You are on page 1of 31

Journal of Environmental Management 299 (2021) 113611

Contents lists available at ScienceDirect

Journal of Environmental Management


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

Pesticide residues in drinking water, their potential risk to human health


and removal options
Ibrahim El-Nahhal a, Yasser El-Nahhal b, *
a
Université de Toulon - CS, 60584, France
b
Dept. of Earth and Environmental Science Faculty of Science, The Islamic University, Gaza, Palestine

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

Keywords: The application of pesticides in agricultural and public health sectors has resulted in substantially contaminated
Contaminated water water resources with residues in many countries. Almost no reviews have addressed pesticide residues in
Pesticide residues drinking water globally; calculated hazard indices for adults, children, and infants; or discussed the potential
Health risk
health risk of pesticides to the human population. The objectives of this article were to summarize advances in
Health hazards
Pesticide removal options
research related to pesticide residues in drinking water; conduct health risk assessments by estimating the daily
intake of pesticide residues consumed only from drinking water by adults, children, and infants; and summarize
options for pesticide removal from water systems. Approximately 113 pesticide residues were found in drinking
water samples from 31 countries worldwide. There were 61, 31, and 21 insecticide, herbicide, and fungicide
residues, respectively. Four residues were in toxicity class IA, 14 residues were in toxicity class IB, 55 residues
were in toxicity class II, 17 residues were in toxicity class III, and 23 residues were in toxicity class IV. The
calculated hazard indices (HIs) exceeded the value of one in many cases. The lowest HI value (0.0001) for
children was found in Canada, and the highest HI value (30.97) was found in Egypt, suggesting a high potential
health risk to adults, children, and infants. The application of advanced oxidation processes (AOPs) showed
efficient removal of many pesticide classes. The combination of adsorption followed by biodegradation was
shown to be an effective and efficient purification option. In conclusion, the consumption of water contaminated
with pesticide residues may pose risks to human health in exposed populations.

1. Introduction various residues.


It is well documented in the literature that contamination-free water
1.1. Water is an essential part of a healthy, balanced diet because the body relies on
it to function properly. The human body consists of approximately
Clean water is an important part of ecosystems and plays an 50–80% water. All chemical processes that take place in the human body
important role in the sustainability of life on earth. It is a well- occur in water. Human bodies need contamination-free water to digest
established fact that access to clean water is a fundamental human food, absorb nutrients, remove waste metabolites, and maintain the
right and vital to sustain healthy life. Water contamination may pose a human body at an isometric body temperature.
risk to many types of organisms, including humans. There is great Water systems may receive anthropogenic pollution in several ways,
concern about water contamination from pesticide residues. such as from chemical industry discharge (Cho et al., 2020; Posthuma
Several efforts have been broadly applied to protect water bodies. et al., 2020), wastewater discharge (Gómez-Sanabria et al., 2020;
Some have succeeded in some countries and failed in others due to the El-Nahhal et al., 2020), agricultural wastes (Wu et al., 2020; Hodomihou
growing demand for food due to population growth. et al., 2020; Garrido et al., 2020; Natasha et al., 2020; Zhi et al., 2020)
To date, the growing demand for food has resulted in the intensive and pesticide residues (Chowdhury et al., 2012; Flores-García et al.,
use of pesticides in agricultural systems for insect, weed, and fungal 2011; Varca, 2012; Donald et al., 2007; Teklu et al., 2015; Wan et al.,
control. Despite the success that pesticide application achieves in terms 2019, 2021; Tröger et al., 2021; Mahai et al., 2021).
of food security, it results in contamination of water systems with

* Corresponding author. Gaza, P. O. Box 108, Palestine.


E-mail address: y_el_nahhal@hotmail.com (Y. El-Nahhal).

https://doi.org/10.1016/j.jenvman.2021.113611
Received 7 March 2021; Received in revised form 13 August 2021; Accepted 23 August 2021
Available online 6 September 2021
0301-4797/© 2021 Elsevier Ltd. All rights reserved.
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

1.2. Occurrence of pesticide residues in water soil samples collected from areas around previous pesticide storage fa­
cilities and in arable soils of Central and Eastern European countries. In
Pesticide residues may become present in water through one or more addition, He et al. (2020) found considerable concentrations of
of the pathways described in the following sections. dichlorodiphenyltrichloroethane (DDT) and hexachlorocyclohexane
(HCH) in the sediments of the Meijiang River Basin. Bento et al. (2019)
1.2.1. Pesticide residues in agricultural ecosystems found that glyphosate and aminomethylphosphonic Acid (AMPA) con­
Pesticides are used in agriculture to control insects, nematodes, ro­ tents were 1.1–17.5 times higher in water-eroded sediment than in soil.
dents, weeds, and fungi. Currently, more than 1100 pesticides (in­
secticides, acaricides, herbicides, fungicides, and rodenticides) are 1.2.5. Dust fall
marketed in European countries (Jeong et al., 2012; Affum et al., 2018; Dust comprises fine solid particles composed of organic and inor­
Lu et al., 2017; Burri et al., 2019) for public use. ganic molecules. Dust may have a variety of shapes and sizes and can be
The public health sector (PHS) uses permethrin, malathion, piper­ carried by winds to different distances based on its size and wind speed
onyl butoxide, and isododecane in homes, schools, offices and stores to and direction. Dust can carry pesticide residues due to adsorption forces.
control insects of public health importance (Giddings et al., 2016; Seven insecticides (carbaryl, propoxur, chlorpyrifos, diazinon, cyflu­
Ankley and Collyard, 1995; Rubin et al., 2002; Markowitz, 1992; Cen­ thrin, cypermethrin, and permethrin), five herbicides (2,4-dichlor­
ters for Disease Control and Prevention, 2011; Liu et al., 2018; Buhl ophenoxyacetic acid [2,4-D], chlorthal, dicamba, mecoprop, and
et al., 2010; Health Consultation, 2010). Additionally, the PHS uses simazine), and one synergist (piperonyl butoxide) were identified in
indoxacarb, dinotefuran, and cyantraniliprole outdoors to control house ≥25% of carpet dust samples (Deziel et al., 2015). Additionally, five
flies (Zahn et al., 2019) and mosquitos (Hustedt et al., 2020; Rohani insecticides (chlorpyrifos, diazinon, permethrin, allethrin, and cyper­
et al., 2020). methrin) and one synergist (piperonyl butoxide) were identified in
Although pesticide application is necessary to achieve high yields of household and urban dust samples (Quirós-Alcalá et al., 2011). Addi­
products, its application has resulted in food and water contamination tionally, six insecticides (carbaryl, chlorpyrifos, chlorthal-dimethyl,
(El-Nahhal 2004, 2020) and continuous damage to ecosystems because diazinon, iprodione, and phosmet) and one herbicide (simazine) were
more than 90% of applied pesticides do not reach the target species identified in carpet dust samples collected within 1250 m of 89 resi­
(Llorent-Martínez et al., 2011). dences in California (Gunier et al., 2011). Moreover, insecticides used in
Water is an important component of public health, and the failure to urban settings, such as fipronil, have been determined in urban dust (Lin
supply safe drinking water results in a substantial health burden on et al., 2009) and in indoor and outdoor dust (Mahler et al., 2009; Shi
humanity (van der Kooij, 2014). et al., 2020). Therefore, dust can be additional source of drinking water
contamination. To date, pyrethroid insecticides have been detected in
1.2.2. Aerial application of pesticides dust from the driveway, curb gutter and street at rates of 53.5–94.8%,
The aerial application of pesticides may contaminate surrounding with median concentrations of 1–46 ng/g (Richards et al., 2016). Urban
areas with macrodroplets and/or microdroplets (drift) of pesticide res­ and residential concrete surfaces are often treated with pesticides to
idues. Several authors have demonstrated that pesticide sprayers influ­ control ants (Jiang et al., 2014).
ence pesticide deposition and distribution far from spraying sites. For
instance, Suarez-Lopez et al. (2020) indicated that pesticide drift caused 1.2.6. Rainfall
acetylcholine esterase inhibition in children living within 1000 m of a Rainfall and runoff characteristics can affect pesticide export from
greenhouse. An et al. (2020) detected myclobutanil and tebuconazole in small headwater catchments due to factors such as pesticide application
soil samples 5 m away from the application site. Wang et al. (2020) dose and timing and the variation in pesticide stocks in soil and soil type.
indicated that wind speed and droplet size play a critical role in the drift Intense rainfall-runoff events, which can fragment soil, control the
contamination potential of water resources. Furthermore, the applica­ export of simazine and tetraconazole (Imfeld et al. (2020). Heavy
tion of ultralow-volume aerosols of insecticides to control mosquitoes rainfall enhances the release of rodenticides from baited sewer
results in contaminated water resources (Britch et al., 2018; Chasko­ systems and outdoor surfaces into receiving streams (Regnery et al.,
poulou et al., 2018; Zhang et al., 2015; Mount, 1998). 2019, 2020). Additionally, water irrigation contributes to pesticide
losses from agricultural lands to water systems (Whelan et al., 2020).
1.2.3. Indoor and outdoor application
Indoor and outdoor application of pesticides causes critical 1.2.7. Runoff
contamination of water bodies. Storm water runoff infiltration systems may deteriorate the quality of
For instance, Lu et al. (2020) revealed cascading effects caused by groundwater. In fact, storm water runoff contains suspended particulate
fenoxycarb in freshwater systems dominated by Daphnia carinata and matter and many ionic organic molecules, such as pesticides and de­
Dolerocypris sinensis. Bao et al. (2020) found that a low concentration of tergents, that can contaminate groundwater. For instance, in one study,
carbendazim in water systems negatively affected zebrafish (Danio concentrations of carbendazim, diuron, fluopyram, imidacloprid, and
rerio). lamotrigine significantly increased in groundwater impacted by infil­
tration, indicating that storm water runoff infiltration systems
1.2.4. Soil erosion contribute to groundwater contamination (Pinasseau et al., 2020). In
Soil contains clay minerals, organic matter, and colloidal particles addition, concentrations of diuron, metolachlor, chlorpyrifos, simazine,
that can determine the fate of pesticides. These materials can actively galaxolide, and triallate were found to have increased in storm water
adsorb pesticides from the aqueous phase in soil (El-Nahhal et al., 1998; recycled via aquifers (Page et al., 2014). Similarly, in another study,
Nir et al., 2000). Dry conditions may result in soil erosion and the concentrations of desethylatrazine, simazine, atrazine, terbuthylazine,
transfer of soil particles from one place to another, resulting in the diuron, metolachlor, and diazinon increased in artificial recharge of
transfer of the soil components that adsorb pesticides to aquatic systems. aquifers used for drinking water production in many countries (Kuster
Rheinheimer Dos Santos et al. (2020) have found five herbicides (2,4-D, et al., 2010). Moreover, high concentrations of 3,4-dichloroaniline,
atrazine, desethyl-atrazine, simazine, and nicosulfuron), three fungi­ molinate, desethylatrazine, dimethoate, simazine, atrazine, metola­
cides (carbendazim, epoxiconazole, and tebuconazole), and one insec­ chlor and chlorpyrifos are present in the Ebro River basin area (Claver
ticide (imidacloprid) in epilithic biofilms due to high rainfall rates and et al., 2006). Additionally, considerable concentrations of terbutryn,
inadequate soil use. Ukalska-Jaruga et al. (2020a, 2020b) identified a diuron, imidacloprid, 2,4-D, and diazinon have been identified in agri­
wide range of organochlorine pesticides (OCPs) (0.01–21,888 μg/kg) in cultural runoff (García-Galán et al., 2020). Furthermore, considerable

2
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

concentrations of four pyrethroids (bifenthrin, permethrin, cyfluthrin, of pesticides for human health risk control. Recently, Olisah et al. (2020)
and cyhalothrin) and fipronil were detected in wash-off water from reviewed the occurrence of organochlorine pesticide residues in bio­
concrete surfaces, indicating an additional source of water contamina­ logical and environmental matrices in Africa. These reviews focused on
tion (Jiang et al., 2010). Additionally, urban surface runoff has been narrow scales and locations, and they did not evaluate the health risk
found to be contaminated with considerable concentrations of fiproles among adults, children, and infants or discuss the potential health risks
and their degradation products (Cryder et al., 2019). of pesticide residues found in water. The objectives of this article are to
summarize the advances in research related to pesticide residues in
1.2.8. Air pollution drinking water; estimate the daily intake and health hazards of pesticide
Pesticides that have high volatility may move via air systems and residues consumed only from drinking water by adults, children, and
substantially contaminate water systems. For instance, Kim et al. (2020) infants; and summarize pesticide removal options from water systems.
found higher concentrations of total (Σ3) endosulfan (50.3–611 pg/m3,
mean: 274 pg/m3) in air samples than in water and soil samples during 2. Methodology
summer.
Similarly, Jeon et al. (2019) determined the concentration of hex­ 2.1. Data collection
abromocyclododecanes (HBCDs) in air, water, soil, and sediment sam­
ples in South Korea and found that the Ʃ3 HBCDs were 22–133 pg/m3, Following the procedure reported by Prosser et al. (2020) with slight
10–128 ng/g, 0.2–151 ng/L, and 0.5–552 ng/g for these samples, modification, an extensive search of scholarly databases, such as Scopus,
respectively. However, Zhou et al. (2020) determined various levels of Web of Science, Science Direct, PubMed, BMC, Research Gate, and
neonicotinoid insecticides in respirable fine particulate matter (PM2.5), Google Scholar, was conducted using the following specific keywords:
suggesting risk factors for human health. ‘pesticide residues in drinking water’, ‘insecticide residues and drinking
water’, ‘fungicide residues and drinking water’, ‘herbicide residues and
1.3. Potential health risk associated with contaminated water drinking water’, ‘water and pesticide contamination’ [(organochlorine
insecticides and water) (organophosphate residues and drinking
It has been shown that exposure to pesticide residues may cause water)], ‘water contamination and pesticides’, ‘human health effect­
reproductive toxicity to both men and women (El-Nahhal, 2020), s/impacts/risks’ and ‘pesticides and health risk’.
disruption in the endocrine system (Miranda-Contreras et al., 2013),
enhancement of human breast cancer (Yang et al., 2020) and prostate 2.2. Inclusion and exclusion criteria
cancer (Pardo et al., 2020), Parkinson’s disease (Pouchieu et al., 2018),
and cardiovascular toxicity (El-Nahhal and El-Nahhal, 2021). The full texts of the articles were downloaded and carefully
reviewed. Articles that identified pesticide residues in drinking water
1.4. Pesticide residue removal options samples collected from different locations in a country, on different
continents, and in different drinking wells/or well survey processes in a
Method for removing pesticide residues from water include physical, country were considered relevant to this review and included in the
chemical and/or biological methods. For instance, physical methods study. Articles that developed analytical methods and validated them for
include adsorption and/or filtration using clay and organoclay (Boyd pesticide monitoring programs were considered relevant to the review
et al., 1988; Boyd and Jaynes, 1992; Nir et al., 2000; El-Nahhal and Safi, and included in the study. Review articles on pesticides and water were
2010), activated carbon (Alves et al., 2019; Gamboa-Carballo et al., considered relevant and included. Articles that developed and validated
2016), nanoparticles (Brooks et al., 2012), and plant materials (El-Ba­ analytical methods and/or techniques to determine pesticide residues in
kouri et al., 2007). Chemical methods include photochemical reactions water samples in the laboratory were excluded. Articles that surveyed
or zerovalent iron that produce hydrogen peroxide radicals that strongly pesticide residues in agricultural water systems, rivers, and/or lakes
interact with organic pollutants and convert them to nontoxic fragments were considered irrelevant and excluded.
(Buxton et al., 1988; Munter, 2001; Stefan, 2017). Additionally, bio­
logical methods use living plants, fungi and/or bacteria to remove 2.3. Data processing and calculations
pesticide residues from contaminated water (Escoto et al., 2019 Lucas
et al., 2008). Pesticide residue information was collected from the included arti­
Furthermore, Snow et al. (2020) detected high concentrations of cles and classified into insecticide, herbicide and fungicide residue
γ-HCH, p,p’-DDE, and dieldrin in water samples collected from the Syr groups.
Darya River. Li et al. (2020) detected the concentration profiles of The majority of the collected articles presented pesticide residues as
organochlorine pesticides in surface seawater and lower atmosphere gas averages, whereas the minority of articles presented pesticide residues
samples between 18◦ N and 40◦ S in the open Pacific Ocean in as a range, minimum, maximum, 25th percentile, 50th percentile and
2006–2007. They found higher concentrations of α-HCH, γ-HCH, 75th percentile. The presented averages in the majority of articles were
trans-chlordane, and cis-chlordane in the Northern Hemisphere than in used to estimate the optimal daily intake (ODI). This value was denoted
the Southern Hemisphere in both the air and water phases. Wu et al. as the optimal scenario.
(2020) found high concentrations of HCHs and low concentrations of p, Some articles presented data as minima, maxima and/or percentiles
p’- and o,p’-DDT in near east Antarctica and in the Ross Sea. (the minority of articles), and the average was calculated from the
A few reviews have focused on pesticide residues in drinking water presented data.
samples at a global level. For instance, Morrissey et al. (2015) reviewed The average concentration of pesticide residue presented as mg/L
global surface water contamination with neonicotinoids and character­ was multiplied by 1000 as a conversion factor to convert it to μg/L.
ized the associated risk to aquatic invertebrates, where they recom­
mended a threshold of 0.013–0.067 μg/L for imidacloprid, which is 2.4. Health risk assessment and estimation of pesticide daily intake
quite different from the USEPA 2017 recommendations.
Bruce-Vanderpuije et al. (2019) characterized the status of persistent The health risk assessment was conducted by estimating pesticide
organic pollutants, whereas Mohammed et al. (2019) reviewed insecti­ intake from drinking water using data on average pesticide residue ob­
cide residues in different matrices in Ghana. Elibariki and Maguta tained from the literature and water consumption for adults, children,
(2017) characterized the status of pesticide pollution in Tanzania. Li and and infants. Hazard quotients (HQs) and hazard indices (HIs) were
Jennings (2017) reviewed the worldwide regulations of standard values calculated based on equations (1)–(3).

3
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

2.4.1. Optimal daily intake (ODI) Health Organization (WHO) guidelines (2009).
The ODI of the pesticide residue was estimated according to equation
(1), 2.7.3. Statistical analysis of pesticide residue means
Based on recently published studies and other modifications
[ODI] = ([APS] × Q)/BW (1) (El-Nahhal, 2020), the reference average (Ref Aver) of the insecticide,
where [APS], Q, and BW are the average concentration of a pesticide herbicide or fungicide concentrations (μg/L) was calculated by summing
residue found in a drinking water sample (μg/L), the amount of water the corresponding concentrations (e.g., insecticide residues) in all
consumed by a person and body weight (kg), respectively. Water con­ countries and dividing them by the corresponding total number of res­
sumption equals 2, 1.5 and 0.75 L for adults, children and infants, idues (e.g., insecticide residues).
respectively, and the BW of adults, children and infants is 60, 15 and 5 The country average insecticide, herbicide or fungicide concentra­
kg, respectively (WHO, 2017). tion (μg/L) was calculated by summing the corresponding concentra­
tions (e.g., insecticide residues) in a country (e.g., Lebanon) and
2.5. Characterization of HQ and HI dividing them by the number of insecticides detected in the water
samples. The calculation mentioned above was used for insecticides,
HQ was estimated according to equation (2): herbicides, and fungicides. The relative average (Rel Aver) in each
country was calculated according to equation (5).
HQ = [ODI]/ARfD (2)
Country average
Rel Aver = (5)
where HQ is the health quotient and ARfD is the acute reference dose of Reference average
pesticide residue expressed in μg/L/day. The use of ARfD was previously
Based on the results of equation (5), water samples from different
reported (Bommuraj et al., 2019; US EPA, 2000). The value of ARfD was
countries were subdivided into four groups as follows:
obtained from the Pesticides Properties DataBase (PPDB) (2007) and its
Group 1 included countries with Rel Aver values below 0.01, group 2
updates.
included countries with Rel Aver values between 0.01 and 0.1, group 3
When a water sample contains more than one pesticide residue, the
included countries with Rel Aver values between 0.11 and 0.5, and
HQ is calculated individually for each sample. HI values ≥ 1 indicate
group 4 included countries with Rel Aver values ≥ 0.51.
additive effects and a high risk, whereas values < 1 indicate low or
To estimate the significant differences between the Rel Aver values of
negligible health risk (US EPA, 2000).
various countries, the reference standard deviation (Ref Stdev) was
HI=HQ1+HQ2+ HQ3+HQ4+HQ5+ … HQn (3) calculated from the general formula of the calculation. Then, the relative
standard deviation of a country was calculated according to equation
When a sample contains only one pesticide residue, (6):
HI=HQ (4) Country Stdev
Rel Stdev = (6)
Reference Stdev
In fact, the HQ was added to equation (3) because a mixture of
pesticide residues in which each individual residue is present in the Then, the Rel Stdev value was used to estimate the relative standard
mixture at a level approximating the no observed effect level elicits a error (Rel Std Error) from equation (7) as follows:
measurable response denoted as a joint additive effect. Thus, the sum­
Rel Stdev
mation of individual effects is given in equation (3). This is in accor­ Rel Std Error = √ ̅̅̅ (7)
2
n
dance with the US Environmental Protection Agency (EPA) (2000) and
Tinwell and Ashby (2004), who emphasized the joint additive effects of where n is the number of pesticide residues used to calculate the Rel
chemical mixtures. Aver and/or Rel Stdev in a country. The Rel Std Error values were used
to indicate the error bars in the figures. The overlapping error bars
2.6. Prediction of HI due to repeated consumption indicate no significant difference in the relative means at a p-value ≤
0.05, whereas no overlap indicates a significant difference in the Rel
Using the equations mentioned above, the HI was estimated at 1, 10, Aver in the studied countries.
25, 50 and 75 times the consumption of drinking water during the year
by multiplying the ODI by a factor of 10, 25, 50, and 75, respectively.
3. Results and discussion
Then, the HI was calculated according to the equations mentioned
above. This approach provided useful information to predict the future
Approximately 39 articles that detected pesticide residues in 31
HI due to the consumption of contaminated water.
countries were included in this review. These articles were published
during 2000–2021, except two articles published in the 1990s. More
2.7. Meta-analysis of pesticide residues
than one relevant article from some countries, such as China, India,
Portugal and the USA, was included in this review to integrate the
2.7.1. Concentration groups
collected information and to obtain a clear picture of the data. The
Concentrations of pesticide residues found in the water samples were
countries included in the study were Brazil, Burkina Faso, Canada, China
subdivided into three groups. Group 1 included concentrations above
(4); Co^te d’Ivoire, Egypt, Ethiopia, France, Germany, Ghana (2); India
the ARfD value, denoted as high concentrations. Group 2 included
(2); Iran (2); Ireland, Japan, Lebanon, Luxembourg, Mexico, Nicaragua,
concentrations equal to or five times lower than the ARfD value, denoted
Nigeria, Philippines, Portugal (2); Slovenia, South Africa, Spain,
as medium concentrations, and group 3 included concentrations below
Thailand, Togo (Africa), Turkey, USA, Vietnam, and West Bengal. The
those in group 2, denoted as diluted concentrations.
number in brackets indicates the included articles from a country.

