50% bromine gets incorporated in THMs for- et al., 2017). Relying on the analysis data of chlorine decay, most of
mation. Generally, chlorinated HAAs are dominant but the presence the chlorine decay in SS pipe is wall decay, which denotes that a
of Br favors the formation of brominated HAAs, hence, making small amount of chlorine is left in bulk water to react with organic
them dominant specie (Hong et al., 2012). Other anions, SO 
4 , NO3 matter to produce DBPs, therefore causing declined formation of

and Cl has no direct effect on DBPs formation (Zhang et al., 2019). DBPs. However, based upon bacterial analysis, it is suggested that
Whereas, F has indirect effect on DBPs formation with respect to DI and SS pipes are more suitable to be used in water distribution
pH. F shows no interaction with NOM or disinfectant, however, it systems as compared to PE. Because DI and SS pipes support less
affects pH of the solution because of same OH and F radius. As, complex bacterial and less abundance of potential opportunistic
the active specie of disinfectant changes with pH, so it influence the pathogens than PE pipes (Zhang et al., 2017). Another study sug-
THMs speciation (Ta et al., 2020). gested that HAAs concentration in the four pipe loop systems fol-
Among cations, Kþ has no notable influence on DBPs formation lowed the order: HDPE > PP > PVC > GS, due to the variation of
during chlorination (Zhang et al., 2019). Hardness reduces the for- chlorine concentration within the pipe loop system. SG pipe loop
mation of THMs. As, Ca2þ and Mg2þ bind with the functional groups showed higher chlorine decay rate as compared to the other pipe
of humic acid in the form of chelates or poly dentate complexes loop systems (Mompremier et al., 2019).
reducing its availability to react with disinfectant to form DBPs (Ta Literature reported that the concentrations of THMs and HAAs
et al., 2020). However, some studies reported that hardness has in the different segments of the distribution systems varied
catalytic effect for THMs and HAAs formation (Zhao et al., 2016; depending on other variables as well, such as age of the pipe
9
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

material, diameter of the pipe and presence of NOM (Zhong et al., in addition to the reduced tail length, whereas HAA (tribromoacetic
2012; Mompremier et al., 2019; Zimoch et al., 2016). Zhong et al. acid and dichloroacetic acid) significantly enhanced the malfor-
(2012) reported that under the conditions of the same initial mation rate (Teixido  et al., 2015). Another study reported the
chlorine concentrations the ranking of the pipe material in different biochemical changes and liver damage in Cyprinus carpio induced
pipe materials was PVC > SS > DI. Similarly, it has also been re- by DBPs (chloroform and iodoform) (Perveen et al., 2019).
ported that the formation of DBPs increases with increasing It has been reported that aromatic DBPs degraded into aliphatic
diameter of the pipe (Zhong et al., 2012). The reaction time is the DBPs (THMs and HAAs) either through ring opening or cleavage of
most important factor determining the formation process of THMs the side chain. The formation of intermediate compound during
and depends on the distance between disinfection points and this conversion are known to be more toxic than the parent/
sampling points in a distribution system. It has been observed that daughter compounds. For example, HBQs formed by the degrada-
THMs concentration drops with the increasing distance (Zimoch tion of halophenols are 1000 times more cytotoxic than regulated
et al., 2016). aliphatic DBPs (Li et al., 2015). Similarly, halofuranones have also
been reported as mutagenic and genotoxic. Aromatic DBPs are
10e100 time more toxic than aliphatic DBPs in term of develop-
5. Toxicity of DBPs
mental toxicity and growth inhibition (Wang et al., 2018; Wagner
and Plewa, 2017). For instance, HBQs exposed to zebrafish em-
Chlorination cause mutation in bacterial DNA, however, its ef-
bryo caused oxidative stress, curved spines, heart malformation
fect on mammalian cell has also been documented previously (de
and other developmental defects, and this toxicity potential of
Castro Medeiros et al., 2019). A study reported that THMs
HBQs was 200 times higher than aliphatic HAAs (Wang et al., 2018).
(750 mg L1) are responsible for hepatotoxicity, hyperplasia in
Similarly, 2,4,6-tribromophenol was 125 times more toxic than
urinary bladder and metaplasia in small intestine of male Wistar
HAAs to Tetraselmis marina and 72 times to the embryo of Platy-
rats at dose 16.17 mg kg1 and time duration of 1 month (Abd El-
nereis dumerilii (Liu and Zhang, 2014). Additionally, HPANs showed
Halim et al., 2017). Another study suggested that THMs could be
higher cytotoxicity in CHO cells as compared to their aliphatic
responsible for almost 700 cancer cases in Canada (Chowdhury
counterparts (HANs) (Zhang et al., 2018). Further, selected groups of
et al., 2011). THMs and HANs formed as a result of chlorination
phenyl DBPs were evaluated in term of oxidative stress in CHO cells
are two times more genotoxic and cytotoxic than same DBPs
in vitro. Toxicity order was HNPs > HPs > HBADs > HBACs (Zhang
formed by chloramination because chlorination yields higher con-
et al., 2020). Toxicity of aromatic DBPs and their LC50 values have
centration of DBPs than chloramination at same disinfectant doses
been described in Table 4.
(Ding et al., 2018). Chloroform and bromoform can cause cytotox-
icity in human cell (Lodhi et al., 2017). Studies showed that HAAs
are genotoxic as they inhibit glyceraldehyde-3-phosphate dehy- 6. International regulations
drogenase (GADPH) which leads to the disruption of cellular
metabolism and energy homeostasis (Dad et al., 2018). Dad et al. Based upon toxicity of DBPs, US EPA has regulated two groups of
(2018) evaluated the effects of three HAAs on Chinese hamster DBPs: THMs and HAAs and EU has regulated THMs only (Table 5).
ovary (CHO) cells. Results revealed that mono-HAAs were the Canada has set different maximum permissible limit (80 mgL1 for
strongest inhibitors of GAPDH and greatly reduced cellular ATP HAAs) and each province of Canada has different guideline values
levels. However, di- and tri-HAAs showed weak inhibition of (Chowdhury et al., 2011). Nigeria has 1 mgL1 maximum allowable
GAPDH and enhanced cellular ATP levels. Further, HAAs also acti- limit for THMs in drinking water (Benson et al., 2017). Similarly,
vated pyruvate dehydrogenase complex (PDC) by inhibiting pyru- South Africa, Australia and Singapore has their own drinking water
vate dehydrogenase kinase (Dad et al., 2018). Further, brominated quality standards (Yang et al., 2018).
HAAs are 201 times more toxic than chlorinated HAAs (Pan and
Zhang, 2013). HANs are known for causing genotoxicity and cyto- 7. Abatement approaches
toxicity in Chinese hamster ovary cell (Muellner et al., 2007).
Further, HANs has been reported to cause DNA damage and ROS The effective removal of DBPs by several treatment processes
generation in human derived hepatoma cell line (Lu et al., 2018). can be achieved such as adsorption, air stripping, ozonation fol-
Although HANs are unregulated, yet they are more toxic than lowed by chlorination, enhanced coagulation, membrane tech-
regulated DBPs. Similarly, HAs are known to cause aneuploidy, DNA niques, and advanced oxidation processes (Chaukura et al., 2020).
damage and chromosomal aberrations in mammalian cells. Liter- Granulated activated carbon (GAC) is also useful for the removal of
ature reported that HAs can cause nephrotoxicity in humans and THMs and HAAs and other aromatic DBPs (Marais et al., 2018; Jiang
cytotoxicity in rats (Jeong et al., 2015). Further, HAcAms are known et al., 2017). High temperature and longer contact time can enhance
to cause cytotoxicity and DNA damage in Chinese hamster ovary the removal of DBPs. However, biological activity of carbon can
(CHO) cells (Plewa et al., 2008). Toxicity data and LC50 values of effectively remove HAAs and other biodegradable DBPs from water
aliphatic DBPs have been summarized in Table 3. distribution systems (Korotta-Gamage and Sathasivan, 2017).
Chlorination can exhibit high acute toxicity for aquatic biota as Physical methods: boiling, addition of sodium carbonate and
well. Literature reported that chlorination of wastewater with ascorbic acid can also be used for the abatement of DBPs (Liu et al.,
2.5_5 mg Cl2 L1 caused toxicity to fishes with enhanced mortality 2021). Summary of all these possible techniques has been provided
of Daphnia magna neonates (Da Costa et al., 2014). This could be in Table 6.
associated with the presence of DBPs. Niu et al. (2017) reported that Among physical methods, boiling is one of the easiest ways to
ecological risk levels of THMs in freshwater are higher than sea reduce DBPs levels in drinking water. Shi et al. (2017) reported the
water. However, risk assessment of THMs in surface water was removal of halo-DBPs through boiling. Removal of HANs after
notably low. Teixido  et al. (2015) evaluated the effects of THMs and boiling involves both volatilization and decomposition steps (Shi
HAA5 on the development of zebra fish embryo at concentration et al., 2017). In addition to boiling, incorporation of lemon in
range of 20e100 mgmL1. Results revealed that THMs caused chlorinated water has been reported as a successful method for the
adverse developmental effects and weak induction of DNA damage abatement of halo-DBPs (Liu et al., 2021). They have also reported