2.7.2. Classification of pesticide residues according to toxicity class and


function 3.1. Pesticide residues in drinking water
The pesticide residues found in the water samples from a country
were subdivided into three groups according to their functions: insec­ 3.1.1. Organochlorine (OC) insecticide residues
ticide, fungicide, and herbicide residues. Each group was subdivided The average concentrations of the OC residues found in water sam­
into four toxicity classes (Ia, Ib, II, III and IV) according to the World ples in several countries are shown in Table 1. Approximately 1116

4
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 1
Average concentration of insecticide residues in drinking water.
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

α-HCH 0.01 6.00 0.20 0.59 0.88 0.00 0.00 0.00 27 14 China Wu et al. (2014)
β-HCH 0.01 6.00 0.48 1.45 2.18 0.00 0.00 0.00
γ-HCH 0.01 0.30 0.29 0.88 1.32 0.00 0.00 0.00
δ-HCH 0.01 6.00 0.49 1.46 2.18 0.00 0.00 0.00
p, p′ -DDE 0.05 0.50 1.56 4.68 7.02 0.00 0.01 0.01
p, p′ -DDD 0.00 0.50 0.03 0.10 0.16 0.00 0.00 0.00
p, p′ -DDT 0.00 0.50 0.15 0.46 0.70 0.00 0.00 0.00
Heptachlor 0.02 0.50 0.73 2.20 3.30 0.00 0.00 0.01
Heptachlor- epo 0.01 0.50 0.41 1.23 1.85 0.00 0.00 0.00
HCB 0.02 6.00 0.55 1.65 2.48 0.00 0.00 0.00
Aldrin 0.03 0.03 0.92 2.76 4.14 0.03 0.09 0.14
Dieldrin 0.01 0.05 0.45 1.34 2.01 0.01 0.03 0.04
Endrin 0.01 0.30 0.38 1.15 1.73 0.00 0.00 0.01
Endrin aldehyde 0.01 0.30 0.21 0.62 0.92 0.00 0.00 0.00
HI 0.05 0.15 0.22
α-HCH 0.002 6.00 0.07 0.21 0.31 0.00 0.00 0.00 17 8 China Liu et al. (2016)
β-HCH 0.003 6.00 0.09 0.27 0.40 0.00 0.00 0.00
γ-HCH 0.001 0.30 0.05 0.15 0.22 0.00 0.00 0.00
δ-HCH 0.000 6.00 0.00 0.00 0.01 0.00 0.00 0.00
p,p’-DDE 0.002 0.50 0.06 0.19 0.29 0.00 0.00 0.00
p,p’-DDD 0.000 0.50 0.01 0.03 0.05 0.00 0.00 0.00
o,p’-DDT 0.001 0.50 0.04 0.11 0.16 0.00 0.00 0.00
p,p’-DDT 0.001 0.50 0.04 0.12 0.18 0.00 0.00 0.00
0.001 0.003 0.005
α-HCH 0.11 6 3.50 10.50 15.75 0.00 0.00 0.00 5 7 China Li et al. (2012)
β-HCH 0.52 6 17.33 52.00 78.00 0.00 0.01 0.01
γ-HCH 2.65 0.3 88.33 265.00 397.50 0.29 0.88 1.33
δ-HCH 0.17 6 5.67 17.00 25.50 0.00 0.00 0.00
o,p’-DDT 0.02 0.5 0.75 2.25 3.38 0.00 0.00 0.01
p,p’ -DDD 0.05 0.5 1.67 5.00 7.50 0.00 0.01 0.02
p,p’-DDT 0.03 0.5 0.92 2.75 4.13 0.00 0.01 0.01
HI 0.31 0.92 1.37
α-HCH 0.004 6 0.14 0.41 0.61 0.000 0.000 0.000 27 13 China Zhou et al. (2008)
β-HCH 0.018 6 0.60 1.81 2.71 0.000 0.000 0.000
γ-HCH 0.003 0.3 0.09 0.26 0.40 0.000 0.001 0.001
δ-HCH 0.008 6 0.28 0.84 1.26 0.000 0.000 0.000
Heptachlor 0.012 0.5 0.40 1.21 1.81 0.001 0.002 0.004
Aldrin 0.007 0.03 0.23 0.68 1.02 0.008 0.023 0.034
Heptachlor epoxide 0.012 0.5 0.39 1.18 1.77 0.001 0.002 0.004
p,p’-DDE 0.002 0.5 0.07 0.20 0.30 0.000 0.000 0.001
Diedrin 0.003 0.05 0.12 0.35 0.52 0.002 0.007 0.010
Endrin 0.002 0.3 0.08 0.23 0.34 0.000 0.001 0.001
p, p’-DDD 0.001 0.5 0.03 0.08 0.11 0.000 0.000 0.000
o, p’-DDT 0.001 0.5 0.02 0.07 0.11 0.000 0.000 0.000
p, p’-DDT 0.004 0.5 0.13 0.39 0.59 0.000 0.001 0.001
HI 0.013 0.038 0.057
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

α-HCH 0.01 6 0.40 1.20 1.80 0.00 0.00 0.00 38 11 India Kaushik et al. (2012)
γ-HCH 0.07 0.3 2.32 6.95 10.43 0.01 0.02 0.03
β-HCH 0.20 6 6.79 20.36 30.54 0.00 0.00 0.01
α-Endosulfan 0.18 6 5.86 17.57 26.36 0.00 0.00 0.00
β-Endosulfan 0.24 6 7.86 23.59 35.39 0.00 0.00 0.01
o,p′ -DDT 0.03 0.5 1.04 3.12 4.68 0.00 0.01 0.01
p,p′ -DDT 0.32 0.5 10.70 32.09 48.13 0.02 0.06 0.10
o,p′ -DDE 0.03 0.5 0.91 2.72 4.08 0.00 0.01 0.01
p,p′ -DDE 0.17 0.5 5.54 16.63 24.95 0.01 0.03 0.05
o,p′ -DDD 0.04 0.5 1.48 4.43 6.65 0.00 0.01 0.01
p,p′ -DDD 0.34 0.5 11.49 34.48 51.71 0.02 0.07 0.10
HI 0.07 0.22 0.33
α-HCH 0.09 6 3.11 9.32 13.98 0.00 0.00 0.00 60 12 India Mohapatra et al. (1995)
β-HCH 0.06 6 2.01 6.04 9.06 0.00 0.00 0.00
γ-HCH 0.07 0.3 2.32 6.97 10.45 0.01 0.02 0.03
pp’ DDT 0.18 0.5 5.95 17.85 26.78 0.01 0.04 0.05
op’-DDT 0.08 0.5 2.61 7.83 11.75 0.01 0.02 0.02
pp’-DDT 0.02 0.5 0.73 2.20 3.30 0.00 0.00 0.01
Op’-DDE 0.01 0.5 0.23 0.70 1.05 0.00 0.00 0.00
α-Endosulfan 0.03 6 1.10 3.30 4.95 0.00 0.00 0.00
β-Endosulfan 0.01 6 0.34 1.03 1.54 0.00 0.00 0.00
(continued on next page)

5
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 1 (continued )
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

Aldrin 0.09 0.03 3.01 9.03 13.55 0.10 0.30 0.45


Dieldrin 0.02 0.05 0.53 1.60 2.40 0.01 0.03 0.05
Heptachlor 0.02 0.5 0.56 1.68 2.52 0.00 0.00 0.01
HI 0.14 0.42 0.63
a-HCH 0.014 6 0.47 1.40 2.10 0.000 0.000 0.000 18 Turkey Bulut et al. (2010)
b-HCH 0.159 6 5.30 15.90 23.85 0.001 0.003 0.004
g-HCH 0.014 0.3 0.47 1.40 2.10 0.002 0.005 0.007
d-HCH 0.034 6 1.13 3.40 5.10 0.000 0.001 0.001
Heptachlor 0.031 0.5 1.03 3.10 4.65 0.002 0.006 0.009
Aldrin 0.02 0.3 0.67 2.00 3.00 0.002 0.007 0.010
Heptachlor epoxide 0.007 0.5 0.23 0.70 1.05 0.000 0.001 0.002
a-Endosulfan 0.01 6 0.33 1.00 1.50 0.000 0.000 0.000
4,4′ -DDE 0.013 0.5 0.43 1.30 1.95 0.001 0.003 0.004
Dieldrin 0.011 0.05 0.37 1.10 1.65 0.007 0.022 0.033
Endrin 0.01 0.3 0.33 1.00 1.50 0.001 0.003 0.005
B-Endosulfan 0.009 6 0.30 0.90 1.35 0.000 0.000 0.000
4,4′ -DDD 0.025 0.5 0.83 2.50 3.75 0.002 0.005 0.008
Endrin aldehyde 0.027 0.3 0.90 2.70 4.05 0.003 0.009 0.014
Endosulfan sulfate 0.022 6 0.73 2.20 3.30 0.000 0.000 0.001
4,4′ -DDT 0.085 0.5 2.83 8.50 12.75 0.006 0.017 0.026
Endrin Ketone 0.02 0.3 0.67 2.00 3.00 0.002 0.007 0.010
Methoxychlor 0.032 5 1.07 3.20 4.80 0.000 0.001 0.001
0.030 0.089 0.134
δ-HCH 0.02 6 0.67 2.00 3.00 0.00 0.00 0.00 15 22 Lebanon Chaza et al. (2018)
2,4′ -DDD 12.82 0.5 427.33 1282.00 1923.00 0.85 2.56 3.85
4,4′ -DDE 0.03 0.5 1.00 3.00 4.50 0.00 0.01 0.01
4,4′ -DDT 0.3 0.5 10.00 30.00 45.00 0.02 0.06 0.09
2,4′ -DDT 0.39 0.5 13.00 39.00 58.50 0.03 0.08 0.12
4,4′ -DDD 0.15 0.5 5.00 15.00 22.50 0.01 0.03 0.05
Aldrin 2.3 0.03 76.67 230.00 345.00 2.56 7.67 11.50
Heptachlor epoxide 9.01 0.5 300.33 901.00 1351.50 0.60 1.80 2.70
Dieldrin 2.72 0.05 90.67 272.00 408.00 1.81 5.44 8.16
Endosulfan 0.05 6 1.67 5.00 7.50 0.00 0.00 0.00
Endrin 0.28 0.3 9.33 28.00 42.00 0.03 0.09 0.14
Endrin aldehyde 0.06 0.3 2.00 6.00 9.00 0.01 0.02 0.03
Endosulfan sulfate 0.63 6 21.00 63.00 94.50 0.00 0.01 0.02
Metoxychlor 0.13 5 4.33 13.00 19.50 0.00 0.00 0.00
Heptachlor 4.22 0.5 140.67 422.00 633.00 0.28 0.84 1.27
O,O,O-Triethylphosphate 0.33 100 11.00 33.00 49.50 0.00 0.00 0.00
Thionazin 2.13 100 71.00 213.00 319.50 0.00 0.00 0.00
Dimethoate 3.78 10 126.00 378.00 567.00 0.01 0.04 0.06
Disulfoton 2.34 100 78.00 234.00 351.00 0.00 0.00 0.00
Methyl-parathion 27.72 6 924.00 2772.00 4158.00 0.15 0.46 0.69
Parathion 2.08 6 69.33 208.00 312.00 0.01 0.03 0.05
Famphur 5.34 100 178.00 534.00 801.00 0.00 0.01 0.01
HI 6.39 19.16 28.74
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

α-HCH 0.03 6 0.88 2.63 3.94 0.00 0.00 0.00 16 15 South Africa Fatoki and Awofolu (2004)
HCB 0.03 6 1.01 3.04 4.56 0.00 0.00 0.00
β-HCH 0.02 6 0.56 1.69 2.53 0.00 0.00 0.00
δ-HCH 0.02 6 0.64 1.91 2.86 0.00 0.00 0.00
Heptachlor 0.01 0.5 0.40 1.20 1.80 0.00 0.00 0.00
Aldrin 0.01 0.03 0.25 0.75 1.13 0.01 0.03 0.04
g-Chlordane 0.02 0.5 0.69 2.06 3.09 0.00 0.00 0.01
2,4-DDE 0.01 0.5 0.26 0.77 1.16 0.00 0.00 0.00
Endosulfan 0.02 6 0.62 1.85 2.78 0.00 0.00 0.00
Dieldrin 0.01 0.05 0.19 0.57 0.86 0.00 0.01 0.02
2,4-DDD 0.01 0.5 0.18 0.55 0.83 0.00 0.00 0.00
Endrin 0.01 0.3 0.49 1.47 2.21 0.00 0.00 0.01
4,4-DDD 0.01 0.5 0.45 1.34 2.01 0.00 0.00 0.00
2,4-DDT 0.01 0.5 0.20 0.60 0.90 0.00 0.00 0.00
4,4-DDT 0.02 0.5 0.63 1.89 2.84 0.00 0.00 0.01
HI 0.02 0.06 0.09
a-HCH 0.06 6 2.00 6.00 9.00 0.00 0.00 0.00 36 14 Mexico Díaz et al., 2009
b-HCH 0.09 6 3.00 9.00 13.50 0.00 0.00 0.00
c-HCH 0.026 0.3 0.87 2.60 3.90 0.00 0.01 0.01
d-HCH 0.028 6 0.93 2.80 4.20 0.00 0.00 0.00
Heptachlor 0.039 0.5 1.30 3.90 5.85 0.00 0.01 0.01
Aldrin 0.021 0.3 0.70 2.10 3.15 0.00 0.01 0.01
(continued on next page)

6
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 1 (continued )
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

Heptachlor epoxide 0.005 6 0.17 0.50 0.75 0.00 0.00 0.00


Endosulfan 0.003 6 0.10 0.30 0.45 0.00 0.00 0.00
p,p’-DDE 0.049 0.5 1.63 4.90 7.35 0.00 0.01 0.01
Endrin 0.005 0.3 0.17 0.50 0.75 0.00 0.00 0.00
p,p’-DDD 0.002 0.5 0.07 0.20 0.30 0.00 0.00 0.00
Endrin aldehyde 0.004 0.3 0.13 0.40 0.60 0.00 0.00 0.00
Endosulfan sulfate 0.024 6 0.80 2.40 3.60 0.00 0.00 0.00
p,p’-DDT 0.004 0.5 0.13 0.40 0.60 0.00 0.00 0.00
00 0.03 0.04
α-HCH 0.294 6 9.80 29.40 44.10 0.00 0.00 0.01 8 16 West Bengal Mondal et al. (2018)
β-HCH 0.563 6 18.77 56.30 84.45 0.00 0.01 0.01
γ-HCH 0.026 0.3 0.87 2.60 3.90 0.00 0.01 0.01
δ-HCH 1.106 6 36.87 110.60 165.90 0.01 0.02 0.03
2,4 ́ -DDE 0.101 0.5 3.37 10.10 15.15 0.01 0.02 0.03
4,4 ́ -DDE 0.049 0.5 1.63 4.90 7.35 0.00 0.01 0.01
2,4 ́ -DDD 0.014 0.5 0.47 1.40 2.10 0.00 0.00 0.00
4,4 ́ -DDD 0.059 0.5 1.97 5.90 8.85 0.00 0.01 0.02
2,4 ́ -DDT 0.091 0.5 3.03 9.10 13.65 0.01 0.02 0.03
4,4 ́ -DDT 0.526 0.5 17.53 52.60 78.90 0.04 0.11 0.16
α-endosulfan 0.068 6 2.27 6.80 10.20 0.00 0.00 0.00
β-endosulfan 0.055 6 1.83 5.50 8.25 0.00 0.00 0.00
Chlorpyrifos 0.092 5 3.07 9.20 13.80 0.00 0.00 0.00
Methyl parathion 0.095 6 3.17 9.50 14.25 0.00 0.00 0.00
Monocrotophos 0.009 0.3 0.30 0.90 1.35 0.00 0.00 0.00
Phorate 0.027 25 0.90 2.70 4.05 0.00 0.00 0.00
HI 0.073 0.218 0.327
γ-HCH 0.04 0.3 1.33 4.00 6.00 0.00 0.01 0.02 64 6 Ghana Fosu-Mensah et al. (2016)
α-Endosulfan 0.03 6 1.00 3.00 4.50 0.00 0.00 0.00
Dieldrin 0.03 0.05 1.00 3.00 4.50 0.02 0.06 0.09
DDT 0.04 0.5 1.33 4.00 6.00 0.00 0.01 0.01
Endosulfanulfate 0.04 6 1.33 4.00 6.00 0.00 0.00 0.00
Heptachlor 0.027 0.5 0.90 2.70 4.05 0.00 0.01 0.01
HI 0.03 0.09 0.13
β-HCH 0.01 6.00 0.33 1.00 1.50 0.00 0.00 0.00 44 13 Ghana Affum et al. (2018)
p,p′ -DDT 0.04 0.50 1.17 3.50 5.25 0.00 0.01 0.01
Endrin 0.01 0.30 0.33 1.00 1.50 0.00 0.00 0.01
Methoxychlor 0.01 5.00 0.33 1.00 1.50 0.00 0.00 0.00
Methamidophos 0.01 3.00 0.43 1.30 1.95 0.00 0.00 0.00
Fenpropathrin 0.04 30.00 1.27 3.80 5.70 0.00 0.00 0.00
L-cyhalothrin 0.01 5 0.33 1.00 1.50 0.00 0.00 0.00
Permethrin 0.03 250.00 1.03 3.10 4.65 0.00 0.00 0.00
Cyfluthrin 0.01 25 0.47 1.40 2.10 0.00 0.00 0.00
Cypermethrin 0.04 100 1.27 3.80 5.70 0.00 0.00 0.00
Deltamethrin 0.05 10.00 1.50 4.50 6.75 0.00 0.00 0.00
Chlorpyrifos 0.41 5 13.73 41.20 61.80 0.00 0.01 0.01
Ethoprophos 0.02 50.00 0.73 2.20 3.30 0.00 0.00 0.00
HI 0.007 0.020 0.030
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

Dieldrin 0.028 0.05 0.93 2.80 4.20 0.02 0.06 0.08 4 10 Philippines Navarrete et al. (2018)
Endrin aldehyde 0.74 0.3 24.67 74.00 111.00 0.08 0.25 0.37
α-HCH 0.039 6 1.30 3.90 5.85 0.00 0.00 0.00
β-HCH 0.036 6 1.20 3.60 5.40 0.00 0.00 0.00
δ-HCH 0.288 0.03 9.60 28.80 43.20 0.32 0.96 1.44
γ-Chlordane 0.02 0.5 0.67 2.00 3.00 0.00 0.00 0.01
Endosulfan 0.056 6 1.87 5.60 8.40 0.00 0.00 0.00
Heptachlor 0.029 0.5 0.97 2.90 4.35 0.00 0.01 0.01
Heptachlor epoxide 0.022 0.5 0.73 2.20 3.30 0.00 0.00 0.01
Methoxychlor 0.026 5 0.87 2.60 3.90 0.00 0.00 0.00
0.427 1.280 1.919
γ-HCH 0.3 0.3 10.00 30.00 45.00 0.03 0.10 0.15 7 9 Egypt Derbalah et al. (2014)
Heptachlor 0.18 0.5 6.00 18.00 27.00 0.01 0.04 0.05
Aldrin 6.00 0.03 200.00 600.00 900.00 6.67 20.00 30.00
Heptachlor epo. 0.23 0.5 7.67 23.00 34.50 0.02 0.05 0.07
Endosulfan 0.1 6 3.33 10.00 15.00 0.00 0.00 0.00
Dieldrin 0.14 0.05 4.67 14.00 21.00 0.09 0.28 0.42
Endrin 0.26 0.3 8.67 26.00 39.00 0.03 0.09 0.13
DDD 0.32 0.5 10.67 32.00 48.00 0.02 0.06 0.10
p,p′ -DDT 0.16 0.5 5.33 16.00 24.00 0.01 0.03 0.05
HI 6.88 20.65 30.97
(continued on next page)

7
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 1 (continued )
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

p,p -DDE

0.06 0.5 2.00 6.00 9.00 0.00 0.01 0.02 3 14 Nicaragua McConnell et al. (1999)
p,p′ -DDD 0.1 0.5 3.33 10.00 15.00 0.01 0.02 0.03
o,p′ -DDT 0.2 0.5 6.67 20.00 30.00 0.01 0.04 0.06
p,p′ -DDT 1.2 0.5 40.00 120.00 180.00 0.08 0.24 0.36
β-HCH 0.7 6 23.33 70.00 105.00 0.00 0.01 0.02
γ-HCH 0.02 0.3 0.67 2.00 3.00 0.00 0.01 0.01
Toxaphene 10.07 100 335.67 1007.00 1510.50 0.00 0.01 0.02
Chlordimeform 2.6 100 86.67 260.00 390.00 0.00 0.00 0.00
Chlorpyrifos 0.3 5 10.00 30.00 45.00 0.00 0.00 0.00
Diazinon 0.07 30 2.33 7.00 10.50 0.00 0.00 0.00
Fenthion 4.5 7 150.00 450.00 675.00 0.02 0.06 0.10
Fenthion sulf. 1.4 7 46.67 140.00 210.00 0.01 0.02 0.03
Mephosfolan 0.7 500 23.33 70.00 105.00 0.00 0.00 0.00
Methyl parathion 0.85 6 28.33 85.00 127.50 0.00 0.01 0.02
HI 0.15 0.44 0.67
,4-DDE 0.02 0.5 0.67 2.00 3.00 0.00 0.00 0.01 9 7 Togo -Africa Mawussi et al. (2009)
4,4-DDT 0.03 0.5 1.00 3.00 4.50 0.00 0.01 0.01
Dieldrin 0.02 0.05 0.67 2.00 3.00 0.01 0.04 0.06
Heptaclor 0.33 0.5 11.00 33.00 49.50 0.02 0.07 0.10
Hepta epoxide 0.09 0.5 3.00 9.00 13.50 0.01 0.02 0.03
α-Endosulfan 0.05 6 1.67 5.00 7.50 0.00 0.00 0.00
β-Endosulfan 0.02 6 0.67 2.00 3.00 0.00 0.00 0.00
HI 0.05 0.14 0.20
Aldrin 0.37 0.03 12.33 37.00 55.50 0.41 1.23 1.85 9 5 Co^te d’Ivoire Wandan and Zabik (1996)
p,p’-DDT 0.47 0.5 15.67 47.00 70.50 0.03 0.09 0.14
Dieldrin 0.38 0.05 12.67 38.00 57.00 0.25 0.76 1.14
Endosulfan 1.73 6 57.67 173.00 259.50 0.01 0.03 0.04
γ-HCH 2.61 0.3 87.00 261.00 391.50 0.29 0.87 1.31
HI 1.00 2.99 4.48
γ-HCH 0.0004 0.3 0.01 0.04 0.07 0.000 0.000 0.000 40 2 Canada Woudneh et al. (2009)
Heptachlor 0.0000 0.5 0.00 0.00 0.00 0.000 0.000 0.000
Chlorpyriphos 0.0001 5 0.00 0.01 0.01 0.000 0.000 0.000
Phosmet 0.0000 45 0.00 0.00 0.00 0.000 0.000 0.000
Pirimiphos 0.0000 100 0.00 0.00 0.00 0.000 0.000 0.000
HI 0.000 0.000 0.000
heptachlor 0.039 0.5 1.30 3.90 5.85 0.00 0.01 0.01 3 2 Ireland McManus et al. (2013)
Lindane 0.042 0.3 1.40 4.20 6.30 0.00 0.01 0.02
HI 00 0.02 0.03
Chlordane 0.22 0.5 7.33 22.00 33.00 0.01 0.04 0.07 15 4 USA Eitzer and Chevalier (1999)
Lindane 0.06 0.3 2.00 6.00 9.00 0.01 0.02 0.03
Chlorpyrifos 0.06 5 2.00 6.00 9.00 0.00 0.00 0.00
Diazinon 0.02 30 0.67 2.00 3.00 0.00 0.00 0.00
HI 0.02 0.07 0.10
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

Azadirachtin 0.20 750.00 6.80 20.41 30.62 0.00 0.00 0.00 106 9 Burkina Faso Lehmann et al. (2017)
Carbofuran 0.06 5 2.01 6.02 9.03 0.00 0.00 0.00
Chlorpyrifos 0.06 5 1.84 5.51 8.27 0.00 0.00 0.00
λ-Cyhalothrin 0.03 5 0.98 2.94 4.41 0.00 0.00 0.00
Cypermethrin 0.22 100.00 7.32 21.97 32.96 0.00 0.00 0.00
Dieldrin 0.11 0.05 3.56 10.69 16.04 0.07 0.21 0.32
Imidacloprid 0.01 57 0.23 0.68 1.02 0.00 0.00 0.00
Profenofos 0.05 1000.00 1.63 4.90 7.35 0.00 0.00 0.00
Triazophos 0.001 1.00 0.04 0.13 0.20 0.00 0.00 0.00
0.07 0.22 0.33
carbaryl 0.170 10 5.65 16.95 25.43 0.001 0.002 0.003 4 16 Japan Tanabe et al. (2001)
chlorfenvinphos 0.012 5 0.40 1.20 1.80 0.000 0.000 0.000
chlorpyrifos 0.009 5 0.28 0.85 1.28 0.000 0.000 0.000
cyanophos 0.016 100 0.54 1.63 2.45 0.000 0.000 0.000
diazinon 0.040 25 1.33 4.00 6.00 0.000 0.000 0.000
dichlofenthion 0.010 100 0.33 1.00 1.50 0.000 0.000 0.000
dichlorvos 0.051 2 1.68 5.05 7.58 0.001 0.003 0.004
dimethoate 0.087 10 2.89 8.67 13.00 0.000 0.001 0.001
etofenprox 0.041 1000 1.37 4.10 6.15 0.000 0.000 0.000
fenitrothion 0.089 13 2.96 8.88 13.31 0.000 0.001 0.001
fenobucarb 0.074 100 2.45 7.35 11.03 0.000 0.000 0.000
malathion 0.026 300 0.85 2.55 3.83 0.000 0.000 0.000
propoxur 0.032 20 1.06 3.17 4.75 0.000 0.000 0.000
pyridaphenthion 0.040 100 1.34 4.03 6.04 0.000 0.000 0.000
isofenphos oxon 0.044 1 1.47 4.40 6.60 0.001 0.004 0.007
(continued on next page)

8
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 1 (continued )
3
RN CF ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L
Ad Ch In Ad Ch In

HI 0.004 0.011 0.017


3-Hydroxycarbofuran, 2.95 100 98.33 295.00 442.50 0.00 0.00 0.00 15 Spain Ccanccapa et al. (2016)
Methiocarb, 1.35 5 45.11 135.33 203.00 0.01 0.03 0.04
Imidacloprid 3.99 57 132.94 398.83 598.25 0.00 0.01 0.01
Azinphos Methyl 1.57 75 52.44 157.33 236.00 0.00 0.00 0.00
Chlorfenvinphos 10.83 20 360.89 1082.67 1624.00 0.02 0.05 0.08
Chlorpyrifos, 4.87 5 162.22 486.67 730.00 0.03 0.10 0.15
Diazinon, 6.94 30 231.22 693.67 1040.50 0.01 0.02 0.03
Diclofenthion, 16.40 500 546.78 1640.33 2460.50 0.00 0.00 0.00
Dimethoate, 21.95 10 731.56 2194.67 3292.00 0.07 0.22 0.33
Fenitrothion, 11.61 13 387.10 1161.29 1741.93 0.03 0.09 0.13
Fenthion, 1.01 100 33.67 101.00 151.50 0.00 0.00 0.00
Fenthion Sulfone, 1.80 100 59.89 179.67 269.50 0.00 0.00 0.00
Malathion 5.40 300 179.89 539.67 809.50 0.00 0.00 0.00
Omethoate 2.36 2 78.67 236.00 354.00 0.04 0.12 0.18
Parathion-Ethyl, 9.88 6 329.44 988.33 1482.50 0.07 0.20 0.30
HI 0.29 0.85 1.26
Dichlorvos 0.257 40 8.57 25.70 38.55 0.00 0.00 0.00 64 7 Iran Heidar et al. (2017)
Diazinon 0.21 30 7.10 21.30 31.95 0.00 0.00 0.00
Methyl parathion 0.57 6 19.10 57.30 85.95 0.00 0.01 0.01
Fenitrothion 0.28 40 9.47 28.40 42.60 0.00 0.00 0.00
Malathion 0.37 300 12.23 36.70 55.05 0.00 0.00 0.00
Profenofos 0.32 1000 10.50 31.50 47.25 0.00 0.00 0.00
Ethion 0.35 0.5 11.60 34.80 52.20 0.02 0.07 0.10
HI 0.03 0.08 0.12
Diazinon 57.53 30 1917.67 5753.00 8629.50 0.06 0.19 0.29 24 2 Iran Shayeghi et al. (2007)
malathion 38.69 300 1289.67 3869.00 5803.50 0.00 0.01 0.02
HI 0.07 0.20 0.31
Carbofuran 0.39 5 13.08 39.25 58.88 0.00 0.01 0.01 23 8 Brazil Albuquerque et al., 2016,
Imidacloprid 1.28 57 42.67 128.00 192.00 0.00 0.00 0.00
Chlorpyrifos 0.26 5 8.50 25.50 38.25 0.00 0.01 0.01
Dimethoate 2.56 10 85.17 255.50 383.25 0.01 0.03 0.04
Ethion 0.12 15 4.00 12.00 18.00 0.00 0.00 0.00
Malathion 0.22 300 7.48 22.43 33.64 0.00 0.00 0.00
Profenofos 0.01 1000 0.20 0.60 0.90 0.00 0.00 0.00
Permethrin 0.82 1500 27.17 81.50 122.25 0.00 0.00 0.00
HI 0.01 0.04 0.06
Diazinon 2.95 30 98.25 294.76 442.14 0.00 0.01 0.01 45 2 Ethiopia Mekonen et al. (2016)
Malathion 9.75 300 324.90 974.70 1462.05 0.01 0.03 0.05
HI 0.014 0.042 0.064
Fenobucarb 0.04 100 1.33 4.00 6.00 0.00 0.00 0.00 54 4 Vietnam Toan et al. (2013)
Fipronil 0.16 3 5.33 16.00 24.00 0.00 0.01 0.01
Buprofezin 0.12 500 4.00 12.00 18.00 0.00 0.00 0.00
Profenofos 0.04 1000 1.33 4.00 6.00 0.00 0.00 0.00
HI 0.002 0.005 0.008

Dimethoate 0.001 10 0.03 0.08 0.12 0.00 0.00 0.00 81 3 Portugal Palma et al. (2009)

Malathion 0.002 300 0.07 0.21 0.32 0.00 0.00 0.00


Diazinon 0.003 30 0.08 0.25 0.38 0.00 0.00 0.00
0.000 0.000 0.000
Diazinon 0.002 30 0.07 0.22 0.33 0.00 0.00 0.00 2 Portugal Palma et al. (2014)
Dimethoate 0.004 10 0.12 0.35 0.53 0.00 0.00 0.00
0.000 0.000 0.000
Diazinon 0.083 30 2.77 8.30 12.45 0.000 0.000 0.000 2 Nigeria Sosan et al., (2008)
Propoxur 0.029 20 0.97 2.90 4.35 0.000 0.001 0.001
HI 0.000 0.001 0.002
Chlorpyrifos 0.11 5 3.67 11.00 16.50 0.00 0.00 0.00 100 3 Thailand Jaipieam et al., (2009)
Dicrotophos 0.145 40 4.83 14.50 21.75 0.00 0.00 0.00
Profenofos 0.258 1000 8.60 25.80 38.70 0.00 0.00 0.00
HI 0.001 0.003 0.004
Diazinon 3.46 30 115.45 346.35 519.53 0.004 0.01 0.02 2 Germany Stehlea et al. 2019
Chlorpyrifos 0.18 5 5.85 17.55 26.33 0.001 0.00 0.01
HI 0.005 0.01 0.03

water samples from 31 countries around the world were analyzed for OC insecticides (14 compounds), organophosphorus (OP) insecticides
pesticide residues. Approximately 58 insecticide residues in addition to (27 compounds), carbamate (CT) insecticides (five compounds), pyre­
some metabolites were detected in the water samples at different fre­ throid (PY) insecticides (eight compounds), neonicotinoids (N) (one
quencies. These insecticides belong to different chemical groups, such as compound) and other chemical groups denoted as others (O) (three

9
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

compounds). Turgut et al. (2009), who found a high concentration (14.3%) of DDT in
Fig. 1a shows insecticide residues that were detected two times or dicofol formulations. Additionally, Pazi et al. (2011) found high con­
more. DDT and its analogs were detected 53 times among all residues, centrations of DDT and other OC residues in sediments from Izmir Bay
whereas disulfoton was detected only one time. The OC residues and from the Gediz River estuary, indicating OC insecticide usage,
detected in water samples included fourteen parent compounds, aldrin, probably linked to public health emergencies. Similarly, the distribution
chlordane, DDT, dieldrin, endrin, α-endosulfan, β-endosulfan, HCH, of OC residues in Lebanon may have emerged from agricultural prac­
heptachlor, lindane, methoxychlor, toxaphene, γ-chlordane, and γ-HCH tices, high rainfall and movement of OC-contaminated sediment (Kou­
(α-HCH, β-HCH, γ-HCH, δ-HCH), and four metabolites, 1,1-dichloro-2,2- zayha et al., 2013). However, the occurrence of OC residues in water
bis(p-chlorophenyl)ethylene (DDE), 1,1-dichloro-2,2-bis(p-chloro­ samples in South Africa may have emerged from the use of OC in­
phenyl)ethane (DDD), heptachlor-epo, and endrin aldehyde. secticides in public health sectors. This is in agreement with a report by
Approximately 42 OC residues with a concentration range of Buah-Kwofie and Humphries (2017), who found high concentrations of
0.001–2.65 μg/L were detected in the water samples from China (Wu OC residues (e.g., DDT, HCH, Aldrin, endrin endosulfan and heptachlor)
et al., 2014; Liu et al., 2016; Li et al., 2012; Zhou et al., 2008), and the in sediments collected from the lakes and wetlands of iSimangaliso.
concentration of one residue exceeded the ARfD value. Twenty-three OC Buah-Kwofie and Humphries attributed these high levels of OC residues
residues with a concentration range of 0.01–0.34 μg/L were detected in to the application of DDT and other insecticides for malaria control.
the water samples from India (Kaushik et al., 2012; Mohapatra et al., Furthermore, the occurrence of OC residues in water samples from
1995). Eighteen OC residues with a concentration range of 0.007–0.159 the USA may emerge from the direct application of OC insecticide on
were identified in the water samples from Turkey (Bulut et al., 2010). cotton farms and contaminated air. This is in accordance with Hoh and
Fifteen OC residues were detected in the water samples from Lebanon, Hites (2004), who identified cotton farms and the Gulf of Mexico as
with a concentration range of 0.02–12.82 μg/L, and concentrations of potential sources of toxaphene or DDEs. Similarly, Alegria et al. (2006)
five residues exceeded the ARfD value (Chaza et al., 2018). The water found elevated levels of some OC pesticides (DDTs, chlordane, toxa­
samples from South Africa had a concentration range of 0.01–0.03 μg/L phene) in the ambient air of Chiapas, Mexico, during 2000–2001, sug­
(Fatoki and Awofolu, 2004). Fourteen OC residues were detected in the gesting a potential source of water contamination through air exchange
water samples from Mexico, with a concentration range of 0.003–0.09 processes. The occurrence of OC residues in water samples from
μg/L (Díaz et al., 2009). Twelve OC residues were detected in the water Nicaragua may emerge from erosion of contaminated soil with OC res­
samples from West Bengal, with a concentration range of 0.014–1.106 idues to water resources (Lacayo-Romero et al., 2006), whereas in the
μg/L (Mondal et al., 2018). Ten OC residues were detected in the water Philippines, DDT and DDE sediment residues were slowly released into
samples from Ghana, with a concentration range of 0.01–0.04 μg/L sterile and nonsterile brackish water systems. DDT residues are easily
(Fosu-Mensah et al., 2016; Affum et al., 2018), and from the Philippines, adsorbed in sediments (95.0%) and slowly released in hard water (8.4%)
with a concentration range of 0.02–0.74 μg/L (Navarrete et al., 2018). and seawater (0.1%) (Bajet and Navarro, 2002). Studies from Ghana
Nine OC residues were detected in the water samples from Egypt, with a revealed high levels of DDT and its isomers in mosquito breeding sites
concentration range of 0.10–6.00 μg/L (Derbalah et al., 2014). Seven OC (Kudom et al., 2018) and contaminated sediments in the harbor (Botwe
residues with a concentration range of 0.02–10.07 μg/L were detected in et al., 2017), suggesting that a history of OC insecticide application for
the water samples from Nicaragua (McConnell et al., 1999) and from mosquito control could be the main source of contamination. In Canada,
Togo-Africa, with a concentration range of 0.02–0.33 μg/L (Mawussi Poole et al. (1998) indicated that the occurrence of OC residues in water
et al., 2009). Five OC residues with a concentration range of 0.37–2.61 samples could be from long-range atmospheric transport, whereas
μg/L were detected in the water samples from Co^te d’Ivoire (Wandan, Bidleman et al. (2002) found that volatilization of OC residues from seas
and Zabik, 1996). Two OC residues were detected in the water samples and large lakes can be a source of water contamination based on the
from Canada (Woudneh et al., 2009), Ireland (McManus et al., 2013) appearance of nonracemic alpha-HCH in the air boundary layer above
and the USA (Eitzer and Chevalier, 1999) with a concentration range of the water. However, Marvin et al. (2004) showed that historical
0.0004–0.22 μg/L. One OC residue was detected in the water samples contamination of sediments and large lakes could be a source of water
from Burkina Faso at a concentration of 0.11 μg/L (Lehmann et al., contamination with OC residues. However, Rawn et al. (2001) noted
2017). The occurrence of OC residues in water samples may be due to that glacial runoff could be a significant source of OC residues for water
direct application of OC insecticides in agricultural sections using contamination. More recently, Jantunen et al. (2015) highlighted the
OC-contaminated pesticides, such as in China, where application of rule of air-water gas exchange as a source of OC residues in drinking
Dicofol in cotton fields was a source of DDT contamination in northern water. Studies from Egypt (Wahaab, and Badawy, 2004) indicated that
Jiangsu Province (Yang et al., 2008). In addition, flooding practices in polluted lakes (Maryut and Manzala) and the Nile River could be the
rice fields may enhance the distribution of OC from contaminated soil to source of OC residue transfer to drinking water.
non-contaminated soil. Yao et al. (2007) emphasized this phenomenon. The names of the OC residues, their concentrations, and the ARfD
However, historical application and illegal use of technical DDT or values in each country are shown in Table 1. The highest concentrations
dicofol may still occur in some urban areas. A similar explanation was of OC residues (2,4′ -DDD, 12.82 μg/L; heptachlor epoxide, heptachlor
given for the distribution of DDT in Italian soil (Thiombane et al., 2018). 9.01; 4.22 μg/L; aldrin 6.0) were detected in water from Lebanon and
Bhadouria et al. (2012) indicated that concentrations of α-HCH, β-HCH, Egypt, and the lowest concentration was in water from Canada (hepta­
γ-HCH, δ-HCH, aldrin, dieldrin, heptachlor, hept.epoxide, endosulfan-I, chlor, 0.000525 μg/L). The highest values of OC residues occurred in
endosulfan-II, endo.sulfate, endrin, 4,4′ -DDE, 4,4′ -DDD, and DDT were water samples from Lebanon, and the lowest number occurred in water
high in the sediments in and out of Keoladeo National Park, an impor­ samples from Burkina Faso.
tant wintering ground in India. Additionally, contamination of surface The detection of OC residues in the water samples from the countries
soil with OC insecticides from agricultural application became a major mentioned above may be attributed to historical agricultural application
source of water pollution in India (Khuman et al., 2020). Additionally, of these compounds. Moreover, OC residues continue to cycle through
OC insecticides may still be in use in India and be a major source of various routes, such as atmospheric transport and runoff (Kim et al.,
water contamination. This finding is in accordance with a previous 2020), and to resist degradation and, thus, remain in the environment
report that indicated that OC insecticides are still being applied in India for longer periods (Yadav et al., 2015). Consequently, OC residues are
(Bakore et al., 2004; Agarwal et al., 2015). transported via environmental components and can be detected again in
The occurrence of OC residues in water samples from Turkey may water (Table 1) and/or accumulate in fruit and vegetables (Mohammed
emerge from using dicofol formulations contaminated with DDT or et al., 2019), honey (El-Nahhal, 2020) and plants (Panseri et al., 2014).
other OC insecticide residues. This is in accordance with a report by Additionally, DDT and/or its analogs have been used in the public health

10
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Fig. 1. a. Frequency of pesticide residues detection. Pesticide detection range is 2–50. The highest frequency detected pesticide is DDT (53 Times) and the lowest one
is tolchlofos methyl oxon. Pesticide residues detected only one time is not presented here. Fig. 1b. Number of pesticide residues detected in water samples from
different countries. Blue, orange and gray colors represent Insecticide, herbicide and fungicide residue detected respectively. (For interpretation of the references to
color in this figure legend, the reader is referred to the Web version of this article.)