10
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Table 5
Guideline values of various DBPs.

DBPs Cancer group (EPA)b WHO (2008) US EPA (1998) European Union (Directive, 1998)

mg L1
THM4 80 100
Chloroform B2 300 70
Bromodichloromethane B2 60 0
Chlorodibromomethane C 100 60
Bromoform B2 100 0
HAA5 60
Dichloroacetic acid B2c 50 0
Trihloroacetic acid C 200 20
Monochloroacetic acid D 20 70
a
Monobromoacetic acid D
a
Dibromoacetic acid D
HANs
Dibromoacetonitrile C 70
Dichloroacetonitrile C 20
a
Bromochloroacetonitrile
a
Trichloroacetonitrile
HKs
a
Chloroacetones (1,1-Dichloroacetone)
HAs
Chloral hydrate (Trichloroacetaldehyde) C 10
Phenyl DBPs
a
2- Chlorophenol D
a
2,4-Dichlorophenol D
2,4,6-Trichlorophenol B2 200
a
Inadequate data in available for setting guideline values.
b
B2¼ Probable human carcinogen based upon enough data from animal studies; C¼ Possible human carcinogen; D ¼ Not classified as human carcinogenic.
c
Probable human carcinogen according to IARC (2004).

that heating of lemon containing water up to 100  C can reduce the Xu et al., 2019; Marais et al., 2018). Fu et al. (2017) conducted a pilot
cytotoxicity of tap water up to 67%. Ascorbic acid (3.7 mg L1) can scale experiment to evaluate a biofiltration process for the removal
quench residual chlorine which restricts the formation of DBPs (Liu of DBPs precursors in order to maintain the formation potential (FP)
et al., 2021). Similar results has been reported by the addition of of 36 different DBPs. The biofiltration process included sand/
sodium carbonate (Liu et al., 2020b). anthracite (SA) filtration coupled with a biologically-active GAC
Membrane technologies involves reverse osmosis, nano- followed by pretreatments. Results revealed that biofiltration pro-
filtration, microfiltration, ultrafiltration and ceramic membranes. cess reduced the FP of total DBPs (24%), more specifically; HAs
These membranes are used for purification of treated water and (33.62%) > HAAs (28.13%) > THMs (20.46%) > HKs (13.46%) > HANs
eliminates NOM and other DBPs precursors which ultimately (8.82%). However, the precursors of HANs were recalcitrant towards
reduce the formation potential of DBPs. Each membrane has biofiltration process, hence showed negative removal. This method
different NOM elimination potential. However, accumulation of proved to be more efficient for the halogenated DBPs as compared
microbes, NOM and other precursors on membranes can fill the to the other (Fu et al., 2017). So, biofiltration processes can be used
voids reducing the efficiency of membranes (Chaukura et al., 2020; as effective method for the removal or control of DBPs formation,

Table 6
Summary of various approaches used for the abatement of DBPs.