11
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

sector, such as for head lice control (El-Nahhal and Hammad et al., (Suarez-Lopez et al., 2020). Similarly, ultralow-volume aerosols of in­
2019), and in the agriculture and public health sectors (Olisah et al., secticides (Britch et al., 2018) could be an additional source of water
2019b). contamination. Moreover, atmospheric dust can carry OP residues
Due to the high KOW values of OC residues, they remain in soil (chlorpyrifos and diazinon) through adsorption to water systems (Deziel
components for a long time and can be adsorbed in clay minerals or soil et al., 2015). Photodegradation, volatilization and hydrolysis enable the
organic matter due to their high hydrophobicity and then slowly mobility of OP residues to contaminate water systems (Kromer et al.,
released in water systems (Umulisa et al., 2020; Ukalska-Jaruga et al. 2004; Kiss and Virág, 2009).
(2020a, 2020b); Bhandari et al., 2020). OC residues may evaporate from The lower frequency of OP residue detection compared with that of
the soil to the atmosphere and diffuse throughout water systems, as OC residue can be explained by several factors, such as sensitivity to
noted in several published studies (Hearon et al., 2020; Kim et al., 2020; water hydrolysis at alkaline pH, as shown for chlorpyrifos and other
Jeon et al., 2019; Snow et al., 2020; Karlsson et al., 2000). organophosphorus esters in drinking water (pubmed.ncbi.nlm.nih.gov/
?sort=pubdate&size=20&term=Lockridge+O&cauthor_id=31401088
3.1.2. Organophosphate (OP) insecticide residues Lockridge et al., 2019). Moreover, OP residues may undergo photo­
The average concentrations of organophosphate (OP) residues in chemical degradation, as shown for methyl parathion (Weber et al.,
the water samples are presented in Table 1. Approximately 27 OP 2009; Bouhala et al., 2020), malathion (Ramos-Delgado et al., 2013),
residues were detected in the water samples from many countries. OP chlorpyrifos (pubmed.ncbi.nlm.nih.gov/?sort=pubdate&size=20&term
residues include azinphos-methyl, chlorfenvinphos, chlorpyrifos, =Femia+J&cauthor_id=24292479Femia et al., 2013), and biodegra­
cyanophos, diazinon, dichlofenthion, dichlorvos, dicrotophos, dimeth­ dation in water systems (Alvarenga et al., 2014).
oate, ethion, ethoprophos, fenitrothion, fenthion, isofenphos-oxon, In general, the detected concentrations of OP residues were higher
malathion, mephosfolan, methamidophos, monocrotophos, omethoate, than those of OC residues because OC residues are highly hydrophobic
parathion-ethyl, parathion-methyl, phosphate, photoset, profenofos, and have low solubility in water (1–10,000 μg/L) and very high lipid
pyridafenthion, thionazin, and triazophos. solubility (log P of 4–7) (PPDB, 2007).
For instance, fourteen OP residues were detected in water samples
from Japan (Tanabe et al., 2001). Twelve OP residues with a concen­ 3.1.3. Carbamate (CT) insecticide residues
tration range of 1.01–21.95 μg/L were identified in the water samples The means of the CT residues in the water samples are presented in
from Spain (Ccanccapa et al., 2016). Nine OP residues with a concen­ Table 1. Five CT residues, carbofuran, carbaryl, methiocarb, fenobucarb
tration range of 0.21–0.57 μg/L were identified in the water samples and propoxur, were detected in the water samples from Burkina Faso,
from Iran (Heidar et al., 2017; Shayeghi et al., 2007), and seven OP Brazil, Spain, and Vietnam. Carbofuran was detected three times with a
residues with a concentration range of 0.33–27.72 μg/L were found in concentration range of 0.06–2.95 μg/L. Carbaryl and methiocarb were
the water samples from Lebanon (Chaza et al., 2018). Six OP residues detected once in the water samples from Spain at concentrations of 0.17
with a concentration range of 0.07–4.5 μg/L were detected in the water and 1.35 μg/L, respectively. Fenobucarb was detected twice at concen­
samples from Nicaragua (McConnell et al., 1999), and five OP residues trations of 0.04 and 0.074 μg/L in Vietnam and Japan, respectively.
with a concentration range of 0.01–2.56 μg/L were detected in the water Propoxur residues were detected in Nigeria and Japan with a concen­
samples from Brazil (Albuquerque et al., 2016). tration range of 0.029–0.023 μg/L. However, some metabolites from
Four OP residues with a concentration range of 0.01–0.41 μg/L were group 3, such as hydroxyl-carbofuran, were detected three times at
detected in the water samples from Ghana (Affum et al., 2018), OP moderate to low concentrations (<5 μg/L), such as 0.06, 0.39, and 2.95
residues with a concentration range of 0.009–0.095 μg/L were found in μg/L in Burkina Faso, Brazil, and Spain, respectively. The occurrence of
the water samples from West Bengal (Mondal et al., 2018), and OP CT residues in water may be attributed to intensive use in agricultural
residues with a similar concentration range were detected in the water systems (Kunstadter et al., 2001; Ruggieri et al., 2008), erosion of
samples from Vietnam (Toan et al., 2013). contaminated soil (Ukalska-Jaruga et al. (2020a, 2020b)), dust fall
Three OP residues with a concentration range of 0.001–0.06 μg/L adsorbing OC residues (Deziel et al., 2015), wastewater (Karpuzcu et al.,
were detected in the water samples from Burkina Faso (Lehmann et al., 2014; Wu et al., 2019a, 2019b), leaching in the soil profile (Hulbert
2017), Portugal (Palma et al. 2009, 2014), and Canada (Woudneh et al., et al., 2020), wash-off from plants by rainfall (Willis et al., 1996) and
2009). Two OP residues with a concentration range of 0.06–0.22 μg/L contaminated rainfall (Vogel et al., 2008).
were detected in the water samples from the USA (Eitzer and Chevalier, The low frequency of carbamate detection in the water samples was
1999), two with a concentration range of 2.95–9.75 μg/L were detected due to its sensitivity to water hydrolysis (Katagi, 2010), photo­
in the water samples from Ethiopia (Mekonen et al., 2016), and 2 with a degradation (Derbalah et al., 2020) and biodegradation (Birolli et al.,
concentration range of 0.18–3.46 μg/L were detected in the water 2016).
samples from Germany (Stehle et al., 2019). Compared with OC resi­
dues, OP residues were less frequently detected in water samples. For 3.1.4. Pyrethroid (PY) neonicotinoid family (N) and other (O) insecticide
instance, diazinon, chlorpyrifos, malathion, dimethoate and fenthion residues
were detected 15, 10, 7, 6 and 4 times, respectively, in the water samples Eight PY residues were detected in the water samples (Table 1). Six
from Ghana, West Bengal, USA, Ethiopia, Iran, Brazil, Vietnam, Spain PY residues were detected at a concentration range of 0.01–0.41 μg/L in
Portugal and Ethiopia. the water samples from Ghana (Affum et al., 2018). Two PY residues
The occurrence of OP residues in drinking water may emerge from were detected in the water samples from Burkina Faso with a concen­
the intensive application of OP pesticides in agricultural systems, such as tration range of 0.001–0.22 μg/L (Lehmann et al., 2017). One PY residue
in China, where they use 1.5–4-fold higher concentrations than the (etofenprox) with a concentration of 0.041 μg/L was detected in the
world average concentration (Zhang et al., 2015). Additionally, the water samples from Japan (Tanabe et al., 2001).
application of high concentrations of pesticides to soil may result in the The low frequency of PY detection was because it is sensitive to
adsorption of pesticides on soil particles, and soil-adsorbed pesticides photochemical degradation (Rafique and Tariq, 2015). The occurrence
may enter the aquatic ecosystem via soil leaching (El-Nahhal and of PY residues in water may emerge from using large amounts of them in
Hamdona, 2017), erosion, surface runoff, and sewer overflow (Xu et al., both agricultural and urban settings (Taylor et al., 2019), the movement
2020). Moreover, the degradation of OP residues in ecosystems could be of PY-contaminated sediments to aquatic systems (Li et al., 2017), sus­
an additional source of drinking water contamination (Y. Zhang et al., pended sediment in surface water such as rivers and/or lakes (Hladik
2020). Furthermore, pesticide application methods, such as aerial and Kuivila, 2009; Cheng et al., 2017), and wastewater and associated
application, can tremendously be a serious source of contamination sediments (Li et al., 2011).

12
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Moreover, the detection of PY residues in the water samples from 2013; Wan et al., 2021).
Ghana and Brazil may have occurred because these two countries Seven herbicide residues were detected in the water samples from
extensively used pyrethroids to control mosquitos because of their high Lebanon with a concentration range of 0.26–1.73 μg/L (Chaza et al.,
activity as insecticides and their low mammalian toxicity (Varloud et al., 2018) and from Germany with a concentration range of 1.22–79.02 μg/L
2015; Willemin et al., 2015). (Stehle et al., 2019). Five herbicide residues were detected in the water
Five N residues (acetamiprid, clothianidin, imidacloprid, thiame­ samples from the USA with a concentration range of 0.03–1.8 μg/L
thoxam, and thiacloprid) at a concentration lower than 0.01 μg/L were (Postle et al., 2004; Eitzer and Chevalier, 1999) and from Luxembourg
detected in water samples from China (Mahai et al., 2021; Wan et al., with a concentration range of 0.001–0.03 μg/L (Bohn et al., 2011).
2019; Wang et al., 2020), Vietnam (Wan et al., 2021), Canada (Sultana Four herbicide residues were detected with a concentration range of
et al., 2018) and the USA (Klarich et al., 2017; Craddock et al., 2019). 0.0001–0.051 μg/L in water samples from Canada (Woudneh et al.,
However, only imidacloprid concentrations of 0.01, 1.28, 3.99 and 8.33 2009), and three herbicide residues were detected with a concentration
μg/L were detected in drinking water samples from Burkina Faso (Leh­ range of 0.001–0.021 μg/L in water samples from China (Shi et al., 2020;
mann et al., 2017), Brazil (Albuquerque et al., 2016), Spain (Ccanccapa Wang et al., 2020).
et al., 2016) and China (Mahai et al., 2021). Two residues were detected in the water samples from Vietnam and
The occurrence of N residues in drinking water may be due to the France, whereas one residue was detected in the water samples from
presence of N residues in many environmental samples, such as fruits Burkina Faso, West Bengal, Ethiopia and Iran. However, glyphosate is a
and vegetables (Lu et al., 2018), indoor dust (Salis et al., 2017), outdoor widely used herbicide, and it is considered a nontoxic pesticide to
dust (Forero et al., 2017), urine (Tao et al., 2019) and runoff loss from humans due to its high LD50 value (>2000 mg/kg) and class III toxicity.
contaminated watersheds (Lamers et al., 2011). It is highly soluble in water (10.5 g/L) and has a high dissociation
Three insecticide residues belonging to the different chemical groups constant (pKa = 2.32) and a low partitioning coefficient (Log P = − 3.2)
(others) were detected in the water samples from Vietnam, Spain, and (PPDB, 2007). For these reasons, less attention may have been given to
Nicaragua. These insecticides were buprofezin, chlordimeform, and glyphosate. Nevertheless, there is growing concern for potential toxicity
fipronil (Table 1). to aquatic organisms due to high solubility in water and high dissocia­
The concentrations of the insecticides listed in Table 1 were divided tion constant (see above). To date, glyphosate has been detected in many
into three groups as described in section 2.7.1. Group 1 included those water bodies, including drinking water, due to possible spillage, runoff,
that had residue concentrations above the ARfD values, such as aldrin, and leaching. Moreover, it has been detected in groundwater with a
γ-HCH (China); aldrin (India); 2,4′ -DDD, aldrin, heptachlor-epoxide, maximum concentration of 1.42 μg/L (Rendon-von Osten and
dieldrin, heptachlor, and parathion-methyl (Lebanon); 4,4́ –DDT (West Dzul-Caamal 2017).
Bengal); endrin-aldehyde and δ-HCH (Philippines); γ-HCH, aldrin, and The occurrence of herbicide residues in the water samples was due to
dieldrin (Egypt); p,p′ -DDT (Nicaragua); aldrin, dieldrin, and γ-HCH direct application to soil followed by water irrigation/rainfall resulting
(Co^te d’Ivoire); dieldrin (Burkina Faso); dimethoate, omethoate, and in herbicide leaching into the groundwater (El-Nahhal et al., 1998,
parathion-ethyl (Spain); and Diazinon (Iran). Almost 11 countries have 2000; Nir et al., 2000; El-Nahhal and Hamdona, 2017) or transport as
elevated levels of pesticide residues. colloidal particles in the soil (Terzaghi et al., 2020). Moreover, it has
Groups 2 and 3 included residues with concentrations below the been shown that soil erosion (Rheinheimer Dos Santos et al., 2020), dust
ARfD values. The insecticide residues listed in Table 1 and not fall (Deziel et al., 2015), rainfall runoff (Imfeld et al., 2020) and dry-wet
mentioned in this section were determined to be at either medium or low conditions have critical impacts on drinking water contamination.
concentrations. Thus, the concentrations of residues in groups 1 and 2
contributed to high HI values in some countries. 3.1.6. Fungicide residues
The concentrations of the fungicide residues determined in the water
3.1.5. Herbicide residues samples are listed below. Approximately 23 residues were detected in 32
Table 2 shows the concentrations of herbicide residues in water water samples from three countries. These residues were azoxystrobin,
worldwide. Approximately 31 parent herbicide residues belonging to carbendazim, chlorothalonil, difenoconazole, edifenphos, epox­
different chemical groups were detected in more than 768 water samples iconazole, etridiazole, flutolanil, flutriafol, imazalil, iprobenfos, iso­
from 18 countries (Brazil, Burkina Faso, Canada, China, Ethiopia, prothiolane, mepronil, pencycuron, picoxystrobin, probenazole,
France, Germany, Iran, Lebanon, Luxembourg, Portugal, Slovenia, prochloraz, quintozene, tebuconazole, thiabendazole, and tricyclazole.
Spain, USA, Vietnam, and West Bengal). These residues were 2,4-D, To date, eleven residues with a concentration range of 0.013–0.473 μg/L
acetochlor, alachlor, ametryn, atrazine, bentazone, butachlor, cloma­ have been detected in Japan (Tanabe et al., 2001). Eight residues with a
zone, cyanazine diuron, hexazinone, imazethapyr, isoproturon, glyph­ concentration range of 0.001–0.39 μg/L were detected in Brazil (Albu­
osate, linuron, mecoprop, metoalachlor, metribuzin, molinate, querque et al., 2016). Five fungicide residues with a concentration range
pendimethalin, prometon, propanil, propazine, simazine, tebuthiuron, of 4.82–101.03 μg/L were detected in the water samples from Spain
terbumeton, trifluralin, terbuthylazine, terbutryn, and pretilachlor. (Ccanccapa et al., 2016), and only one residue exceeded the ARfD value.
Additionally, seven herbicide metabolites (deethylatrazine, deisopro­ Additionally, the fungicide residues carbendazim, difenoconazole,
pylatrazine, deisopropylatrazine, desethylatrazine, desethyl terbuthy­ epoxiconazole, flutolanil, picoxystrobin, prochloraz, tebuconazole,
lazine, terbumeton-deethyl, terbuthylazine deethyl, and − 2 hydroxy- thiabendazole, and tricyclazole were detected in water samples from
terbutylene) were detected in the water samples. China with a concentration range of 0.0011–0.077 μg/L, and the fre­
For instance, twenty-eight herbicide residues were detected in the quency of detection reached 43.4% (n = 166 water samples) in some
water samples from Portugal with a concentration range of 0.002–0.27 cases (Li et al., 2020). Similar fungicide residues were also detected in
μg/L (Palma et al., 2009, 2014). Twenty-two herbicide residues were water samples from Malaysia (Elfikrie et al., 2020).
detected in the water samples from Brazil with a concentration range of The occurrence of fungicide residues in the drinking water samples
0.01–4.90 μg/L (Albuquerque et al., 2016; Dores et al., 2008). Fourteen was due to the widespread use of these chemicals in soil to control soil-
herbicide residues were detected in the water samples from Spain with a borne diseases, often as seed dressings and other times as foliar sprays to
concentration range of 1.16–32.32 μg/L (Ccanccapa et al., 2016). Eight control fungi during the growing seasons (Tosi et al., 2018). Further­
herbicide residues were detected in the water samples from Slovenia more, eroded sediments (Rheinheimer Dos Santos et al., 2020), indus­
with a concentration range of 0.0001–0.051 μg/L (Koroša et al., 2016). trial wastewater (Kapsi et al., 2019), contaminated estuaries (Kapsi
Eight herbicide residues were detected in the water samples from et al., 2019), and animal manure (Kreuzig et al., 2010) could be an
Vietnam with a concentration range of 0.0001–0.47 μg/L (Toan et al., additional source of water contamination with fungicide residues.

13
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 2
Average concentration of herbicide residues in drinking water.
3
RN CFμg/L ARfD μg/L Average daily intake ng/g *10− HQ AS DR Cn Au

Ad Ch In Ad Ch In

Simazine 0.033 5 1.11 3.33 5.00 0.00 0.00 0.00 81 12 Portugal Palma et al. (2009)
Isoproturon 0.003 15 0.11 0.33 0.50 0.00 0.00 0.00
Chlortoluron 0.008 40 0.26 0.78 1.17 0.00 0.00 0.00
Atrazine 0.270 100 9.01 27.02 40.54 0.00 0.00 0.00
Diuron 0.049 3 1.65 4.95 7.42 0.00 0.00 0.00
Cyanazine 0.000 100 0.01 0.04 0.06 0.00 0.00 0.00
Molinate 0.002 100 0.06 0.17 0.25 0.00 0.00 0.00
Linuron 0.001 30 0.04 0.11 0.16 0.00 0.00 0.00
Alachlor 0.009 10 0.31 0.94 1.41 0.00 0.00 0.00
Metolachlor 0.010 97 0.34 1.02 1.53 0.00 0.00 0.00
HI 0.001 0.003 0.004
Atrazine 0.005 100 0.18 0.54 0.81 0.00 0.00 0.00 NR 16 Portugal Palma et al. (2014)
Cyanazine 0.001 100 0.02 0.06 0.10 0.00 0.00 0.00
Simazine 0.008 5 0.28 0.84 1.26 0.00 0.00 0.00
Terbuthylazine 0.155 8 5.18 15.53 23.29 0.00 0.00 0.00
Chlortoluron 0.015 40 0.49 1.48 2.22 0.00 0.00 0.00
Diuron 0.004 3 0.14 0.42 0.63 0.00 0.00 0.00
Isoproturon 0.002 15 0.08 0.23 0.34 0.00 0.00 0.00
Linuron 0.001 30 0.04 0.13 0.20 0.00 0.00 0.00
Alachlor 0.001 10 0.04 0.13 0.20 0.00 0.00 0.00
Metolachlor 0.037 97 1.23 3.69 5.54 0.00 0.00 0.00
Propanil 0.001 70 0.03 0.10 0.15 0.00 0.00 0.00
Molinate 0.012 100 0.38 1.15 1.73 0.00 0.00 0.00
2,4-D 0.036 750 1.20 3.59 5.39 0.00 0.00 0.00
Bentazone 0.195 250 6.49 19.46 29.19 0.00 0.00 0.00
MCPA 0.025 150 0.83 2.48 3.72 0.00 0.00 0.00
Mecoprop 0.005 10 0.17 0.50 0.74 0.00 0.00 0.00
HI 0.001 0.003 0.004
2,4-D 1.85 750 61.67 185.00 277.50 0.00 0.00 0.00 23 16 Brazil Albuquerque et al., 2016,
Propanil 4.90 70 163.17 489.50 734.25 0.00 0.01 0.01
Bentazone 1.43 250 47.64 142.93 214.40 0.00 0.00 0.00
Acetochlor 0.01 20 0.25 0.75 1.13 0.00 0.00 0.00
Metolachlor 0.42 97 14.00 42.00 63.00 0.00 0.00 0.00
Alachlor 0.01 10 0.37 1.10 1.65 0.00 0.00 0.00
Pendimethalin 0.01 300 0.37 1.10 1.65 0.00 0.00 0.00
Trifuralin 1.16 15 38.67 116.00 174.00 0.00 0.01 0.01
Imazethapyr 0.65 440 21.67 65.00 97.50 0.00 0.00 0.00
Diuron 0.12 3 4.13 12.40 18.60 0.00 0.00 0.01
Clomazone 4.48 2500 149.24 447.73 671.60 0.00 0.00 0.00
Hexazinone 0.40 50 13.33 40.00 60.00 0.00 0.00 0.00
Metribuzin 0.07 20 2.39 7.18 10.76 0.00 0.00 0.00
Ametryn 0.75 15 24.89 74.68 112.02 0.00 0.00 0.01
Atrazine 1.06 100 35.26 105.79 158.69 0.00 0.00 0.00
Simazine 0.17 5 5.66 16.99 25.48 0.00 0.00 0.01
atrazine 0.063 100 2.10 6.30 9.45 0.000 0.000 0.000 23 6 Brazil Dores et al. (2008)
DEA 0.158 100 5.27 15.80 23.70 0.000 0.000 0.000
simazine 0.064 5 2.13 6.40 9.60 0.000 0.001 0.002
metribuzin 0.151 5 5.03 15.10 22.65 0.001 0.003 0.005
trifluralin 0.142 7.5 4.73 14.20 21.30 0.001 0.002 0.003
metolachlor 0.253 97 8.43 25.30 37.95 0.000 0.000 0.000
HI 0.012 0.037 0.060
Metoalachlor 2.17 97 72.33 217.00 325.50 0.00 0.00 00 NR 14 Spain Ccanccapa et al. (2016)
Atrazine, 7.45 100 248.22 744.67 1117.00 0.00 0.01 0.01
Deisopropylatrazine 7.94 100 264.67 794.00 1191.00 0.00 0.01 0.01
Deethylatrazine 13.88 100 462.50 1387.50 2081.25 0.00 0.01 0.02
Simazine, 27.31 5 910.44 2731.33 4097.00 0.18 0.55 0.82
Terbumeton, 3.55 75 118.44 355.33 533.00 0.00 0.00 0.01
Terbumeton-Deethyl, 3.68 75 122.56 367.67 551.50 0.00 0.00 0.01
Terbuthylazine 4.14 8 138.00 414.00 621.00 0.02 0.05 0.08
Terbuthylazine Deethyl, 1.16 8 38.56 115.67 173.50 0.00 0.01 0.02
Terbuthylazine-2 Hydroxy 4.41 8 147.00 441.00 661.50 0.02 0.06 0.08
Terbutryn 11.58 27 385.94 1157.83 1736.75 0.01 0.04 0.06
Diuron, 32.32 3 1077.44 3232.33 4848.50 0.36 1.08 1.62
Isoproturon 5.76 15 192.11 576.33 864.50 0.01 0.04 0.06
HI 0.62 1.87 2.80
Atrazine 0.0508 100 1.69 5.08 7.62 0.00 0.00 0.00 NR 8 Slovenia Koroša et al. (2016)
Desethylatrazine 0.051 100 1.70 5.10 7.65 0.00 0.00 0.00
Deisopropylatrazine 0.005 100 0.17 0.50 0.75 0.00 0.00 0.00
Simazine 0.0052 5 0.17 0.52 0.78 0.00 0.00 0.00
Terbuthylazine 0.0011 8 0.04 0.11 0.17 0.00 0.00 0.00
Desethylterbuthylazine 0.0026 8 0.09 0.26 0.39 0.00 0.00 0.00
Metolachlor 0.001 97 0.03 0.10 0.15 0.00 0.00 0.00
Propazine 0.001 100 0.03 0.10 0.15 0.00 0.00 0.00
HI 0.000 0.000 0.000

14
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

3
RN CF ARfD Average daily intake ng/g *10− HQ AS DR Cn Au
μg/L μg/L
Ad Ch In Ad Ch In

Isoprothiolane 0.17 100 5.67 17.00 25.50 0.000 0.000 0.000 54 2 Vietnam Toan et al. (2013)
Butachlor 0.47 100 15.67 47.00 70.50 0.000 0.000 0.001
Atrazine 0.02 100 0.66 1.99 2.99 0.000 0.000 0.000 39 6 Wan et al. (2021)
Simazine 0.00 5 0.00 0.01 0.01 0.000 0.000 0.000
Terbuthylazine 0.00 8 0.00 0.00 0.00 0.000 0.000 0.000
MCPA 0.00 150 0.01 0.04 0.06 0.000 0.000 0.000
2,4-D 0.00 750 0.01 0.03 0.05 0.000 0.000 0.000
Bentazone 0.00 250 0.00 0.01 0.01 0.000 0.000 0.000
HI 0.000 0.001 0.001
Molinate 1.08 100 36.00 108.00 162.00 0.00 0.00 0.00 15 7 Lebanon Chaza et al. (2018)
Atrazine 1.24 100 41.33 124.00 186.00 0.00 0.00 0.00
Acetochlor 1.53 20 51.00 153.00 229.50 0.00 0.01 0.01
Alachlor 1.73 10 57.67 173.00 259.50 0.01 0.02 0.03
Metolachlor 0.26 97 8.67 26.00 39.00 0.00 0.00 0.00
Butachlor 1.08 100 36.00 108.00 162.00 0.00 0.00 0.00
HI 0.01 0.03 0.04
Atrazine 7.01 100 233.63 700.90 1051.35 0.00 0.01 0.01 NR 7 Germany Stehlea et al., 2009
Simazine 79.02 5 2633.93 7901.80 11852.70 0.53 1.58 2.37
Prometon 2.12 100 70.57 211.70 317.55 0.00 0.00 0.00
Metolachlor 1.22 97 40.58 121.75 182.63 0.00 0.00 0.00
Tebuthiuron 1.43 70 47.53 142.60 213.90 0.00 0.00 0.00
2,4-D 17.96 750 598.70 1796.10 2694.15 0.00 0.00 0.00
Diuron 5.54 3 184.67 554.00 831.00 0.06 0.18 0.28
HI 0.59 1.78 2.67
alachlor 1.8 10 60.00 180.00 270.00 0.01 0.02 0.03 237 3 USA Postle et al. (2004)
acetochlor 0.15 20 5.00 15.00 22.50 0.00 0.00 0.00
metolachlor 0.79 97 26.33 79.00 118.50 0.00 0.00 0.00
Dacthal (DCPA) 0.03 100 1.00 3.00 4.50 0.00 0.00 0.00 15 2 USA Eitzer and Chevalier
(1999)
Trifluralin 0.04 7.5 1.33 4.00 6.00 0.00 0.00 0.00
HI 0.01 0.02 0.03
Atrazine 0.009 100 0.30 0.90 1.35 0.00 0.00 0.00 69 5 Luxembourg Bohn et al. (2011)
Desethylatrazine 0.019 100 0.63 1.90 2.85 0.00 0.00 0.00
Deisopropylatrazine 0.003 100 0.10 0.30 0.45 0.00 0.00 0.00
Simazine 0.001 5 0.03 0.10 0.15 0.00 0.00 0.00
2 P,6-Dichlorobenzamide 0.03 100 1.00 3.00 4.50 0.00 0.00 0.00
HI 0.000 0.000 0.000
Ametryn 0.0001 15 0.00 0.01 0.01 0.000 0.000 0.000 NR 4 Canada Woudneh et al. (2009)
Desethylatrazine 0.0001 100 0.00 0.01 0.01 0.000 0.000 0.000
Metribuzin 0.0000 20 0.00 0.00 0.01 0.000 0.000 0.000
Alachlor 0.0003 10 0.01 0.03 0.04 0.000 0.000 0.000
000 000 000
2, 4-D 0.00 750 0.04 0.12 0.17 0.000 0.000 0.000 789 2 China Shi et al., 2020
MCPA 0.00 150 0.04 0.12 0.18 0.000 0.000 0.000
Atrazine 0.02 100 0.70 2.11 3.17 0.000 0.000 0.000 884 1 Wang et al. (2020)
HI 0.000 0.000 0.000
Glyphosate 0.145 100 4.83 14.50 21.75 0.00 0.00 0.00 2 France Botta et al., (2009)
AMPA (Metabolite of 0.464 100 15.47 46.40 69.60 0.00 0.00 0.00
glyphosate)
HI 0.000 0.001 0.001
Atrazine 0.01 100 0.45 1.35 2.03 0.00 0.00 0.00 106 1 Burkina Faso Lehmann et al. (2017)
HI 0.00 0.00 0.00
Butachlor 0.005 100 0.17 0.50 0.75 0.00 0.00 0.00 1 West Bengal Mondal et al. (2018)
0.00 0.00 0.00
2,4-D 2.58 759 86.00 258.00 387.00 0.00 0.00 0.00 45 1 Ethiopia Mekonen et al. (2016)
0.00 0.00 0.001
Trifluralin 0.21 7.5 6.90 20.70 31.05 0.001 0.003 0.004 64 1 Iran Heidar et al. (2017)
0.001 0.003 0.004

Fig. 1b shows the total number of pesticide residues distributed by A includes countries with Rel Aver values of insecticide residues below
country. It appears that the highest number of insecticides was found in 0.01 and includes water samples from five countries (Canada, Portugal,
Lebanon, followed by Spain and West Bengal, whereas the lowest China, South Africa, and Mexico). Section B includes countries with Rel
number was found in Ireland. The highest amount of herbicide residues Aver values in the range of 0.01–0.1 and includes water samples from
was found in Brazil, and the lowest amount was found in Iran. The thirteen countries (Ireland, Japan, Ghana, Nigeria, India, Togo-Africa,
highest amount of fungicide residue was found in Japan, and the lowest Burkina Faso, Vietnam, the USA, the Philippines, Thailand, West Ben­
amount was found in Spain. The differences in the detected number of gal, and Iran). Section C includes countries with Rel Aver values of
pesticide residues among counties may be attributed to different regu­ insecticide residues in the range of 0.11–0.5 and includes water samples
lations and restrictions. from seven countries (Ghana, China, Brazil, Egypt, Co^te d’Ivoire, and
Nicaragua), and Section D includes countries with Rel Aver values of
3.1.7. Concentrations of pesticide residues at the country level insecticide residues >0.5 and includes water samples from six countries
Fig. 2 shows the Rel Aver of pesticide residues in water. Fig. 2a shows (Germany, Lebanon, Ethiopia, Spain, Iran, and India).
the Rel Aver of insecticides. This figure consists of four sections. Section Similarly, Fig. 2b shows the Rel Aver of herbicide residues. This

15
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

a b

Fig. 2. a. Rel Aver of insecticide residues in water from different countries. Rel Aver ranges are <0.01; 0.011–0.099; 0.1–0.5; >0.5 represented by A, B, C, and D
respectively. Countries have the same letter on Rel Aver in a box are not significantly different at p-value >0.05. Error bars represent relative standard error of a
country. Fig. 2b. Relative average (Rel Aver) of herbicide residues in water from different countries. Rel Aver ranges are <0.01; 0.011–0.099; 0.1–0.5; >0.5 rep­
resented by A, B, C, and D respectively. Countries have the same letter on Rel Aver in a box are not significantly different at p-value ≤ 0.05. Error bars represent
relative standard error of a country. Fig. 2c. Relative average (Rel Aver) of fungicide residues in water from different countries. Countries have the same letter on Rel
Aver in a box are not significantly different at p-value ≤ 0.05. Error bars represent relative standard error of a country.

16
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

figure consists of four sections. Section A includes countries with Rel Table 3 shows the name, toxicity class and chemical groups of the
Aver values < 0.01 and includes water samples from five countries pesticides considered in this study. Table 3a shows the insecticides,
(Canada, West Bengal, Luxembourg, Burkina Faso, and Slovenia). Sec­ which included six chemical groups (OC, OP, CT, PY, N, and/or O) and
tion B includes countries with Rel Aver values in the range of 0.01–0.1. four toxicity classes.
This section includes water samples from three countries (Portugal Four and thirteen insecticide residues belonged to toxicity classes Ia
2014, Portugal 2009; Brazil, and Iran). Section C includes countries with and Ib, respectively, and belonged to four chemical classes, OP, CT, OC
Rel Aver values in the range of 0.11–0.5 and includes water samples and PY. Additionally, thirty-one, four and six insecticide residues
from five countries (France, Vietnam, USA, Japan, and Brazil), and belonged to toxicity classes II, III and IV, respectively (Table 3a).
Section D includes countries with Rel Aver values for herbicide residues Furthermore, 16, 10 and 5 herbicide residues belonged to toxicity
of >0.5 and includes water samples from four countries (Lebanon, classes II, III and IV, respectively (Table 3b). However, 1, 8, 3 and 11
Ethiopia, Spain, and Germany). fungicide residues belonged to classes Ib, II, III and IV, respectively
Additionally, Fig. 2c shows the Rel Aver of fungicide residues and (Table 3c).
consists of two sections only. Section A includes countries with Rel Aver
values < 0.01 and includes water samples from Brazil and Japan. Sec­
tions B and C do not include any country, whereas section D includes 3.2. Potential health risks associated with drinking water
countries with Rel Aver values for fungicide residues of >0.5 and in­
cludes water samples from Spain. Applying equations 1–4 to pesticide residues (Tables 1 and 2) pro­
Statistically significant differences among the Rel Aver values in vided the average daily intake HQs and HIs for adults, children, and
different countries are illustrated by letters a, b and c in the vicinity of infants for insecticide, herbicide, and fungicide residues.
the error bars. HIs for adults, children and infants had values ranging from <1 to
30.97 for infants in Egypt. The HI values for infants were the highest
3.1.8. Classification of pesticide residues among all age groups, indicating a high risk. The HI values for infants
Pesticide residues in this review were summarized and classified into were classified into four groups: group 1 included countries with HI
four toxicity classes as follows: toxicity class Ia (extremely toxic com­ values less than 0.01, such as Portugal, Canada, Nigeria, Thailand,
pounds, with LD50 values < 5 μg/g.b.w.); toxicity class Ib (highly toxic China, and Vietnam; group 2 included countries with HI values ranging
compounds, with LD50 values in the range of 5-<50 μg/g.b.w.); toxicity from 0.01 to 0.1, such as Japan, Germany, Ghana, Ireland, China,
class II (moderately toxic compounds, with LD50 values in the range of Mexico, Brazil, Ethiopia, South Africa and the USA; group 3 included
50-<500 μg/g.b.w.); toxicity class III (slightly toxic compounds, with countries with HI values ranging from 0.1 to 1.0, such as Iran, Turkey,
LD50 values in the range of 500-<2000 μg/g.b.w.); and toxicity class IV Togo-Africa, China, Iran, Burkina Faso, West Bengal, Nicaragua, India,
(less toxic compounds, with LD50 values > 2000 μg/g.b.w.). A summary and Ghana; and group 4 included countries with HI values > 1.0, such as
of pesticide toxicity classes is presented in Table 3a, b, c. China, Spain, the Philippines, Cote d’lvoire, Lebanon, Egypt, and India.
Fig. 3 demonstrates the differences between all groups.

Table 3a
Name, toxicity class and chemical group of insecticide residues found in water samples from several countries.
Na ToxC ChemG Na ToxC ChemG Na ToxC ChemG Na ToxC ChemG

Parathion-methyl Ia OP Azinphos Methyl Ib OP Chlorpyrifos II OP Malathion III OP


Parathion-Ethyl Ia OP Isofenphosoxon Ib OP cyanophos II OP Buprofezin III Others
Phorate Ia OP Dicrotophos Ib OP Fenitrothion II OP Methoxychlor III OC
Ethoprophos Ia OP Monocrotophos Ib OP Pyridafenthion II OP etofenprox III PY
Dichlorvos Ib OP Profenofos II OP Thionazin IV OP
Methamidophos Ib OP Dimethoate II OP Chlordimeform IV Others
Omethoate Ib OP Ethion II OP Aldrin IV OC
Triazophos Ib OP Diazinon II OP Endrin IV OC
Chlorfenvinphos Ib OP Phosmet II OP dichlofenthion IV OP
Methiocarb Ib CT Fenthion II OP Mephosfolan IV OP
Carbofuran Ib CT Toxaphene II OC thiamethoxam IV N
Dieldrin Ib OC γ-Chlordane II OC Clothianidin IV N
Cyfluthrin Ib PY γ-HCH II OC
Chlordane II OC
py = 8 DDT II OC
CT = 5 heptachlor II OC
N=4 α-Endosulfan II OC
OP = 27 β-Endosulfan II OC
OC = 14 HCH II OC
OTHERS = 3 Lindane II OC
Total = 61 Fenpropathrin II PY
Deltamethrin II PY
Cypermethrin II PY
L-cyhalothrin II PY
Permethrin II PY
λ-Cyhalothrin II PY
carbaryl II CT
Fenobucarb II CT
Propoxur II CT
Thiacloprid II N
Imidacloprid II N
Fipronil II Others
Total 4 13 32 12

Abbreviations: CT, OP, N, PY, OC, are carbamate, organophosphorus, neonicotinoid, pyrethroids, organochlorine compounds respectively. O= (Obsolete substance).
Na = name, ToxC = toxicity class, ChemG = chemical group.

17
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 3b
Name, toxicity class and chemical group of herbicide residues found in water samples from several countries.
Na ToxC ChemG Na ToxC ChemG Na ToxC ChemG

Alachlor II chloroacetanilide Acetochlor III chloroacetanilide Imazethapyr IV Imidazolinone


Ametryn II triazine Atrazine III triazine Pretilachlor IV Chloroacetamide
2,4-D II Alkylchlorophenoxy Butachlor III Chloroacetamide Propazine IV Triazine
Bentazone II diazinone Diuron III Phenylamide Simazine IV Triazine
Clomazone II Isoxazolidinone Glyphosate III Phosphonoglycine Trifluralin IV Dinitroaniline
Cyanazine II Triazine Linuron III Urea
Hexazinone II Triazinone Metoalachlor III Chloroacetamide
Isoproturon II Urea Prometon III Methoxytriazine
Mecoprop II Aryloxyalkanoic acid Terbutryn III Triazine
Metribuzin II Triazinone Terbuthylazine III Triazine
Molinate II Thiocarbamate
Pendimethalin II Dinitroaniline
Propanil II Anilide
Tebuthiuron II Urea
Terbumeton II Triazine
MCPA II Aryloxyalkanoic acid
Total 16 10 5

The high HI values in group 4 may be explained by the fact that many OP and carbamate pesticides. Arbuckle et al. (2001) also found an as­
OC insecticide residues were detected in the water samples from those sociation between spontaneous abortion in Canadian women and
countries with concentration values exceeding the ARfD. exposure to 2,4-D. Moreover, acetamiprid (Kong et al., 2017), chlor­
The HI index of insecticides increased in the water samples from pyrifos (Brandt et al., 2015), cypermethrin (Saillenfait et al., 2017),
different countries in the following order: Portugal < Canada < Nigeria diazinon, permethrin (Wang et al., 2012), dimethoate (Jallouli et al.,
< Thailand < China < Vietnam < Japan < Germany < Ghana < Ireland 2016), endosulfan (Yan et al., 2019), esfenvalerate, (Saillenfait et al.,
< China < Mexico < Brazil < Mexico < Ethiopia < South Africa < USA 2018), and glyphosate (Romano et al., 2010) interfered with testos­
< Iran < Turkey < Togo-Africa < China < Iran < Burkina Faso < West terone biosynthesis.
Bengal < Nicaragua < India < Ghana < China < Spain < Philippines <
Cote d’Ivoire < Lebanon < Egypt. 3.2.3. Potential carcinogenic agents
The HIs associated with herbicide residues (Table 2) had values in It has been suggested that exposure to pesticide residues may
the range of 0.0001–2.81. According to the classification mentioned enhance carcinogenicity among populations. For instance, Yang et al.
above, group 1 included countries with HI values less than 0.01 (Burkina (2020) indicated that exposure to malathion, terbufos, and chlorpyrifos
Faso, West Bengal, Luxembourg, Ethiopia, Vietnam, France, and Iran), was positively associated with human breast cancer risk. Additionally,
group 2 included countries with HI values between 0.01 and 0.1 (Brazil, Pardo et al. (2020) showed an increased aggressive prostate cancer risk
the USA and Lebanon), group 3 included countries with HI values be­ among ever users of the organodithioate insecticide dimethoate (n = 54
tween 0.11 and 1.0 (no countries), and group 4 included countries with exposed cases, health risk = 1.37, 95% CI = 1.04, 1.80) compared with
HI values > 1.0 (Germany and Spain). never users. However, they observed an inverse association between
The high HI values in group 4 may be attributed to the fact that the aggressive prostate cancer and the herbicide triclopyr (n = 35 exposed
water samples contained high concentrations of herbicides that were cases, health risk = 0.68, 95% CI = 0.48, 0.95), with the strongest in­
close to or exceeded the value of the ARfD. verse association for those reporting durations of use above the median
The HIs associated with fungicide residues in the water samples (≥4 years; n = 13 exposed cases, health risk = 0.44, 95% CI = 0.26,
ranged from 0.001 to 0.23, suggesting that these were low-risk groups 0.77). In contrast, Brasil et al. (2018) indicated that the literature review
(data not shown). According to these data, group 1 included Brazil and does not provide clear evidence for an association between pesticide
Japan, with HIs below 0.01; no water samples were in group 2 (HI values exposure and the occurrence of head and neck cancer. Furthermore,
between 0.01 and 0.1); group 3 included Spain (HI between 0.11 and Hyland et al. (2018) observed that self-reported maternal insecticide use
1.0); and no water samples were in group 4 due to the low fungicide inside the home in the year before pregnancy, during pregnancy, and
residues found. while breastfeeding was associated with increased odds of acute
lymphoblastic leukemia among boys [adjusted odds ratio (aOR) = 1.63
3.2.1. Reproductive toxicity (95% confidence interval [95% CI]: 1.05–2.53), 1.75 (1.13–2.73), and
It has been shown that exposure to pesticide residues may cause 1.75 (1.12–2.73), respectively]. Additionally, Liang et al. (2016) indi­
reproductive toxicity in both men and women (El-Nahhal, 2020). For cated that pesticide exposure was associated with an increased risk of
instance, exposure to OP insecticides caused semen quality deterioration bladder cancer. Similarly, Alavanja and Bonner (2012) found that in­
among Chinese OP pesticide factory workers (Padungtod et al. 2000; secticides, herbicides, fungicides, and fumigants have significant asso­
Recio-Vega et al., 2008). Similar effects were observed among farm ciations with an array of cancer sites. Furthermore, Roberts et al. (2012)
workers occupationally exposed to OP and carbamate pesticides (Mir­ demonstrated the associations between parental use of insecticides and
anda-Contreras et al., 2013). Similarly, deterioration in semen quality acute lymphocytic leukemia and brain tumors. Van Maele-Fabry et al.
has been found among workers exposed to different organochlorine (2006) pointed out that occupational exposure to pesticides is a possible
insecticides (e.g., DDT, β-HCH, γ-HCH) (Jamal et al., 2016). Similarly, risk factor for prostate cancer, but the question of causality remains
exposure to herbicides (e.g., glyphosate) decreased the semen quality unanswered.
parameters of exposed farmers (Sutyarso and Kanedi, 2014). Additionally, Pouchieu et al. (2018) suggested that the risk of Par­
kinson’s disease is increased in farmers exposed to pesticides used on
3.2.2. Potential endocrine disruptors several French crops and livestock, and their research provides addi­
Furthermore, exposure to pesticide residues may create disruption in tional evidence of an association of Parkinson’s disease with dithiocar­
the endocrine system. For instance, Miranda-Contreras et al. (2013) bamate fungicides, rotenone and the herbicides diquat and paraquat.
found alterations in testosterone levels due to occupational exposure to

18
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

3.2.4. Potential cardiovascular toxicity


It has been shown that exposure to organochlorine insecticides may
cause cardiotoxicity in populations. For instance, DDT and its isomers
cause cardiotoxicity in humans (Lamichhane et al., 2019), the devel­

Strobilurin type
Benzisothiazole
Benzimidazole

Chlorophenyl
Chlorophenyl
Chloronitrile
opment of arteriosclerosis and arterial hypertension (Morgan et al.,

Benzanilide
Phenylurea
strobilurin
1980; Donat-Vargas et al. 2018; La Merrill et al. 2013; La Merrill et al.