Techniques Principle DBPs References

Boiling Decomposition and Volatilization All halo-DBPs Shi et al. (2017);


Liu et al. (2020b); 2021
Pan et al., 2014
Ascorbic Acid Quenching of residual chlorine All halo-DBPs Liu et al., 2021
Sodium Carbonate Increase pH which consequently increase the All halo-DBPs except THMs Liu et al. (2020b)
dehalogenation of DBPs
Adsorption (GAC) Concentration of DBPs on adsorbent; NOM removal THMs, HAAs, HKs, HAs, Chen et al. (2019);
aromatic DBPs Jiang et al. (2017); Chowdhury
(2013)
Air stripping Transfer of volatile compounds into air/gas THMs Valdivia-Garcia et al. (2019)
Enhanced Co-agulation NOM/DOC removal, THMs, HAAs, phenyl DBPs €a
Sillanpa € et al. (2018)
Membrane technologies Filtration/removal of NOM, microbes, TOC and other THMs, HAAs Guo et al. (2019);
DBP precursors Xu et al. (2019);
Marais et al. (2018);
Osawa et al. (2017)
Advanced oxidation processes NOM removal, mineralization and degradation of DBP THMs, HAAs, aromatic DBPs Sharma et al. (2018)
precursors
Combined approaches (Ozonation followed by Oxidation/breakdown of DBP precursors THMs, HAAs, N-DBPs, HKs, Chen et al. (2019);
chlorination) Phenyl DBPs Sillanpa€a€ et al. (2018)
Huang et al., 2018
Stepwise Chlorination Reduce chlorine dose Overall DBPs Li et al. (2017)

11
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

however, efficiency of the method depends upon the indigenous 9. Future perspectives
microbial communities, used in supporting media, and their
respective efficiencies (Chen et al., 2019; Fu et al., 2017). Literature asserts that following gaps exist
Advanced oxidation processes (AOPs) involves both photo-
chemical (UV technology) and non-photochemical (ozonation and  Little attention has been paid on emerging DBPs (HANs), though
Fenton reaction) processes. Another technique, non-thermal some studies reported that HANs are more toxic than the
plasma is also used for the degradation of NOM and DBPs. This is regulated DBPs.
one of the most feasible AOP due to its low operational cost and  Future studies should focus on generating the data to regulate
high efficiency (Ma et al., 2020; Aziz et al., 2019). A study reported emerging DBPs.
that ozonation process (0.7 mg of O3/mg of DOC) followed by  Specific NOM with respect to water source should be evaluated.
chlorination reduced the FP of THMs, HAAs, HANs and HKs but no  New methods should be developed to quantify aromatic DBPs at
notable change was observed in the FP of HAs (Chen et al., 2019). such low concentrations.
This process reduced the FP of overall DBPs up to 25%, however, it  Economical feasible technologies must be improved.
enhance the bromate formation. Up-flow biological activated car-  Toxicological studies are limited and human risk data is also
bon (UBAC) at 15 min empty bed contact time (EBCT) followed by scarce.
chlorination reduced the formation potential of DBPs up to 35%.  Most of the technologies deal with the post-formation removal
Further, combined O3/UBAC process (with same condition) not only of DBPs, pretreatment removal of precursors should be
reduce the formation potential of total DBPs up to 50% but also explored.
reduce their toxicities (Chen et al., 2019). Combined approaches are
also used to reduce the formation potential of DBPs. For instance,
UV combined with chlorination reduces the formation of HAcAms Authors contribution
(Liu et al., 2020c). Details of all these methods have already been
explained elsewhere (Chaukura et al., 2020). Sundas Kali, collected the research articles, reviewed and pre-
Several points must be considered while selecting of appro- pared first original draft, screened and analyzed the key findings.
priate control strategies i.e. control efficiency, cost effectiveness Marina Khan, collected the research articles, reviewed and pre-
and potential health hazards etc. It is obvious that boiling, coagu- pared first original draft. Muhammad Sheraz Ghaffar, collected the
lation and optimization of disinfection process are economical research articles, reviewed and prepared first original draft. Sajida
techniques. However, adsorption, membrane technologies, AOPs Rasheed, screened and analyzed the key findings. Amir Waseem,
and combined treatments are costly, yet efficient methods for the screened the papers, constructed the database, review and editing.
control of DBPs. Therefore, appropriate technique must be selected Muhammad Mazhar Iqbal, refined and improved the document and
and further methods should be explored which are cost effective Revisions. Muhammad Bilal khan Niazi, refined and improved the
and efficient. document and Revisions. Mazhar Iqbal Zafar, supervised the whole
work, Conceptualization, writing and review, and finalized the
manuscript.
8. Conclusions and recommendations
Declaration of competing interest
This paper summarized the literature on the formation of THMs,
HAAs, HANs, HKs and HAs in drinking water. It has been concluded The authors declare that they have no known competing
that increase in chlorine dose, NOM, contact time and temperature financial interests or personal relationships that could have
enhances the formation of THMs and HAAs. Whereas, high pH fa- appeared to influence the work reported in this paper.
vors the formation of THMs but oppose HAAs. Hydrophobic NOM
favors the formation of THMs, N-DBPs and aromatic DBPs. Unstable Acknowledgement
DBPs hydrolyzed into THMs and HAAs, however, they become
stable at low temperature, moderate pH and low residual chlorine This work is supported by a grant awarded by Higher Education
concentration. Similarly, constituents of water also influence DBPs Commission (HEC), Pakistan and part of the project #: 9139/Fed-
formation in various ways. Presence of Br and I promotes the eral/NRPU/HEC. Further, authors would like to pay their sincere
formation of brominated and iodinated DBPs, respectively, while indebtedness to M. Phil student Zenab Azam for her help in
SO4, NO3, Cl and F have no direct effect. Besides, the presence improving the graphical abstract and figures. and international
of multiple metal ions also showed analogous but not synergistic scientific community for providing valuable and necessary scien-
effect on the formation of DBPs. Similarly, PE pipes enhance the tific information for the completion of review article.
formation of DBPs. Toxicity data revealed that not only humans, but
aquatic biota is also at risk. Though N-DBPs and aromatic DBPs are References
not regulated, they are also genotoxic and cytotoxic. Aromatic DBPs
are found in very low concentration (ng L1), yet they are more Abbas, S., Hashmi, I., Rehman, M.S.U., Qazi, I.A., Awan, M.A., Nasir, H., 2014. Moni-
toring of chlorination disinfection by-products and their associated health risks
toxic than aliphatic DBPs. Further, most of the remediation tech-
in drinking water of Pakistan. J. Water Health 13 (1), 270e284.
niques are post-formation and expensive. Economically feasible Abd El-Halim, M.I., Mohammed, S.Y., Idris, O.F., Sabahelkhier, M.K., 2017. Histo-
technologies should be developed for the abatement of DBPs. pathological effect on different rat tissues induced by the trihalomethane-
Stepwise chlorination approach can be applied as it reduces DBPs chloroform administered in drinking water. Int. J. Adv. Res. 3 (11), 1319e1329.
Al-Otoum, F., Al-Ghouti, M.A., Ahmed, T.A., Abu-Dieyeh, M., Ali, M., 2016. Disin-
formation, increase disinfection efficiency and reduce operational fection by-products of chlorine dioxide (chlorite, chlorate, and tri-
cost. halomethanes): occurrence in drinking water in Qatar. Chemosphere 164,

12
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

649e656. treatment. Water Res. 123, 224e235.