Oxathiin

Triazole
ChemG 2018; Vafeiadi et al. 2015; Teixeira et al. 2015).
Additionally, organophosphorus insecticide residues may cause
cardiac complications due to direct or indirect exposure. For instance,
occupational exposure to chlorpyrifos increased the risk factor to a value
ToxC

above 1 for acute infarction, whereas exposure to fenitrothion and

11
IV
IV
IV
IV
IV
IV
IV
IV
IV
IV
IV
malathion residues increased blood pressure, leading to cardiovascular
complications among male farmworkers (Zago et al., 2020) and female
tolchlofos methyl oxon farmworkers (Dayton et al., 2010).
Moreover, exposure to herbicides such as alachlor and butachlor
Chlorothalonil

Epoxiconazole

caused hypotension and coma among humans (Lo et al., 2008), whereas
Picoxystrobin
Carbendazim
Azoxystrobin

probenazole
pencycuron
quintozene

atrazine caused heart and lung complications at low concentrations in


flutolanil
mepronil

drinking water (EC 1998; Meghdad. 2013).


Na

3.3. Health hazards of repeated doses


Aromatic hydrocarbon

The influence of repeated consumption of contaminated water on


adults, children and infants is calculated and discussed below.
Benzimidazole

It appears that the HI value increased as the amount of time drinking


Morpholine

water was consumed increased from 0 to 1; 10, 25, 50 and 75 positive


ChemG

associations were obtained in all cases with a regression coefficient


equal to 1. These data indicated that repeated drinking of contaminated
water may result in health risks to populations. However, exposure of
infants and children was worse than that of adults because adults receive
ToxC

less residue with respect to their body weight. These data are provided in
III
III
III

Tables 1 and 2 Furthermore, the highest HI value was found in Germany,


and the lowest HI value was found in Vietnam.
Fenpropimorph

Vietnam and Brazil had HI values < 0.4, and Ghana and Ethiopia had
Thiabendazole
Name, toxicity class and chemical group of fungicide residues found in water samples from several countries.

HI values > 1 and < 3. India, Burkina Faso, and China 2 had HI values >
Etridiazole

5 and < 10, whereas the USA had an HI value of 11.67. West Bengal,
India 2, and China had HI values > 20 and < 50, and Germany had an HI
Na

of 89.57.
These HI values are expected to be two-to threefold higher for chil­
dren and infants. This result suggests that the potential HI for infants is
Triazolobenzothiazole

higher than the HI for children and adults.


Phosphorothiolate
Organophosphate

In conclusion, repeatedly drinking water containing pesticide resi­


dues at concentrations greater than, equal to or a few times lower than
Imidazole

Imidazole
Triazole
Triazole

Triazole

the ARfD may expose the population to potential health risks.


ChemG

Keeping in mind the high values of HI, it is necessary to propose ways


to remove pesticide residues from drinking water to make it suitable for
human consumption.
ToxC

II
II
II
II
II
II
II
II

3.4. Proposed pesticide residue removal options

Selection or design of a suitable method for pesticide removal from


Difenoconazole

Isoprothiolane
Tebuconazole

drinking water requires a full understanding of the physicochemical


tricyclazole
Prochloraz
Iprobenfos
Flutriafol

properties of pesticides, such as their water solubility, chemical struc­


Imazalil

ture, ionic status of molecules, and KOW. The pesticides listed in this
Na

manuscript can be subdivided into seven groups (Table 4) based on their


water solubility. Group 1 included pesticides with a solubility limit
below 1 mg/L and denoted as practically insoluble. This group included
ChemG

23 insecticides, 3 herbicides and 4 fungicides. Group 2 included pesti­


OP

cides with a solubility range of 1–9.99 mg/L and was denoted as poorly
soluble. This group included 6 insecticides, 7 herbicides, and 6 fungicide
residues. Group 3 included pesticides with a solubility range of
ToxC

Ib

10–29.99 mg/L and was denoted as mildly soluble. This group included
1

9 insecticide residues, 7 herbicide residues and 3 fungicides. Group 4


included pesticides with a solubility range of 30–99.99 mg/L and was
Edifenphos

denoted as slightly soluble. This group included 6 insecticide residues, 8


Table 3c

Total

herbicide residues and 6 fungicide residues. Group 5 included pesticides


Na

with a solubility range of 100–999.99 mg/L and was denoted as

19
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Fig. 3. Hazards index (HI) of insecticide calculated for infant from different countries. There are four groups of HI among counties. HI range is shown in each graph.
Values above 1 suggest potential cardiotoxicity among infant. Country marked by * has value 4.5 times higher than presented.

moderately soluble. This group included 6 insecticide residues, 8 her­ 1992; Nir et al., 2000; El-Nahhal and Safi, 2010), activated carbon
bicide residues and 4 fungicides. Group 6 included pesticides with a (Alves et al., 2019; Gamboa-Carballo et al., 2016; Grassi et al., 2012;
solubility range of 1000–9999.99 mg/L and was denoted soluble. This Gupta et al., 2011), carbon nanoparticles (Brooks et al., 2012), and plant
group included 3 insecticide residues, 4 herbicide residues and 0 fungi­ materials (El-Bakouri et al., 2007).
cide residues. Group 7 included pesticides with a solubility range of Organoclays remove several chloroacetamide herbicides (El-Nahhal
>10000 mg/L and was denoted as highly soluble. This group included 6 et al., 1998; Nir et al., 2000). Porous materials removed simazine from
insecticide residues, 6 herbicide residues and 0 fungicide residues. water (Esposito et al., 2016); activated carbon removed carbaryl,
Furthermore, these pesticides are organic molecules classified into methomyl and carbofuran (Alves et al., 2019); granular activated carbon
different chemical groups. and carbon nanotubes removed atrazine, simazine and diuron (Alves
et al., 2020); carbon slurry removed endosulfan and methoxychlor
3.4.1. Removal by physical processes using adsorption (Gupta and Ali, 2008); and activated carbon removed methoxychlor,
Adsorption is denoted here as the adhesion of pesticide ions or methyl parathion and atrazine (Gupta et al., 2011). The authors
molecules from water to the surfaces of solid materials. This may create emphasized the importance of adsorptive materials used, specific su­
an accumulation of pesticide molecules, forming a monolayer or bi­ perficial area, and size and distribution of specific pore volume. Bio­
layers, depending on the physicochemical properties of pesticides and adsorbents removed atrazine, alachlor, endosulfan sulfate and trifluralin
solid materials (El-Nahhal et al., 1998). from water (Rojas et al., 2014). Plant materials removed 4 herbicide
The removal of pesticide residues is affected by physicochemical residues, alachlor, atrazine, simazine and trifluralin (Sharma et al.,
properties such as hydrophobicity, water solubility, molecular weight, 2008), and 8 insecticide residues, aldrin, chlorpyrifos, chlorfenvinphos,
evaporation and sensitivity to photodegradation (Rodriguez et al., 2016; dieldrin, α-endosulfan, endrin, hexachlorobenzene, β-HCH, γ-HCH,
Gonzalez et al., 2013). (El-Bakouri et al., 2007), carbaryl and fenuron (Loffredo and Taskin,
Accordingly, we propose an adsorption method to remove highly 2017).
hydrophobic residues (groups 1–3, Table 4) and use organic adsorbing In contrast, atrazine (Yang et al., 2018) and λ-cyhalothrin (An et al.,
materials such as activated carbon, activated charcoal, nanoorganic 2020) were removed using colloidal biochar. Similarly, carbendazim
materials, organoclays, and bioadsorbents such as plant residues. For and linuron were removed from aqueous solutions with activated carbon
instance, highly hydrophobic pesticide residues (groups 1–3, Table 4) produced from spent coffee grounds (Hgeig et al., 2019).
were removed using modified clay (Boyd et al., 1988; Boyd and Jaynes, In summary, adsorption is a suitable, easily applicable, cost effective

20
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

and efficient method for removing OC insecticides (Gamboa-Carballo The mechanism of generating hydroxyl radicals in each step has
et al., 2016), herbicides (Brooks et al., 2012) and fungicides (El-Nahhal previously been described in detail (Munter, 2001).
and Safi, 2008). AOPs have successfully removed OC residues (e.g., DDT, DDE, DDD),
OP residues (e.g., chlorpyrifos), chloroacetamide herbicides (e.g., ala­
3.4.2. Removal by physical processes using membrane filtration chlor, metolachlor), triazines (e.g., atrazine) and dinitroaniline (e.g.,
The membrane filtration process is a physical separation method pendimethalin) from water (Sayles et al., 1997; Pirnie et al., 2006;
used to separate molecules of different sizes and characteristics. Con­ Thomas et al., 2008; Zhu et al., 2016; Marican and Durán-Lara, 2018).
centration differences between the two sides of a special membrane may Similarly, they increased the degradation of triazine (e.g., atrazine) and
control the separation. Separation by membranes may depend on the chloroacetamide (e.g., metolachlor) herbicides due to the addition of
sieving power and the physiochemical properties of the membrane. To zerovalent iron to soil containing residues (Marican and Durán-Lara,
date, membrane filtration processes (e.g., nanofiltration and ultrafil­ 2018). Moreover, AOPs removed 100% of ametryn (Sangami and Manu,
tration) have been used for natural organic matter and pesticide 2017), 98% of malathion and parathion-methyl (Roe and Lemley, 1997),
removal. The general separation trend of these membranes is based on and caused rapid degradation of alachlor, chlorpyrifos, and pendime­
hydrophobic adsorption as a function of the octanol–water partition thalin due to the addition of aluminum sulfate and acetic acid with
coefficient between the hydrophobic compounds and porous hydro­ zerovalent iron (Shea et al., 2004). Nurul Amin et al. (2008) completely
phobic membrane during membrane filtration (Yoon et al., 2007). removed thiobencarb from water using a powdered zero valent iron and
Therefore, this method may be suitable for pesticide residues in groups batch experimental system. Yu (2002) revealed efficient and fast
1–3 (Table 4). However, a single method usually does not achieve 100% removal of OP residues from water (fenitrothion, chlorpyrifos, diazinon,
removal of pesticides. Accordingly, it is suggested to use membrane methamidophos, edifenphos, mevinphos, fenthion, and acephate) by
filtration in tandem with adsorption using activated carbon to further supercritical carbon dioxide extraction followed by Fenton’s solution
enhance the removal efficiency of pesticide residues. A combination of treatment. da Costa et al. (2019) showed the ability of the photo-Fenton
methods (activated carbon + and membrane filtration) has successfully reaction to degrade the fungicide carbendazim within 15 min with 96%
removed pharmaceuticals and pesticide residues from water (Acero removal, indicating that both solar photo-Fenton and artificially irra­
et al., 2012; Margot et al., 2013; Rodriguez et al., 2016; Snyder et al., diated systems are promising routes for carbendazim degradation. Erol
2007). However, Alves et al. (2020) used a microfiltration technique et al. (2019) removed DDT and its analogs very quickly from water by
combined with granular activated carbon fixed bed adsorption tech­ magnetic diatomite. However, Kamel et al. (2009) showed the efficiency
nology to remove large amounts (100%) of carbaryl and carbofuran of chlorine as a strong oxidizing reagent in completely oxidizing ten OP
from public water supplies. insecticides (phorate, disulfoton, terbufos, methidathion, bensulide,
chlorethoxyfos, phosmet, parathion-methyl, phostebupirim, and teme­
3.4.3. Removal by advanced oxidation processes phos) in water samples within 72 h under laboratory conditions.
Advanced oxidation processes (AOPs) are based on initiating chem­
ical reactions that produce hydroxyl radicals (⋅OH) to partially or 3.4.4. Removal of pesticide residues via combined processes
completely degrade pesticide residues to nontoxic fragments. The for­ It has been shown that using a combination of two advanced
mation of hydroxide radicals was previously reported (Fenton, 1894). oxidation systems (ZnO(normal)/H2O2/artificial sunlight and ZnO
Recent works emphasized the importance of peroxide reactions in the (nano)/H2O2/artificial sunlight) successfully removed imidacloprid
removal of organic and inorganic pollutants from water (Buxton et al., from water (pubmed.ncbi.nlm.nih.gov/?term=Derbalah+A&cauthor
1988). Later, several attempts were made to enhance the production of _id=30942775Derbalah et al., 2019).
hydroxyl radicals (⋅OH), including chemical and photochemical at­ Including electro-oxidation-ozone treatment removed all imidaclo­
tempts (Munter, 2001; Stefan, 2017). AOPs may be suitable for prid within 120 min and increased mineralization to 50% TOC (Segura
removing pesticide residues with different chemical groups in their et al., 2008; Malato et al., 2002; González et al., 2020). Using
chemical structure with different electronegativities, such as chlorinated electro-Fenton processes further improved the degradation rate of imi­
hydrocarbon (OC insecticides and chloroacetamide herbicide), organo­ dacloprid (Iglesias et al., 2014) and carbofuran (Abdessalem et al., 2010;
phosphates, carbamates, pyrethroids, and dinitroaniline derivatives. Ma et al., 2009). Similarly, 94% of imidacloprid (pubmed.ncbi.nlm.nih.
Hydroxyl radicals are highly reactive and nonselective oxidizing gov/?term=Turabik+M&cauthor_id=24671401Turabik et al., 2014),
agents for decomposing organic pollutants in water systems (Glaze et al., 100% of λ-cyhalothrin (pubmed.ncbi.nlm.nih.gov/?term=Colombo
1987). Accordingly, pesticide residues such as 2,4-D, 2,4,5-T, chlor­ +R&cauthor_id=23380363Colombo et al., 2013) and 99% of atrazine
ophenols, paraquat, diquat, glyphosate, parathion, malathion, pyrethrin were removed from water within 2 h by electrochemical advanced
and carbaryl may be degraded or mineralized. This is in accordance with oxidation processes and electro-Fenton processes with a boron-doped
the findings reported by Munter (2001), who revealed fast decomposi­ diamond anode (pubmed.ncbi.nlm.nih.gov/?term=Komtchou+S&
tion of organic molecules (e.g., acids, alcohols, aldehydes, aromatics, cauthor_id=28837868Komtchou et al., 2017).
amines, ethers, ketones). Advanced oxidation treatment resulted in Barbosa et al. (2016) showed the possibility of removing imidaclo­
detoxification of many chlorinated hydrocarbons and pesticides prid, methiocarb, and oxadiazon using a UV filter (2-ethyl­
(Andreozzi et al., 1999). hexyl-4-methoxycinnamate) and an antioxidant (2,6-di-tert-butyl
The following reactions are used to produce *OH radicals: -4-methylphenol).
Deltamethrin (Lafi and Al-Qodah, 2006) and chlorothalonil (Kiefer
1. Chemical reactions et al., 2020) were completely removed within 210 min using O3 and
1.1 Ozonation at alkaline pH (>8.5) O3/UV oxidation systems followed by biological treatments.
1.2 O3 + H2O2 Moreover, using a combination of oxidation methods (UV/H2O2,
1.3 O3 + catalyst UV/S2O82− , and UV/H2O2/S2O82− ) further completed the photo­
1.4 Fenton system (H2O2/Fe2+) degradation of carbofuran (Chu et al., 2006), diazinon (Rasoulifard
2. Photochemical reactions et al., 2015), metolachlor (Wu et al., 2007), and endosulfan (Shah et al.,
2.1 O3/UV 2013) in water systems.
2.2 H2O2/UV Simazine was removed from water by various techniques, with a
2.3 O3/H2O2/UV removal rate following this order: ozone < ozone/hydrogen peroxide <
2.4 Photo-Fenton/Fenton-like systems UV radiation < ozone/UV radiation and UV radiation/hydrogen
2.5 Photocatalytic oxidation (UV/TiO2) peroxide (Beltrán et al., 2000).