Aziz, K.H.H., Miessner, H., Mahyar, A., Mueller, S., Kalass, D., Moeller, D., Omer, K.M., Guo, J., Khan, S., Cho, S.H., Kim, J., 2019. ZnS nanoparticles as new additive for
2019. Removal of dichloroacetic acid from aqueous solution using non-thermal polyethersulfone membrane in humic acid filtration. J. Ind. Eng. Chem. 79,
plasma generated by dielectric barrier discharge and nano-pulse corona 71e78.
discharge. Separ. Purif. Technol. 216, 51e57. Hao, R., Zhang, Y., Du, T., Yang, L., Adeleye, A.S., Li, Y., 2017. Effect of water chemistry
Baig, S.A., Mahmood, Q., Nawab, B., Shafqat, M.N., Pervez, A., 2011. Improvement of on disinfection by-product formation in the complex surface water system.
drinking water quality by using plant biomass through household bio sand Chemosphere 172, 384e391.
filter e a decentralized approach. Ecol. Eng. 37, 1842e1848. Hassan, A., Thacker, N.P., Bassin, J., 2010. Trihalomethane formation potential in
Benson, N.U., Akintokun, O.A., Adedapo, A.E., 2017. Disinfection byproducts in treated water supplies in urban metro city. Environ. Monit. Assess. 168,
drinking water and evaluation of potential health risks of long-term exposure 489e497.
in Nigeria. Journal of environmental and public health 2017, 7535797. Hisam, A., Rahman, M.U., Kadir, E., Tariq, N.A., Masood, S., 2014. Microbiological
Bond, T., Templeton, M.R., Kamal, N.H.M., Graham, N., Kanda, R., 2015. Nitrogenous contamination in water filtration plants in Islamabad. J Coll Phys Surg Pak 24,
disinfection byproducts in English drinking water supply systems: occurrence, 345e350.
bromine substitution and correlation analysis. Water Res. 85, 85e94. Hong, H., Xiong, Y., Ruan, M., Liao, F., Lin, H., Liang, Y., 2013. Factors affecting THMs,
Chaukura, N., Marais, S.S., Moyo, W., Mbali, N., Thakalekoala, L.C., Ingwani, T., HAAs and HNMs formation of Jin Lan Reservoir water exposed to chlorine and
Mamba, B.B., Jarvis, P., Nkambule, T.T., 2020. Contemporary issues on the monochloramine. Sci. Total Environ. 444, 196e204.
occurrence and removal of disinfection byproducts in drinking water-A review. How, Z.T., Kristiana, I., Busetti, F., Linge, K.L., Joll, C.A., 2017. Organic chloramines in
Journal of Environmental Chemical Engineering 8 (2), 103659. chlorine-based disinfected water systems: a critical review. J. Environ. Sci. 58,
Chaves, R.S., Guerreiro, C.S., Cardoso, V.V., Benoliel, M.J., Santos, M.M., 2019. Hazard 2e18.
and mode of action of disinfection by-products (DBPs) in water for human Hu, S., Gong, T., Ma, J., Tao, Y., Xian, Q., 2018. Simultaneous determination of
consumption: evidences and research priorities. Comp. Biochem. Physiol. C iodinated haloacetic acids and aromatic iodinated disinfection byproducts in
Toxicol. Pharmacol. 223, 53e61. waters with a new SPE-HPLC-MS/MS method. Chemosphere 198, 147e153.
Chen, H., Lin, T., Chen, W., Tao, H., Xu, H., 2019. Removal of disinfection byproduct Hu, S., Gong, T., Wang, J., Xian, Q., 2019. Trihalomethane yields from twelve aro-
precursors and reduction in additive toxicity of chlorinated and chloraminated matic halogenated disinfection byproducts during chlor (am) ination. Chemo-
waters by ozonation and up-flow biological activated carbon process. Chemo- sphere 228, 668e675.
sphere 216, 624e632. Huang, G., Jmaiff Blackstock, L.K., Jiang, P., Liu, Z., Lu, X., Li, X.F., 2019. Formation,
Chowdhury, I.R., Chowdhury, S., 2017. Desalinated and blended water in Saudi identification, and occurrence of new bromo-and mixed halo-tyrosyl dipeptides
Arabia: human exposure and risk analysis from disinfection byproducts. In: in chloraminated water. Environ. Sci. Technol. 53 (7), 3672e3680.
MATEC Web of Conferences, vol. 120. EDP Sciences, 5002. Huang, H., Zhu, H., Gan, W., Chen, X., Yang, X., 2017. Occurrence of nitrogenous and
Chowdhury, S., 2013. Trihalomethanes in drinking water: effect of natural organic carbonaceous disinfection byproducts in drinking water distributed in Shenz-
matter distribution. WaterSA 39 (1), 1e8. hen, China. Chemosphere 188, 257e264.
Chowdhury, S., Rodriguez, M.J., Sadiq, R., 2011. Disinfection byproducts in Canadian Ioannou, P., Charisiadis, P., Andra, S.S., Makris, K.C., 2016. Occurrence and variability
provinces: associated cancer risks and medical expenses. J. Hazard. Mater. 187 of iodinated trihalomethanes concentrations within two drinking-water dis-
(1e3), 574e584. tribution networks. Sci. Total Environ. 543, 505e513.
Chowdhury, S., Rodriguez, M.J., Sadiq, R., Serodes, J., 2012. Modeling DBPs formation Jeong, C., 2014. Drinking Water Disinfection By-Products: Toxicological Impacts and
in drinking water in residential plumbing pipes and hot water tanks. Water Res. Biological Mechanisms Induced by Individual Compounds or as Complex Mix-
45 (1), 337e347. tures. Doctoral dissertation, University of Illinois.
Chu, W., Gao, N., Yin, D., Krasner, S.W., Templeton, M.R., 2012. Trace determination Jeong, C.H., Postigo, C., Richardson, S.D., Simmons, J.E., Kimura, S.Y., Marin ~ as, B.J.,
of 13 haloacetamides in drinking water using liquid chromatography triple et al., 2015. Occurrence and comparative toxicity of haloacetaldehyde disin-
quadrupole mass spectrometry with atmospheric pressure chemical ionization. fection byproducts in drinking water. Environ. Sci. Technol. 49 (23),
J. Chromatogr. A 1235, 178e181. 13749e13759.
Cortes, C., Marcos, R., 2018. Genotoxicity of disinfection byproducts and disinfected Jiang, J., Han, J., Zhang, X., 2020. Nonhalogenated aromatic DBPs in drinking water
waters: a review of recent literature. Mutat. Res. Genet. Toxicol. Environ. chlorination: a gap between NOM and halogenated aromatic DBPs. Environ. Sci.
Mutagen 831, 1e12. Technol. 54 (3), 1646e1656.
Da Costa, J.B., Rodgher, S., Daniel, L.A., Espíndola, E.L.G., 2014. Toxicity on aquatic Jiang, J., Zhang, X., Zhu, X., Li, Y., 2017. Removal of intermediate aromatic haloge-
organisms exposed to secondary effluent disinfected with chlorine, peracetic nated DBPs by activated carbon adsorption: a new approach to controlling
acid, ozone and UV radiation. Ecotoxicology 23 (9), 1803e1813. halogenated DBPs in chlorinated drinking water. Environ. Sci. Technol. 51 (6),
Daud, M.K., Nafees, M., Ali, S., Rizwan, M., Bajwa, R.A., Shakoor, M.B., et al., 2017. 3435e3444.
Drinking water quality status and contamination in Pakistan. BioMed Res. Int. Karim, Z., Mumtaz, M., Kamal, T., 2011. Health risk assessment of trihalomethanes
2017, 7908183. from tap water in Karachi, Pakistan. J. Chem. Soc. Pakistan 33, 215e219.
De Castro Medeiros, L., de Alencar, F.L.S., Navoni, J.A., de Araujo, A.L.C., do Korotta-Gamage, S.M., Sathasivan, A., 2017. A review: potential and challenges of
Amaral, V.S., 2019. Toxicological aspects of trihalomethanes: a systematic re- biologically activated carbon to remove natural organic matter in drinking
view. Environ. Sci. Pollut. Control Ser. 26 (6), 5316e5332. water purification process. Chemosphere 167, 120e138.
Diana, M., Felipe-Sotelo, M., Bond, T., 2019. Disinfection byproducts potentially Kurajica, L., Bosnjak, M.U., Stankov, M.N., Kinsela, A.S., Stigli  
c, J., Waite, D.T.,
responsible for the association between chlorinated drinking water and bladder Capak, K., 2020. Disinfection by-products in Croatian drinking water supplies
cancer: a review. Water Res. 162, 492e504. with special emphasis on the water supply network in the city of Zagreb.
Diemert, S., Wang, W., Andrews, R.C., Li, X.F., 2013. Removal of halo-benzoquinone J. Environ. Manag. 276, 111360.
(emerging disinfection by-product) precursor material from three surface wa- Li, D., Zeng, S., Gu, A.Z., He, M., Shi, H., 2013. Inactivation, reactivation and regrowth
ters using coagulation. Water Res. 47 (5), 1773e1782. of indigenous bacteria in reclaimed water after chlorine disinfection of a
Ding, S., Chu, W., Krasner, S.W., Yu, Y., Fang, C., Xu, B., Gao, N., 2018. The stability of municipal wastewater treatment plant. J. Environ. Sci. 25 (7), 1319e1325.
chlorinated, brominated, and iodinated haloacetamides in drinking water. Li, Y., Jiang, J., Li, W., Zhu, X., Zhang, X., Jiang, F., 2020. Volatile DBPs contributed
Water Res. 142, 490e500. marginally to the developmental toxicity of drinking water DBP mixtures
Ding, W., Jin, W., Cao, S., Zhou, X., Wang, C., Jiang, Q., Huang, H., Tu, R., Han, S.F., against Platynereis dumerilii. Chemosphere 252, 126611.
Wang, Q., 2019. Ozone disinfection of chlorine-resistant bacteria in drinking Li, Y., Zhang, X., Yang, M., Liu, J., Li, W., Graham, N.J., Yang, B., 2017. Three-step
water. Water Res. 160, 339e349. effluent chlorination increases disinfection efficiency and reduces DBP forma-
Ding, X., Zhu, J., Zhang, J., Dong, T., Xia, Y., Jiao, J., Wang, X., Zhou, W., 2020. tion and toxicity. Chemosphere 168, 1302e1308.
Developmental toxicity of disinfection by-product monohaloacetamides in Liberatore, H.K., Plewa, M.J., Wagner, E.D., VanBriesen, J.M., Burnett, D.B.,
embryo-larval stage of zebrafish. Ecotoxicol. Environ. Saf. 189, 110037. Cizmas, L.H., Richardson, S.D., 2017. Identification and comparative mammalian
Dobaradaran, S., Fard, E.S., Tekle-Ro €ttering, A., Keshtkar, M., Karbasdehi, V.N., cell cytotoxicity of new iodo-phenolic disinfection byproducts in chloraminated
Abtahi, M., Gholamnia, R., Saeedi, R., 2020. Age-sex specific and cause-specific oil and gas wastewaters. Environ. Sci. Technol. Lett. 4 (11), 475e480.
health risk and burden of disease induced by exposure to trihalomethanes Liu, J., Zhang, X., 2014. Comparative toxicity of new halophenolic DBPs in chlori-
(THMs) and haloacetic acids (HAAs) from drinking water: an assessment in four nated saline wastewater effluents against a marine alga: halophenolic DBPs are
urban communities of Bushehr Province, Iran, 2017. Environ. Res. 182, 109062. generally more toxic than haloaliphatic ones. Water Res. 65, 64e72.
Dominguez-Tello, A., Dominguez-Alfaro, A., Go  mez-Ariza, J.L., Arias-Borrego, A., Liu, X., Chen, L., Yang, M., Tan, C., Chu, W., 2020a. The Occurrence, Characteristics,
García-Barrera, T., 2020. Effervescence-assisted spiral hollow-fibre liquid-phase Transformation and Control of Aromatic Disinfection By-Products: A Review.
microextraction of trihalomethanes, halonitromethanes, haloacetonitriles, and Water Research, p. 116076.
haloketones in drinking water. J. Hazard Mater. 397, 122790. Liu, J., Li, Y., Jiang, J., Zhang, X., Sharma, V.K., Sayes, C.M., 2020b. Effects of ascorbate
Du, Y., Lv, X.T., Wu, Q.Y., Zhang, D.Y., Zhou, Y.T., Peng, L., Hu, H.Y., 2017. Formation and carbonate on the conversion and developmental toxicity of halogenated
and control of disinfection byproducts and toxicity during reclaimed water disinfection byproducts during boiling of tap water. Chemosphere 126890.
chlorination: a review. J. Environ. Sci. 58, 51e63. Liu, J., Zhang, X., Li, Y., Li, W., Hang, C., Sharma, V.K., 2019. Phototransformation of
Ewaid, S.H., Rabee, A.M., Al-Naseri, S.K., 2018. Carcinogenic risk assessment of tri- halophenolic disinfection byproducts in receiving seawater: kinetics, products,
halomethanes in major drinking water sources of Baghdad City. Water Resour. and toxicity. Water Res. 150, 68e76.
45 (5), 803e812. Liu, X., Chen, Z., Wang, L., Shen, J., 2012. Effects of metal ions on THMs and HAAs
Fu, J., Lee, W.N., Coleman, C., Nowack, K., Carter, J., Huang, C.H., 2017. Removal of formation during tannic acid chlorination. Chem. Eng. J. 211, 179e185.
disinfection byproduct (DBP) precursors in water by two-stage biofiltration Liu, Z., Lin, Y.L., Xu, B., Hu, C.Y., Zhang, T.Y., Cao, T.C., Gao, N.Y., 2020c. Degradation of