21
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Moreover, 99% of atrazine was removed from water within 15–45 et al., 2016). In contrast, Garcinuño et al. (2006) removed 57% of
min using Fenton reactions (photoelectro-Fenton, electro-Fenton and carbaryl, 53% of linuron and 55% of permethrin using aquatic plants
anodic oxidation with simultaneous H2O2 formation (AO - H2O2)). (Upinus angustifolius), indicating that some plants (e.g., L. angustifolius)
Approximately 98% of chlorothalonil, parathion-methyl and meth­ could be a useful biotechnology for pesticide removal from water.
amidophos were removed from water using the Fenton reaction within Similarly, Su et al. (2019) showed the ability of genetically modified
40 min (Gutiérrez et al., 2007), whereas 75% of alachlor and diuron plants to remove acetochlor from water, with a percentage removal that
were removed within 30 min after treatment using a falling film reached 51.7–57.4%. However, Hwang et al. (2020) completely
dielectric barrier discharge plasma reactor combined with adsorption on removed atrazine residues in 12 weeks from water by floating treatment
activated carbon (Vanraes et al., 2018). in wetlands planted with Canna flaccida.
Inclusion of an aerobic reactor along with the Photo-Fenton tech­ However, organochlorine, strobilurin/strobin, organosphosphate
nique yielded acceptable degradation rates of chlorothalonil, chlorpyr­ and pyrethroid residues were removed from agricultural wastewater
ifos, and cypermethrin, improving water quality up to local standards using constructed wetlands (Vymazal and Březinová, 2015). To date,
(Affam et al., 2014). pesticide removal generally increases in direct proportion with KOC;
High percentages of 2,4-D, glyphosate, trifluralin, and butachlor however, the relationship is not strong. Similarly, many pesticides of
were removed using chemical oxidation (graphene oxide/TiO) followed different chemical structures (e.g., fluoranthene, trifluralin, alachlor,
by adsorption (Hosseini and Toosi, 2019). Similarly, 83.4% of trifluralin isoproturon, diuron, tributyltin compounds, simazine, atrazine, chlor­
was removed by ozonation after 20 min (Milhome et al., 2018), and 2, pyrifos, chlorfenvinphos, hexachlorobenzene, pentachlorophenol,
4-D residues were removed using bioadsorbents (Brito et al., 2020). endosulfan, DDT and dieldrin) were removed at high weekly removal
Chlorothalonil residue was removed from different materials after 5 efficiencies (>87%) (Gorito et al., 2017, 2018). Moreover, a high
min of ozonation (Kusvuran et al., 2012) and from water by catalytic removal rate (87.22 ± 16.61%) of OP residues was found using con­
degradation using bimetallic iron (Ghauch and Tuqan, 2008). Similarly, structed wetlands (Liu et al., 2019). Similarly, Sherrard et al. (2004)
> 92% of both dieldrin and DDT were removed by ozonation followed showed the efficiency of constructed wetland in removing chlorothalonil
by microfiltration (Kenny et al., 2018). Approximately 95% of fenpro­ and chlorpyrifos from water. Similar observations were obtained with
pathrin and λ-cyhalothrin were removed by UV light and simulated simazine and wetlands (Wilson et al., 2011). However, Xia and Ma
sunlight (Liu et al., 2014), whereas 90% of λ-cyhalothrin and cyper­ (2006) emphasized the phytoremediation potential of water hyacinth
methrin were removed by photodegradation and chemical oxidation (Eichhornia crassest) to remove ethion from water. In summary, this
using copper (Xie et al., 2011). However, Yabalak and Gizir (2020) method is suitable for pesticide residues in groups 1–4 (Table 4) because
demonstrated the ability of oxidation methods to remove pesticide res­ their water solubility enables them to stay in the root zone, hence
idues with different chemical structures (e.g., chlorpyrifos, penconazole, exposing them to potential microbial degradation associated with plant
brassilexin, buprofezin, etoxazole, and pyriproxyfen). Wardenier et al. roots.
(2019) removed atrazine, alachlor, diuron, dichlorvos and pentachlo­
rophenol using an innovative AOP based on electrical discharges in a 3.4.6. Removal of pesticide residues via biodegradation
continuous-flow pulsed dielectric barrier discharge reactor. Addition­ Pesticide biodegradation is a partial or total destruction of pesticide
ally, Yang and Zhang (2019) revealed rapid degradation and elimination molecules by bacteria or fungi throughout three stages. This includes
of dimethoate, atrazine and propoxur due to exposure to vacuum biodeterioration (partial metabolism that may introduce minor changes
ultraviolet/ultraviolet/chlorine. Westlund et al. (2018) eliminated to the chemical structure of pesticide molecules), biofragmentation (a
50–70% of the biological activity of atrazine, 2,4-D, terbutryn, diuron, complete conversion of pesticide molecules to carbon dioxide, water
and tebuconazole by ozonation. Furthermore, Aliste et al. (2020) and/or other gases), and assimilation (the conversion of carbon dioxide
removed seven fungicide residues, including difenoconazole, from and water by plants to glucose). A similar definition was given for other
leaching water using different homogeneous (photo-Fenton and cases (Lucas et al., 2008). Several pesticides have been removed from
photo-Fenton-like) and heterogeneous (ZnO and TiO2) photocatalytic water by biodegradation. For instance, chlorpyrifos and diazinon (OP
systems. They indicated that mineralization of initial dissolved organic residues) were removed from liquid medium, soil and a biobed bio­
carbon was not complete, but the conversion rate under ZnO/Na2S2O8 mixture using a Streptomyces mixed culture with removal rates of 0.036
treatment was approximately 1.3 times higher than that using and 0.015/h and half-lives of 19 and 46/h, respectively (Briceño et al.,
TiO2/Na2S2O8. 2018). Similarly, phorate (OP residue) was removed from water using a
Furthermore, advanced chemical method application (chlorination, biodegradation method consisting of three microorganisms (Brevibacte­
ozonation, chemical precipitation) followed by adsorption using acti­ rium frigoritolerans, Bacillus aerophilus and Pseudomonas fulva), and the
vated carbon successfully removed 44 pesticide residues from water, biodegradation efficiency ranged from 97.65 to 98.31% (Jariyal et al.,
with removal percentages that reached 90%. The tested pesticides are 2018).
herbicides from different groups, and insecticides from OC and OP- In contrast, carbaryl was completely (100%) removed from water
residues are listed in Table 1 (Ormad et al., 2008). In contrast, Khedr using Rhodopseudomonas sphaeroides (Wu et al., 2019a, 2019b), DDT was
et al. (2019) showed the ability of gamma irradiation to completely removed using Stenotrophomonas sp (Pan et al., 2016), and lindane was
eliminate chlorfenvinphos, diazinon, dimethoate, and profenofos from removed using Gram-negative bacteria (Zhang et al., 2020). Alvar­
deionized and ionized water at low pesticide concentrations. It can be ado-Gutiérrez et al. (2020) showed the ability of three bacteria (Klebsi­
concluded that AOPs are suitable and effective methods to remove ella oxytoca, Flavobacterium johnsoniae, and Stenotrophomonas
pesticide residues in groups 1–7 (Table 4). maltophilia) to biodegrade carbendazim. Additionally, Wu et al. (2018)
showed the ability of Sphingomonas sp. to remove highly toxic and
3.4.5. Removal of pesticide residues via phytoremediation recalcitrant tricyclazole from wastewater.
The use of living plants to remove pesticide residues from water is
denoted as phytoremediation, although the phytoremediation definition 3.5. Comparison between this study and studies in the literature
includes using living plants to clean up soil, air and water contaminated
with hazardous contaminants (Reichenauer and Germida, 2008). Phy­ A comparison of the present study with previously published studies
toremediation has been successfully used to remove clomazone residues on this subject matter is shown in Table 5. The unique aspect of the
in water (Escoto et al., 2019), 56% for acetochlor and 51% for alachlor, present review is that it collected concentrations of pesticide residues
whereas metolachlor was resistant to removal with an efficiency of 23% from many countries and then classified these pesticide residues into
(Elsayed et al., 2014), and similarly for imazalil and tebuconazole (Lv different functional and toxicity groups. It also estimated the ODI values

22
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Table 4
Classes of some pesticide according to solubility limit in water.
Solubility class Solubility range Insecticides Herbicides Fungicides
(mg/L)

Practically <1 PY (L-Cyhalothrin, λ-Cyhalothrin, Cyfluthrin, Dinitroaniline (Trifluralin, Phenylurea (pencycuron)


insoluble Cypermethrin, Deltamethrin, Etofenprox, Permethrin, Pendimethalin) Chlorophenyl (quintozene,
Fenpropathrin) Oxidiazole (oxadiazon) Chloronitrile
OC (Aldrin, heptachlor, α-/β-/γ-Chlordane, Methoxychlor, (Chlorothalonil)
Dieldrin, Endrin, α-/β-Endosulfan)
O (Buprofezin, Pyriproxyphen,
OP (dichlofenthion)
Poorly soluble 1–9.99 OC (Lindane) Amide (bromobutide) Methoxyacrylate
OP (Chlorpyrifos, Ethion, Fenthion) Oxyacetamide (mefenacet) (Picoxystrobin)
CT (carbaryl) Thiocarbamate (esprocarb) Morpholine (Fenpropimorph)
O (Fipronil) OP (butamifos) Strobilurin (Azoxystrobin)
Triazine (Simazine, Terbuthylazine, Triazole (Epoxiconazole)
Propazine) Benzimidazole
(Carbendazim)
Oxathiin (flutolanil)
Low soluble 10–29.99 OP (Azinphos-Methyl, Disulfoton, Fenitrothion, isofenphos Chloroacetamide (alachlor, Butachlor) Benzanilide (mepronil)
oxon, Parathion-Ethyl, Phosmet, Profenofos) Thiocarbamate (thiobencarb, Triazole (Difenoconazole)
CT (Methiocarb) dimepiperate) Imidazole (Prochloraz)
Benzonitrile (dichlobenil)
OP (piperophos)
Triazine (Terbutryn)
Slightly soluble 30–99.99 OP (Triazophos, cyanophos, Phorate, Parathion-methyl, Triazine (prometryn, Atrazine) Benzimidazole
Mephosfolan, Diazinon) Phenylamide (Diuron) (Thiabendazole)
Methylthiotriazine (dimethametryn) Triazole (Tebuconazole,
Urea (Linuron, Isoproturon, Flutriafol)
Chlorotoluron) OP(edifenphos)
Anilide (Propanil) O (Etridiazole)
Moderately 100–999.99 OP (pyridaphenthion, Chlorfenvinphos, malathion) Triazine (Terbumeton, Cyanazine, Benzisothiazole
soluble CT (Carbofuran, Fenobucarb) Ametryn, simetryn, Prometon) (probenazole)
N (Imidacloprid) Chloroacetanilide (Acetochlor, Alachlor, Imidazole (Imazalil)
Pretilachlor, Metoalachlor) OP (iprobenfos)
Triazolobenzothiazole
(tricyclazole)
soluble 1000–9999.99 OP (Thionazin, Ethoprophos) Thiocarbamate (Molinate)
CT (Propoxur) Isoxazolidinone (Clomazone)
Imidazolinone (Imazethapyr)
Urea (Tebuthiuron)
highly soluble >10000 OP (Dichlorvos, Dimethoate, Methamidophos, Omethoate, Phosphonoglycine (Glyphosate)
Monocrotophos, Dicrotophos) Triazinone (Metribuzin, Hexazinone)
Alkylchlorophenoxy (2,4-D, MCPA,
Mecoprop)

of pesticide residues and compared them with ARfD values; calculated health risk among populations.
HQ and HI for adults, children and infants in each country; predicted In conclusion, the present review highlights the current status of
future toxicity; classified countries into four groups; and discussed pesticide residues in countries and provides suitable options for their
human health risk. The present review provides removal options. removal. Thus, it fills the research gap in the field of pesticide residues in
In contrast, previously published studies neither classified pesticide drinking water and their purification options that can reduce the po­
residues into toxic classes nor provided removal options for these resi­ tential health risk and provide valuable information that can serve as a
dues. Additionally, the published articles did not review the potential baseline for further developments in the field.

Table 5
Comparison between this review and those published previously.
Reviewed parameter Morrissey et al. Mohammed et al. Elibariki and Bruce-Vanderpuije et al. Li and Jennings Olisah et al. Present
(2015) (2019) Maguta (2017) (2019) (2017) (2020) review

Countries Many countries Ghana Tanzania Ghana Few countries Few countries, Many
countries
Insecticide residues ✓ ✓ ✓ ✓ ✓ X ✓
Acaricide residues X X X X X X ✓
Fungicide residues X X X X ✓ X ✓
Herbicide residues X X ✓ X ✓ X ✓
Classification of residues X X X X X X ✓
Estimation of ADI X X X X X X ✓
Estimation of MDI X X X X X X ✓
Predicted toxics scenario X X X X X ✓
HI for adult, X X X X X X ✓
HI for child X X X X X X ✓
HI for infant X X X X X X ✓
Removal options X X X X X X ✓
guidelines for a safe water X X X X X X ✓
protection

23
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

4. Conclusion Alvarado-Gutiérrez, M.L., Ruiz-Ordaz, N., Galíndez-Mayer, J., Curiel-Quesada, E.,


Santoyo-Tepole, F., 2020. Degradation kinetics of carbendazim by Klebsiella
oxytoca, Flavobacterium johnsoniae, and Stenotrophomonas maltophilia strains.
This review highlighted the levels of pesticide residues in drinking Environ. Sci. Pollut. Res. Int. 27 (23), 28518–28526. https://doi.org/10.1007/
water samples in several countries and demonstrated the presence of s11356-019-07069-8.
hazardous residues of pesticides with class Ia, Ib and II toxicity. More­ Alvarenga, N., Birolli, W.G., Seleghim, M.H., Porto, A.L., 2014. Biodegradation of methyl
parathion by whole cells of marine-derived fungi Aspergillus sydowii and
over, pesticides with class Ia or Ib toxicity are extremely toxic pesticides, Penicillium decaturense. Chemosphere 117, 47–52. https://doi.org/10.1016/j.
and most of them are cholinesterase inhibitors. chemosphere.2014.05.069.
This review also estimated the average ODI of pesticides in many Alves Pimenta, J.A., Francisco Fukumoto, A.A., Madeira, T.B., Alvarez Mendez, M.O.,
Nixdorf, S.L., Cava, C.E., Kuroda, E.K., 2020. Adsorbent selection for pesticides
countries, calculated the corresponding HI and demonstrated high removal from drinking water. Environ. Technol. 1–12. https://doi.org/10.1080/
values in many countries (4.48, 28.74, 30.97), suggesting potential risk 09593330.2020.1847203. Advance online publication.
among consumers. Moreover, this review demonstrated health risks Alves, A., Ruiz, G., Nonato, T., Müller, L.C., Sens, M.L., 2019. Performance of the fixed-
bed of granular activated carbon for the removal of pesticides from water supply.
among water consumers. The review highlights the value of the Rel Std Environ. Technol. 40 (15), 1977–1987. https://doi.org/10.1080/
Error as a statistical tool with which to compare the concentration of 09593330.2018.1435731.
pesticide residues between countries. Furthermore, this review presents Alves, A., Ruiz, G., Nonato, T., Pelissari, C., Dervanoski, A., Sens, M.L., 2020. Combined
microfiltration and adsorption process applied to public water supply treatment:
pesticide removal options. water quality influence on pesticides removal. Environ. Technol. 41 (18),
Nevertheless, the limitation of this study is that there is a lack of 2382–2392. https://doi.org/10.1080/09593330.2019.1567605.
published data regarding pesticide residues, and only old literature can An, Q., Li, D., Wu, Y., Pan, C., 2020. Deposition and distribution of myclobutanil and
tebuconazole in a semidwarf apple orchard by hand-held gun and air-assisted
be found in some cases.
sprayer application. Pest Manag. Sci. 76 (12), 4123–4130. https://doi.org/10.1002/
Nonetheless, the main point of the study is that the number of ps.5968.
countries included is a representative sample of all countries, as well as Andreozzi, R., Caprio, V., Insola, A., Marotta, R., 1999. Advanced oxidation processes
the fact that many pesticide residues were similarly found in some (AOP) for water purification and recovery. Catal. Today 53, 51_59. https://doi.org/
10.1016/S0920-5861(99)00102-9. Available from:
countries. Furthermore, this review also provides a systematic reference Ankley, G., Collyard, S., 1995. Influence of piperonyl butoxide on the toxicity of
with which to estimate the daily intake of pesticide residues, allowing organophosphate insecticides to three species of freshwater benthic invertebrates.
researchers to determine HIs that are best suited for the study. In Comp. Biochem. Physiol. C Pharmacol. Toxicol. Endocrinol. 110 (2), 149–155.
https://doi.org/10.1016/0742-8413(94)00098-U.
conclusion, this review updates the state of the art in the field of pesti­ Bajet, C.M., Navarro, M.P., 2002. Degradation, release and bioavailability of 14C DDT
cide residues in drinking water and their potential health risks. This and 14C DDE sediment residues to oysters and mussels. Environ. Technol. 23 (11),
manuscript may help young scholars in the field of pesticide residues in 1293–1302. https://doi.org/10.1080/09593332308618318.
Bakore, N., John, P.J., Bhatnagar, P., 2004. Organochlorine pesticide residues in wheat
water. and drinking water samples from Jaipur, Rajasthan, India. Environ. Monit. Assess.
98 (1–3), 381–389.
Bao, Z., Zhao, Y., Wu, A., Lou, Z., Lu, H., Yu, Q., Fu, Z., Jin, Y., 2020. Sub-chronic
Declaration of competing interest carbendazim exposure induces hepatic glycolipid metabolism disorder accompanied
by gut microbiota dysbiosis in adult zebrafish (Daino rerio). Sci. Total Environ. 739,
140081. https://doi.org/10.1016/j.scitotenv.2020.140081.
No competing interest to declare. Barbosa, M.O., Moreira, N., Ribeiro, A.R., Pereira, M., Silva, A., 2016. Occurrence and
removal of organic micropollutants: an overview of the watch list of EU Decision
2015/495. Water Res. 94, 257–279. https://doi.org/10.1016/j.watres.2016.02.047.
Acknowledgment
Beltrán, F.J., García-Araya, J.F., Rivas, J., Alvarez, P.M., Rodríguez, E., 2000. Kinetics of
simazine advanced oxidation in water. J. Environ. Sci. Health - Part B Pesticides,
Prof Dr Yasser El-Nahhal thanks the AvH Foundation for funding the Food Contam. Agric. Wastes 35 (4), 439–454. https://doi.org/10.1080/
03601230009373281.
research stay in Berlin, Germany.
Bento, C., van der Hoeven, S., Yang, X., Riksen, M., Mol, H., Ritsema, C.J., Geissen, V.,
2019. Dynamics of glyphosate and AMPA in the soil surface layer of glyphosate-
References resistant crop cultivations in the loess Pampas of Argentina. Environ. Pollut. 244,
323–331. https://doi.org/10.1016/j.envpol.2018.10.046.
Bhadouria, B.S., Mathur, V.B., Kaul, R., 2012. Monitoring of organochlorine pesticides in
Abdessalem, A.K., Bellakhal, N., Oturan, N., Dachraoui, M., Oturan, M.A., 2010.
and around Keoladeo national Park, bharatpur, Rajasthan, India. Environ. Monit.
Treatment of a mixture of three pesticides by photo-and electro-Fenton processes.
Assess. 184 (9), 5295–5300. https://doi.org/10.1007/s10661-011-2340-z.
Desalination 250 (1), 450–455. https://doi.org/10.1016/j.desal.2009.09.072.
Bhandari, G., Atreya, K., Scheepers, P., Geissen, V., 2020. Concentration and distribution
Acero, J.L., Javier Benitez, F., Real, F.J., Teva, F., 2012. Coupling of adsorption,
of pesticide residues in soil: non-dietary human health risk assessment. Chemosphere
coagulation, and ultrafiltration processes for the removal of emerging contaminants
253, 126594. https://doi.org/10.1016/j.chemosphere.2020.126594.
in a secondary effluent. Chem. Eng. J. 210, 1_8. https://doi.org/10.1016/j.
Bidleman, T.F., Leone, A.D., Falconer, R.L., Harner, T., Jantunen, L.M., Wiberg, K.,
cej.2012.08.043. Available from:
Helm, P.A., Diamond, M.L., Loo, B., 2002. Chiral pesticides in soil and water and
Affam, A.C., Chaudhuri, M., Kutty, S.R.M., Muda, K., 2014. UV Fenton and sequencing
exchange with the atmosphere. Sci. World J. 2, 357–373. https://doi.org/10.1100/
batch reactor treatment of chlorpyrifos, cypermethrin and chlorothalonil pesticide
tsw.2002.109.
wastewater. Int. Biodeterior. Biodegrad. 93, 195–201. https://doi.org/10.1016/j.
Birolli, W.G., Alvarenga, N., Seleghim, M.H., Porto, A.L., 2016. Biodegradation of the
ibiod.2014.06.002.
pyrethroid pesticide esfenvalerate by marine-derived fungi. Mar. Biotechnol. 18 (4),
Affum, A.O., Acquaah, S.O., Osae, S.D., Kwaansa-Ansah, E.E., 2018. Distribution and risk
511–520. https://doi.org/10.1007/s10126-016-9710-z.
assessment of banned and other current-use pesticides in surface and groundwaters
Bohn, T., Cocco, E., Gourdol, L., Guignard, C., Hoffmann, L., 2011. Determination of
consumed in an agricultural catchment dominated by cocoa crops in the Ankobra
atrazine and degradation products in Luxembourgish drinking water: origin and fate
Basin, Ghana. Sci. Total Environ. 633, 630–640. https://doi.org/10.1016/j.
of potential endocrine-disrupting pesticides. Food Addit. Contam. Part A, Chemistry,
scitotenv.2018.03.129.
analysis, control, exposure & risk assessment 28 (8), 1041–1054. https://doi.org/
Agarwal, A., Prajapati, R., Singh, O.P., Raza, S.K., Thakur, L.K., 2015. Pesticide residue in
10.1080/19440049.2011.580012.
water–a challenging task in India. Environ. Monit. Assess. 187 (2), 54. https://doi.
Bommuraj, V., Chen, Y., Klein, H., Sperling, R., Barel, S., Shimshoni, J.A., 2019. Pesticide
org/10.1007/s10661-015-4287-y.
and trace element residues in honey and beeswax combs from Israel in association
Alavanja, M.C., Bonner, M.R., 2012. Occupational pesticide exposures and cancer risk: a
with human risk assessment and honey adulteration. Food Chem. 299, 125123.
review. J. Toxicol. Environ. Health B Crit. Rev. 15 (4), 238–263. https://doi.org/
Botwe, B.O., Kelderman, P., Nyarko, E., Lens, P., 2017. Assessment of DDT, HCH and
10.1080/10937404.2012.632358.
PAH contamination and associated ecotoxicological risks in surface sediments of
Albuquerque, A.F., Ribeiro, J.S., Kummrow, F., Nogueira, A.J., Montagner, C.C.,
coastal Tema Harbour (Ghana). Mar. Pollut. Bull. 115 (1–2), 480–488. https://doi.
Umbuzeiro, G.A., 2016. Pesticides in Brazilian freshwaters: a critical review.
org/10.1016/j.marpolbul.2016.11.054.
Environmental science. Processes & impacts 18 (7), 779–787. https://doi.org/
Bouhala, A., Lahmar, H., Benamira, M., Moussi, A., Trari, M., 2020. Photodegradation of
10.1039/c6em00268d.
organophosphorus pesticides in honey medium by solar light irradiation. Bull.
Alegria, H., Bidleman, T.F., Figueroa, M.S., 2006. Organochlorine pesticides in the
Environ. Contam. Toxicol. 104 (6), 792–798. https://doi.org/10.1007/s00128-020-
ambient air of Chiapas, Mexico. Environ. Pollut. 140 (3), 483–491. https://doi.org/
02858-1.
10.1016/j.envpol.2005.08.007.
Boyd, S.A., Jaynes, W.F., 1992. Role of layer charge in organic contaminant sorption by
Aliste, M., Pérez-Lucas, G., Vela, N., Garrido, I., Fenoll, J., Navarro, S., 2020. Solar-driven
organo clays. In: Proceedings of the CMS Workshop on Layer Charge Characteristics
photocatalytic treatment as sustainable strategy to remove pesticide residues from
of Clays. Canada, University of Saskatchewan, Minneapolis, MN, pp. 89–120.
leaching water. Environ. Sci. Pollut. Res. Int. 27 (7), 7222–7233. https://doi.org/
10.1007/s11356-019-07061-2.