13
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

diiodoacetamide in water by UV/chlorination: kinetics, efficiency, influence Selvam, R., Muniraj, S., Duraisamy, T., Muthunarayanan, V., 2018. Identification of
factors and toxicity evaluation. Chemosphere 240, 124761. disinfection by-products (DBPs) halo phenols in drinking water. Applied Water
Lodhi, A., Hashmi, I., Nasir, H., Khan, R., 2017. Effect of trihalomethanes (chloroform Science 8 (5), 1e8.
and bromoform) on human haematological count. J. Water Health 15 (3), Sharma, A., Ahmad, J., Flora, S.J.S., 2018. Application of advanced oxidation pro-
367e373. cesses and toxicity assessment of transformation products. Environ. Res. 167,
Ma, S., Kim, K., Chun, S., Moon, S.Y., Hong, Y., 2020. Plasma-assisted advanced 223e233.
oxidation process by a multi-hole dielectric barrier discharge in water and its Shi, W., Wang, L., Chen, B., 2017. Kinetics, mechanisms, and influencing factors on
application to wastewater treatment. Chemosphere 243, 125377. the treatment of haloacetonitriles (HANs) in water by two household heating
Mahmood, K., 2015. Naegleria fowleri in Pakistan-an emerging catastrophe. J Coll devices. Chemosphere 172, 278e285.
Physicians Surg Pak 25, 159e160. Sillanp€ €, M., Ncibi, M.C., Matilainen, A., Vepsa
aa €l€
ainen, M., 2018. Removal of natural
Mao, Y.Q., Wang, X.M., Guo, X.F., Yang, H.W., Xie, Y.F., 2016. Characterization of organic matter in drinking water treatment by coagulation: a comprehensive
haloacetaldehyde and trihalomethane formation potentials during drinking review. Chemosphere 190, 54e71.
water treatment. Chemosphere 159, 378e384. Srivastav, A.L., Kaur, T., 2020. Factors affecting the formation of disinfection by-
Marais, S.S., Ncube, E.J., Msagati, T.A.M., Mamba, B.B., Nkambule, T.T., 2018. Com- products in drinking water: human health risk. In: Disinfection By-Products
parison of natural organic matter removal by ultrafiltration, granular activated in Drinking Water. Butterworth-Heinemann, pp. 433e450.
carbon filtration and full scale conventional water treatment. Journal of envi- Stalter, D., O’Malley, E., von Gunten, U., Escher, B.I., 2020. Mixture effects of drinking
ronmental chemical engineering 6 (5), 6282e6289. water disinfection by-products: implications for risk assessment. Environ. Sci.:
Mompremier, R., Fuentes Mariles, O.A.,  Becerril Bravo, J.E., Ghebremichael, K., 2019. Water Research & Technology 6 (9), 2341e2351.
Study of the variation of haloacetic acids in a simulated water distribution Stefan, D., Erdelyi, N., Izsa
k, B., Za
ray, G., Vargha, M., 2019. Formation of chlorination
network. Water Supply 19 (1), 88e96. by-products in drinking water treatment plants using breakpoint chlorination.
Moradi, S., Liu, S., Chow, C.W., van Leeuwen, J., Cook, D., Drikas, M., Amal, R., 2017. Microchem. J. 149, 104008.
Chloramine demand estimation using surrogate chemical and microbiological Sulehria, A.Q.K., Mustafa, Y.S., Kanwal, B., Nazish, A., 2013. ASSESSMENT OF
parameters. J. Environ. Sci. 57, 1e7. DRINKING WATER QUALITY IN ISLAMPURA, distt. LAHORE. (Local report). Sci.
Muellner, M.G., Wagner, E.D., McCalla, K., Richardson, S.D., Woo, Y.T., Plewa, M.J., Int. 25 (2), 359e361.
2007. Haloacetonitriles vs. regulated haloacetic acids: are nitrogen-containing Sun, X., Wei, D., Liu, W., Geng, J., Liu, J., Du, Y., 2019. Formation of novel disinfection
DBPs more toxic? Environ. Sci. Technol. 41 (2), 645e651. by-products chlorinated benzoquinone, phenyl benzoquinones and polycyclic
Nabeela, F., Azizullah, A., Bibi, R., Uzma, S., Murad, W., Shakir, S.K., H€ader, D.P., 2014. aromatic hydrocarbons during chlorination treatment on UV filter 2, 4-
Microbial contamination of drinking water in Pakistanda review. Environ. Sci. dihydroxybenzophenone in swimming pool water. J. Hazard Mater. 367,
Pollut. Control Ser. 21 (24), 13929e13942. 725e733.
Navalon, S., Alvaro, M., Garcia, H., 2009. Ca2þ and Mg2þ present in hard waters Sun, Y., Chen, Z., Wu, G., Wu, Q., Zhang, F., Niu, Z., Hu, H.Y., 2016. Characteristics of
enhance trihalomethane formation. J. Hazard Mater. 169 (1e3), 901e906. water quality of municipal wastewater treatment plants in China: implications
Nieuwenhuijsen, M.J., Toledano, M.B., Eaton, N.E., Fawell, J., Elliott, P., 2000. Chlo- for resources utilization and management. J. Clean. Prod. 131, 1e9.
rination disinfection byproducts in water and their association with adverse Ta, N., Li, C., Wang, Y., An, W., 2020. Effects of ions on THM formation during
reproductive outcomes: a review. Occup. Environ. Med. 57 (2), 73e85. chlorination of bromide-containing water. Water, Air, Soil Pollut. 231 (8), 1e17.
Nihemaiti, M., Le Roux, J., Hoppe-Jones, C., Reckhow, D.A., Croue , J.P., 2017. For- Tada, Y., Cordero, J.A., Echigo, S., Itoh, S., 2020. Effect of coexisting manganese ion on
mation of haloacetonitriles, haloacetamides, and nitrogenous heterocyclic the formation of haloacetic acids during chlorination. Chemosphere 263,
byproducts by chloramination of phenolic compounds. Environ. Sci. Technol. 51 127862.
(1), 655e663. Tahir, M.A., Rasheed, H., Imran, S., 2010. Water Quality Status in Rural Areas of
Niu, Z., Li, X., Zhang, Y., 2017. Composition profiles, levels, distributions and Pakistan. Pakistan Council of Research in Water Resources.
ecological risk assessments of trihalomethanes in surface water from a typical Tian, D., Moe, B., Huang, G., Jiang, P., Ling, Z.-C., Li, X.-F., 2020. Cytotoxicity of
estuary of Bohai Bay, China. Mar. Pollut. Bull. 117 (1e2), 124e130. halogenated tyrosyl compounds, an emerging class of disinfection byproducts.
Osawa, H., Lohwacharin, J., Takizawa, S., 2017. Controlling disinfection by-products Chem. Res. Toxicol. 33 (4), 1028e1035.
and organic fouling by integrated ferrihydriteemicrofiltration process for sur- Valdivia-Garcia, M., Weir, P., Graham, D.W., Werner, D., 2019. Predicted impact of
face water treatment. Separ. Purif. Technol. 176, 184e192. climate change on trihalomethanes formation in drinking water treatment. Sci.
Padhi, R.K., Subramanian, S., Satpathy, K.K., 2019. Formation, distribution, and Rep. 9 (1), 1e10.
speciation of DBPs (THMs, HAAs, ClO2, andClO3) during treatment of Van Haute, S., Tryland, I., Escudero, C., Vanneste, M., Sampers, I., 2017. Chlorine
different source water with chlorine and chlorine dioxide. Chemosphere 218, dioxide as water disinfectant during fresh-cut iceberg lettuce washing disin-
540e550. fectant demand, disinfection efficiency, and chlorite formation. Lebensm. Wiss.
Pan, Y., Zhang, X., 2013. Four groups of new aromatic halogenated disinfection Technol. 75, 301e304.
byproducts: effect of bromide concentration on their formation and speciation Villanueva, C.M., Cordier, S., Font-Ribera, L., Salas, L.A., Levallois, P., 2015. Overview
in chlorinated drinking water. Environ. Sci. Technol. 47 (3), 1265e1273. of disinfection by-products and associated health effects. Current environ-
Pardakhti, A.R., Bidhendi, G.R.N., Torabian, A., Karbassi, A., Yunesian, M., 2011. mental health reports 2 (1), 107e115.
Comparative cancer risk assessment of THMs in drinking water from well water Villanueva, C.M., Gracia-Lavedan, E., Bosetti, C., Righi, E., Molina, A.J., Martín, V.,
sources and surface water sources. Environ. Monit. Assess. 179, 499e507. et al., 2016. Colorectal cancer and long-term exposure to trihalomethanes in
Perveen, S., Hashmi, I., Khan, R., 2019. Evaluation of genotoxicity and hematological drinking water: a multicenter caseecontrol study in Spain and Italy. Environ.
effects in common carp (Cyprinus carpio) induced by disinfection by-products. Health Perspect. 125 (1), 56e65.
J. Water Health 17 (5), 762e776. Wagner, E.D., Plewa, M.J., 2017. CHO cell cytotoxicity and genotoxicity analyses of
Plewa, M.J., Wagner, E.D., Richardson, S.D., 2017. TIC-Tox: a preliminary discussion disinfection by-products: an updated review. J. Environ. Sci. 58, 64e76.
on identifying the forcing agents of DBP-mediated toxicity of disinfected water. Wang, C., Yang, X., Zheng, Q., Moe, B., Li, X.F., 2018. Halobenzoquinone-induced
J. Environ. Sci. 58, 208e216. developmental toxicity, oxidative stress, and apoptosis in zebrafish embryos.
Plewa, M.J., Simmons, J.E., Richardson, S.D., Wagner, E.D., 2010. Mammalian cell Environ. Sci. Technol. 52 (18), 10590e10598.
cytotoxicity and genotoxicity of the haloacetic acids, a major class of drinking Wang, S., Tian, D., Zheng, W., Jiang, S., Wang, X., Andersen, M.E., Qu, W., 2013.
water disinfection by-products. Environ. Mol. Mutagen. 51 (8-9), 871e878. Combined exposure to 3-chloro-4-dichloromethyl-5-hydroxy-2 (5H)-furanone
Plewa, M.J., Muellner, M.G., Richardson, S.D., Fasano, F., Buettner, K.M., Woo, Y.T., and microsytin-LR increases genotoxicity in Chinese hamster ovary cells
McKague, A.B., Wagner, E.D., 2008. Occurrence, synthesis, and mammalian cell through oxidative stress. Environ. Sci. Technol. 47 (3), 1678e1687.
cytotoxicity and genotoxicity of haloacetamides: an emerging class of nitrog- Wang, W., Moe, B., Li, J., Qian, Y., Zheng, Q., Li, X.F., 2016. Analytical characterization,
enous drinking water disinfection byproducts. Environ. Sci. Technol. 42 (3), occurrence, transformation, and removal of the emerging disinfection
955e961. byproducts halobenzoquinones in water. Trac. Trends Anal. Chem. 85, 97e110.
Plewa, M.J., Wagner, E.D., Jazwierska, P., Richardson, S.D., Chen, P.H., McKague, A.B., Xu, D., Bai, L., Tang, X., Niu, D., Luo, X., Zhu, X., Liang, H., 2019. A comparison study of
2004. Halonitromethane drinking water disinfection byproducts: chemical sand filtration and ultrafiltration in drinking water treatment: removal of
characterization and mammalian cell cytotoxicity and genotoxicity. Environ. Sci. organic foulants and disinfection by-product formation. Sci. Total Environ. 691,
Technol. 38 (1), 62e68. 322e331.
Poleneni, S.R., Inniss, E.C., 2019. Array of prediction tools for understanding extent Yang, M., Zhang, X., 2016. Current trends in the analysis and identification of
of wall effects on DBP formation in drinking water distribution systems. emerging disinfection byproducts. Trends in Environmental Analytical Chem-
J. Water Supply Res. Technol. - Aqua 68 (6), 390e398. istry 10, 24e34.
Proch azka, E., Escher, B.I., Plewa, M.J., Leusch, F.D., 2015. In vitro cytotoxicity and Yang, M., Zhang, X., Liang, Q., Yang, B., 2019. Application of (LC/) MS/MS precursor
adaptive stress responses to selected haloacetic acid and halobenzoquinone ion scan for evaluating the occurrence, formation and control of polar haloge-
water disinfection byproducts. Chem. Res. Toxicol. 28 (10), 2059e2068. nated DBPs in disinfected waters: a review. Water Res. 158, 322e337.
Ratpukdi, T., Sinorak, S., Kiattisaksiri, P., Punyapalakul, P., Siripattanakul- Yin, J., Wu, B., Liu, S., Hu, S., Gong, T., Cherr, G.N., Zhang, X.X., Ren, H., Xian, Q., 2018.
Ratpukdi, S., 2019. Occurrence of trihalomethanes and haloacetonitriles in Rapid and complete dehalogenation of halonitromethanes in simulated
water distribution networks of Khon Kaen Municipality, Thailand. Water Supply gastrointestinal tract and its influence on toxicity. Chemosphere 211,
19 (6), 1748e1757. 1147e1155.
Scarlett, K., Collins, D., Tesoriero, L., Jewell, L., van Ogtrop, F., Daniel, R., 2016. Effi- Yu, Y., Ma, X., Chen, R., Li, G., Tao, H., Shi, B., 2019. The occurrence and trans-
cacy of chlorine, chlorine dioxide and ultraviolet radiation as disinfectants formation behaviors of disinfection byproducts in drinking water distribution
against plant pathogens in irrigation water. Eur. J. Plant Pathol. 145 (1), 27e38. systems in rural areas of eastern China. Chemosphere 228, 101e109.

14
S. Kali, M. Khan, M.S. Ghaffar et al. Environmental Pollution 281 (2021) 116950

Zhai, H., Zhang, X., 2011. Formation and decomposition of new and unknown polar Zhang, Z., Zhu, Q., Huang, C., Yang, M., Li, J., Chen, Y., Yang, B., Zhao, X., 2020.
brominated disinfection byproducts during chlorination. Environ. Sci. Technol. Comparative cytotoxicity of halogenated aromatic DBPs and implications of the
45 (6), 2194e2201. corresponding developed QSAR model to toxicity mechanisms of those DBPs:
Zhai, H., He, X., Zhang, Y., Du, T., Adeleye, A.S., Li, Y., 2017. Disinfection byproduct binding interactions between aromatic DBPs and catalase play an important
formation in drinking water sources: a case study of Yuqiao reservoir. Che- role. Water Res. 170, 115283.
mosphere 181, 224e231. Zhao, Y., Yang, H.W., Liu, S.T., Tang, S., Wang, X.M., Xie, Y.F., 2016. Effects of metal
Zhai, H., Zhang, X., Zhu, X., Liu, J., Ji, M., 2014. Formation of brominated disinfection ions on disinfection byproduct formation during chlorination of natural organic
byproducts during chloramination of drinking water: new polar species and matter and surrogates. Chemosphere 144, 1074e1082.
overall kinetics. Environ. Sci. Technol. 48 (5), 2579e2588. Zheng, J., Su, C., Zhou, J., Xu, L., Qian, Y., Chen, H., 2017. Effects and mechanisms of
Zhang, D., Chu, W., Yu, Y., Krasner, S.W., Pan, Y., Shi, J., et al., 2018. Occurrence and ultraviolet, chlorination, and ozone disinfection on antibiotic resistance genes
stability of chlorophenylacetonitriles: a new class of nitrogenous aromatic DBPs in secondary effluents of municipal wastewater treatment plants. Chem. Eng. J.
in chlorinated and chloraminated drinking waters. Environ. Sci. Technol. Lett. 5 317, 309e316.
(6), 394e399. Zhong, D., Yuan, Y., Ma, W., Cui, C., Wu, Y., 2012. Influences of pipe materials and
Zhang, M., Ma, H., Wang, H., Du, T., Liu, M., Wang, Y., et al., 2019. Effects of ion hydraulic conditions on the process of trihalomethanes formation in water
species on the disinfection byproduct formation in artificial and real water. distribution network. Desalination and Water treatment 49 (1e3), 165e171.
Chemosphere 217, 706e714. Zimoch, I., Szymura, E., Moraczewska-Majkut, K., 2016. Changes of trihalomethanes
Zhang, Y., Chu, W., Yao, D., Yin, D., 2017. Control of aliphatic halogenated DBP (THMs) concentration in water distribution system. Desalination and Water
precursors with multiple drinking water treatment processes: formation po- Treatment 57 (3), 1399e1408.
tential and integrated toxicity. J. Environ. Sci. 58, 322e330.

15

You might also like