24
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Boyd, S.A., Shaobai, S., Lee, J.F., Mortland, M.M., 1988. Pentachlorophenol sorption by 1999-2015. Environ. Health : a global access science source 18 (1), 7. https://doi.
organo-clays. Clay Clay Miner. 36, 125–130. org/10.1186/s12940-018-0441-7.
Brasil, V., Ramos Pinto, M.B., Bonan, R.F., Kowalski, L.P., da Cruz Perez, D.E., 2018. Cryder, Z., Greenberg, L., Richards, J., Wolf, D., Luo, Y., Gan, J., 2019. Fiproles in urban
Pesticides as risk factors for head and neck cancer: a review. J. Oral Pathol. Med. : surface runoff: understanding sources and causes of contamination. Environ. Pollut.
official publication of the International Association of Oral Pathologists and the 250, 754–761. https://doi.org/10.1016/j.envpol.2019.04.060.
American Academy of Oral Pathology 47 (7), 641–651. https://doi.org/10.1111/ da Costa, E.P., Bottrel, S., Starling, M., Leão, M., Amorim, C.C., 2019. Degradation of
jop.12701. carbendazim in water via photo-Fenton in Raceway Pond Reactor: assessment of
Briceño, G., Vergara, K., Schalchli, H., Palma, G., Tortella, G., Fuentes, M.S., Diez, M.C., acute toxicity and transformation products. Environ. Sci. Pollut. Res. Int. 26 (5),
2018. Organophosphorus pesticide mixture removal from environmental matrices by 4324–4336. https://doi.org/10.1007/s11356-018-2130-z.
a soil Streptomyces mixed culture. Environ. Sci. Pollut. Res. Int. 25 (22), Derbalah, A., Chidya, R., Kaonga, C., Iwamoto, Y., Takeda, K., Sakugawa, H., 2020.
21296–21307. https://doi.org/10.1007/s11356-017-9790-y. Carbaryl residue concentrations, degradation, and major sinks in the Seto Inland Sea.
Britch, S.C., Linthicum, K.J., Aldridge, R.L., Golden, F.V., Pongsiri, A., Khongtak, P., Japan. Environmental science and pollution research international 27 (13),
Ponlawat, A., 2018. Ultra-low volume Application of spinosad (natular 2EC) 14668–14678. https://doi.org/10.1007/s11356-020-08010-0.
larvicide as a residual in a tropical environment against Aedes and Anopheles Derbalah, A., Ismail, A., Hamza, A., Shaheen, S., 2014. Monitoring and remediation of
species. J. Am. Mosq. Contr. Assoc. 34 (1), 58–62. https://doi.org/10.2987/17- organochlorine residues in water. Water Environ. Res. : a research publication of the
6692.1. Water Environment Federation 86 (7), 584–593. https://doi.org/10.2175/
Brito, G.M., Roldi, L.L., Schetino Jr., M.Â., Checon Freitas, J.C., Cabral Coelho, E.R., 106143014x13975035525221.
2020. High-performance of activated biocarbon based on agricultural biomass waste Derbalah, A., Sunday, M., Chidya, R., Jadoon, W., Sakugawa, H., 2019. Kinetics of
applied for 2,4-D herbicide removing from water: adsorption, kinetic and photocatalytic removal of imidacloprid from water by advanced oxidation processes
thermodynamic assessments. J. Environ. Sci. Health - Part B Pesticides, Food with respect to nanotechnology. J. Water Health 17 (2), 254–265. https://doi.org/
Contam. Agric. Wastes 55 (9), 767–782. https://doi.org/10.1080/ 10.2166/wh.2019.259.
03601234.2020.1783178. Deziel, N.C., Colt, J.S., Kent, E.E., Gunier, R.B., Reynolds, P., Booth, B., Metayer, C.,
Brooks, A.J., Lim, H., Kilduff, J.E., 2012. Adsorption uptake of synthetic organic Ward, M.H., 2015. Associations between self-reported pest treatments and pesticide
chemicals by carbon nanotubes and activated carbons. Nanotechnology 23 (29), concentrations in carpet dust. Environ. Health : a global access science source 14, 27.
294008. https://doi.org/10.1088/0957-4484/23/29/294008. https://doi.org/10.1186/s12940-015-0015-x.
Bruce-Vanderpuije, P., Megson, D., Reiner, E.J., Bradley, L., Adu-Kumi, S., Gardella Jr., J. Díaz, G., Ortiz, R., Schettino, B., Vega, S., Gutiérrez, R., 2009. Organochlorine pesticides
A., 2019. The state of POPs in Ghana- A review on persistent organic pollutants: residues in bottled drinking water from Mexico City. Bull. Environ. Contam. Toxicol.
environmental and human exposure. Environ. Pollut. 245, 331–342. https://doi.org/ 82 (6), 701–704. https://doi.org/10.1007/s00128-009-9687-7.
10.1016/j.envpol.2018.10.107. Donald, D.B., Cessna, A.J., Sverko, E., Glozier, N.E., 2007. Pesticides in surface drinking-
Buah-Kwofie, A., Humphries, M.S., 2017. The distribution of organochlorine pesticides in water supplies of the northern Great Plains. Environ. Health Perspect. 115 (8),
sediments from iSimangaliso Wetland Park: ecological risks and implications for 1183–1191. https://doi.org/10.1289/ehp.9435.
conservation in a biodiversity hotspot. Environ. Pollut. 229, 715–723. https://doi. Dores, E.F., Carbo, L., Ribeiro, M.L., De-Lamonica-Freire, E.M., 2008. Pesticide levels in
org/10.1016/j.envpol.2017.07.031. ground and surface waters of primavera do leste region, mato grosso, Brazil.
Buhl, K., Stone, D., Powers, L., 2010. Bedbug-related Pesticide Incidents Reported to the J. Chromatogr. Sci. 46 (7), 585–590. https://doi.org/10.1093/chromsci/46.7.585.
National Pesticide Information Center (Poster). Oregon State University, Department Eitzer, B.D., Chevalier, A., 1999. Landscape care pesticide residues in residential drinking
of Environmental and Molecular Toxicology. National Pesticide Information Center water wells. Bull. Environ. Contam. Toxicol. 62 (4), 420–427. https://doi.org/
(NPIC) .pdf Last accessed September 1, 2019. http://npic.orst.edu/NPICbedbugpos 10.1007/s001289900892.
ter101510. El Bakouri, H., Morillo, J., Usero, J., Ouassini, A., 2007. Removal of prioritary pesticides
Bulut, S., Erdoðmuþ, S.F., Konuk, M., Cemek, M., 2010. The organochlorine pesticide contamining r’mel ground water by using organic waste residues. Commun. Agric.
residues in the drinking waters of afyonkarahisar, Turkey. Ekoloji 19 (74), 24–31. Appl. Biol. Sci. 72 (2), 197–207.
Burri, N.M., Weatherl, R., Moeck, C., Schirmer, M., 2019. A review of threats to Elfikrie, N., Ho, Y.B., Zaidon, S.Z., Juahir, H., Tan, E., 2020. Occurrence of pesticides in
groundwater quality in the anthropocene. Sci. Total Environ. 684, 136–154. https:// surface water, pesticides removal efficiency in drinking water treatment plant and
doi.org/10.1016/j.scitotenv.2019.05.236. potential health risk to consumers in Tengi River Basin, Malaysia. Sci. Total Environ.
Buxton, V.G., Greenstock, L.C., Helman, P.W., Ross, B.A., 1988. Critical Review of the 712, 136540. https://doi.org/10.1016/j.scitotenv.2020.136540.
rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl Elibariki, R., Maguta, M.M., 2017. Status of pesticides pollution in Tanzania - a review.
radicals (*OH/*O-) in aqueous solution. J. Phys. Chem. Ref. Data 17, 513–886. Chemosphere 178, 154–164. https://doi.org/10.1016/j.chemosphere.2017.03.036.
Ccanccapa, A., Masiá, A., Andreu, V., Picó, Y., 2016. Spatio-temporal patterns of El-Nahhal, Y., 2020. Pesticide residues in honey and their potential reproductive toxicity.
pesticide residues in the Turia and Júcar Rivers (Spain). Sci. Total Environ. 540, Sci. Total Environ. 741, 139953. https://doi.org/10.1016/j.scitotenv.2020.139953.
200–210. https://doi.org/10.1016/j.scitotenv.2015.06.063. El-Nahhal, Y., El-Nahhal, I., 2021. Cardiotoxicity of some pesticides and their
Centers for Disease Control and Prevention, 2011. Acute illnesses associated with amelioration. Environmental science and pollution research. https://doi.
insecticides used to control bed bugs – seven states, 2003-2010. Morb. Mortal. Wkly. org/10.1007/s11356-021-14999-9.
Rep. 60 (37), 1269–1274. Last accessed September 1, 2019. https://www.cdc. El-Nahhal, Y., Nir, S., Polubesova, T., Margulies, L., Rubin, B., 1998. Leaching,
gov/mmwr/preview/mmwrhtml/mm6037a1.htm. phytotoxicity and weed control of new formulations of alachlor. J. Agric. Food
Chaskopoulou, A., Miaoulis, M., Kashefi, J., 2018. Ground ultra low volume (ULV) space Chem. 46, 3305–3313.
spray applications for the control of wild sand fly populations (Psychodidae: El-Nahhal, Y., Safi, J., 2008. Removal of pesticide residues from water by organo-
phlebotominae) in Europe. Acta Trop. 182, 54–59. https://doi.org/10.1016/j. bentonites. In: Proc. The Twelfth International Water Technology Conference,
actatropica.2018.02.003. pp. 1711–1724 (Alexandria, Egypt).
Chaza, C., Sopheak, N., Mariam, H., David, D., Baghdad, O., Moomen, B., 2018. El-Nahhal, I., Redon, R., Raynaud, M., El-Nahhal, Y., Mounier, S., 2020. Characterization
Assessment of pesticide contamination in Akkar groundwater, northern Lebanon. of the fate and changes of post-irradiance fluorescence signal of filtered
Environ. Sci. Pollut. Res. Int. 25 (15), 14302–14312. https://doi.org/10.1007/ anthropogenic effluent dissolved organic matter from wastewater treatment plant in
s11356-017-8568-6. the coastal zone of Gapeau river. Environ. Sci. Pollut. Res. Int. 27 (18),
Cheng, F., Li, H., Qi, H., Han, Q., You, J., 2017. Contribution of pyrethroids in large 23141–23158. https://doi.org/10.1007/s11356-020-08842-w.
urban rivers to sediment toxicity assessed with benthic invertebrates Chironomus El-Nahhal, Y., 2004. Contamination and safety status of plant food in arab countries.
dilutus: a case study in South China. Environ. Toxicol. Chem. 36 (12), 3367–3375. J. Appl. Sci. 4 (3), 411–417. https://doi.org/10.1016/10.3923/jas.2004.411.417.
https://doi.org/10.1002/etc.3919. El-Nahhal, Y., Hamdona, N., 2017. Adsorption, leaching and phytotoxicity of some
Cho, Y.S., Lee, Y., Park, J.K., 2020. Fabrication of silica microspheres containing TiO₂ or herbicides as single and mixtures to some crops. Journal of the Association of Arab
aluminum zinc oxide nanoparticles via self-assembly: application in water Universities for Basic and Applied Sciences 17–25. https://doi.org/10.1016/j.
purification. J. Nanosci. Nanotechnol. 20 (11), 6738–6746. https://doi.org/ jaubas.2016.01.001.
10.1166/jnn.2020.18793. El-Nahhal, Y., Hammad, S., 2019. The prevalence of head lice and health consequence of
Chowdhury, M.A., Banik, S., Uddin, B., Moniruzzaman, M., Karim, N., Gan, S.H., 2012. their medical treatments among female students from different schools. Int. J. Pure
Organophosphorus and carbamate pesticide residues detected in water samples App. Biosci. 7 (1), 35–45. https://doi.org/10.18782/2320-7051.7196.
collected from paddy and vegetable fields of the Savar and Dhamrai Upazilas in El-Nahhal, Y., Nir, S., Serban, C., Rabinovitch, O., Rubin, B., 2000. Montmorillonite-
Bangladesh. Int. J. Environ. Res. Publ. Health 9 (9), 3318–3329. https://doi.org/ phenyltrimethylammonium yields environmentally improved formulations of
10.3390/ijerph9093318. hydrophobic herbicides. J. Agric. Food Chem. 48 (10), 4791–4801. https://doi.org/
Chu, W., Lau, T.K., Fung, S.C., 2006. Effects of combined and sequential addition of dual 10.1021/jf000327j.
oxidants (H2O2/S2O8(2-)) on the aqueous carbofuran photodegradation. J. Agric. El-Nahhal, Y., Safi, J., 2010. Adsorption of bromoxynil to modified bentonite: influence
Food Chem. 54 (26), 10047–10052. https://doi.org/10.1021/jf062018k. of pH, and temperature. J. Pestic. Sci. 35 (3), 333–338.
Claver, A., Ormad, P., Rodríguez, L., Ovelleiro, J.L., 2006. Study of the presence of Elsayed, O.F., Maillard, E., Vuilleumier, S., Nijenhuis, I., Richnow, H.H., Imfeld, G.,
pesticides in surface waters in the Ebro river basin (Spain). Chemosphere 64 (9), 2014. Using compound-specific isotope analysis to assess the degradation of
1437–1443. https://doi.org/10.1016/j.chemosphere.2006.02.034. chloroacetanilide herbicides in lab-scale wetlands. Chemosphere 99, 89–95. https://
Colombo, R., Ferreira, T.C., Alves, S.A., Carneiro, R.L., Lanza, M.R., 2013. Application of doi.org/10.1016/j.chemosphere.2013.10.027.
the response surface and desirability design to the Lambda-cyhalothrin degradation Erol, K., Yıldız, E., Alacabey, İ., Karabörk, M., Uzun, L., 2019. Magnetic diatomite for
using photo-Fenton reaction. J. Environ. Manag. 118, 32–39. https://doi.org/ pesticide removal from aqueous solution via hydrophobic interactions. Environ. Sci.
10.1016/j.jenvman.2012.12.035. Pollut. Res. Int. 26 (32), 33631–33641. https://doi.org/10.1007/s11356-019-
Craddock, H.A., Huang, D., Turner, P.C., Quirós-Alcalá, L., Payne-Sturges, D.C., 2019. 06423-0.
Trends in neonicotinoid pesticide residues in food and water in the United States,

25
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Escoto, D.F., Gayer, M.C., Bianchini, M.C., da Cruz Pereira, G., Roehrs, R., Denardin, E., Gupta, V.K., Gupta, B., Rastogi, A., Agarwal, S., Nayak, A., 2011. Pesticides removal from
2019. Use of Pistia stratiotes for phytoremediation of water resources contaminated waste water by activated carbon prepared from waste rubber tire. Water Res. 45
by clomazone. Chemosphere 227, 299–304. https://doi.org/10.1016/j. (13), 4047–4055. https://doi.org/10.1016/j.watres.2011.05.016.
chemosphere.2019.04.013. Gutiérrez, R.F., Santiesteban, A., Cruz-López, L., Bello-Mendoza, R., 2007. Removal of
Esposito, S., Garrone, E., Marocco, A., Pansini, M., Martinelli, P., Sannino, F., 2016. chlorothalonil, methyl parathion and methamidophos from water by the Fenton
Application of highly porous materials for simazine removal from aqueous solutions. reaction. Environ. Technol. 28 (3), 267–272. https://doi.org/10.1080/
Environ. Technol. 37 (19), 2428–2434. https://doi.org/10.1080/ 09593332808618787.
09593330.2016.1151461. He, W., Ye, M., He, H., Zhu, M., Li, Y., 2020. The decomposition and ecological risk of
Fatoki, O.S., Awofolu, O.R., 2004. Levels of organochlorine pesticide residues in marine-, DDTs and HCHs in the soil-water system of the Meijiang River. Environ. Res. 180,
surface-, ground- and drinking waters from the Eastern Cape Province of South 108897. https://doi.org/10.1016/j.envres.2019.108897.
Africa. J. Environ. Sci. Health - Part B Pesticides, Food Contam. Agric. Wastes 39 (1), Health Consultation, 2010. Pesticide misapplication in a private residence, parsons,
101–114. https://doi.org/10.1081/pfc-120027442. labette county, KS. U.S. Department of health and human services. Agency for Toxic
Femia, J., Mariani, M., Zalazar, C., Tiscornia, I., 2013. Photodegradation of chlorpyrifos Substances and Disease Registry. https://www.atsdr.cdc.gov/HAC/pha/Parsons
in water by UV/H2O2 treatment: toxicity evaluation. Water Sci. Technol. : a journal MalathionStrikeTeam/ParsonsMalathionStrikeTeamHC1117.pdf. (Accessed 1
of the International Association on Water Pollution Research 68 (10), 2279–2286. September 2019).
https://doi.org/10.2166/wst.2013.493. Hearon, S.E., Wang, M., Phillips, T.D., 2020. Strong adsorption of dieldrin by parent and
Fenton, H.J.H., 1894. Oxidation of tartaric acid in the presence of iron. J. Chem. Soc. processed montmorillonite clays. Environ. Toxicol. Chem. 39 (3), 517–525. https://
Trans. 65, 899–910. https://doi.org/10.1039/CT8946500899. doi.org/10.1002/etc.4642.
Flores-García, M.E., Molina-Morales, Y., Balza-Quintero, A., Benítez-Díaz, P.R., Miranda- Heidar, H., Seyed Taghi Omid, N., Abbasali, Z., 2017. Monitoring organophosphorous
Contreras, L., 2011. Residuos de plaguicidas en aguas para consumo humano en una pesticides residues in the shahid rajaei dam reservoir, sari, Iran. Bull. Environ.
comunidad agrícola del estado Mérida, Venezuela [Pesticide residues in drinking Contam. Toxicol. 98 (6), 791–797. https://doi.org/10.1007/s00128-017-2080-z.
water of an agricultural community in the state of Mérida, Venezuela]. Invest. Clin. Hgeig, A., Novaković, M., Mihajlović, I., 2019. Sorption of carbendazim and linuron from
52 (4), 295–311. aqueous solutions with activated carbon produced from spent coffee grounds:
Forero, L.G., Limay-Rios, V., Xue, Y., Schaafsma, A., 2017. Concentration and movement equilibrium, kinetic and thermodynamic approach. J. Environ. Sci. Health - Part B
of neonicotinoids as particulate matter downwind during agricultural practices using Pesticides, Food Contam. Agric. Wastes 54 (4), 226–236. https://doi.org/10.1080/
air samplers in southwestern Ontario, Canada. Chemosphere 188, 130–138. 03601234.2018.1550307.
Fosu-Mensah, B.Y., Okoffo, E.D., Darko, G., Gordon, C., 2016. Assessment of Hladik, M.L., Kuivila, K.M., 2009. Assessing the occurrence and distribution of
organochlorine pesticide residues in soils and drinking water sources from cocoa pyrethroids in water and suspended sediments. J. Agric. Food Chem. 57 (19),
farms in Ghana. SpringerPlus 5 (1), 869. https://doi.org/10.1186/s40064-016- 9079–9085. https://doi.org/10.1021/jf9020448.
2352-9. Hodomihou, N.R., Feder, F., Legros, S., Formentini, T.A., Lombi, E., Doelsch, E., 2020.
Gamboa-Carballo, J.J., Melchor-Rodríguez, K., Hernández-Valdés, D., Enriquez- Zinc speciation in organic waste drives its fate in amended soils. Environ. Sci.
Victorero, C., Montero-Alejo, A.L., Gaspard, S., Jáuregui-Haza, U.J., 2016. Technol. 54 (19), 12034–12041. https://doi.org/10.1021/acs.est.0c02721.
Theoretical study of chlordecone and surface groups interaction in an activated Hoh, E., Hites, R.A., 2004. Sources of toxaphene and other organochlorine pesticides in
carbon model under acidic and neutral conditions. J. Mol. Graph. Model. 65, 83–93. North America as determined by air measurements and potential source contribution
https://doi.org/10.1016/j.jmgm.2016.02.008. function analyses. Environ. Sci. Technol. 38 (15), 4187–4194. https://doi.org/
García-Galán, M.J., Monllor-Alcaraz, L.S., Postigo, C., Uggetti, E., López de Alda, M., 10.1021/es0499290.
Díez-Montero, R., García, J., 2020. Microalgae-based bioremediation of water Hosseini, N., Toosi, M.R., 2019. Removal of 2,4-D, glyphosate, trifluralin, and butachlor
contaminated by pesticides in peri-urban agricultural areas. Environ. Pollut. 265 (Pt herbicides from water by polysulfone membranes mixed by graphene oxide/TiO2
B), 114579. https://doi.org/10.1016/j.envpol.2020.114579. nanocomposite: study of filtration and batch adsorption. Journal of environmental
Garcinuño, R.M., Fernandez Hernando, P., Camara, C., 2006. Removal of carbaryl, health science & engineering 17 (1), 247–258. https://doi.org/10.1007/s40201-
linuron, and permethrin by Lupinus angustifolius under hydroponic conditions. 019-00344-3.
J. Agric. Food Chem. 54 (14), 5034–5039. https://doi.org/10.1021/jf060850j. Hulbert, D., Raja Jamil, R.Z., Isaacs, R., Vandervoort, C., Erhardt, S., Wise, J., 2020.
Garrido, I., Fenoll, J., Flores, P., Hellín, P., Pérez-Lucas, G., Navarro, S., 2020. Solar Leaching of insecticides used in blueberry production and their toxicity to red worm.
photocatalysis as strategy for on-site reclamation of agro-wastewater polluted with Chemosphere 241, 125091. https://doi.org/10.1016/j.chemosphere.2019.125091.
pesticide residues on farms using a modular facility. Environ. Sci. Pollut. Res. Int. Hustedt, J.C., Boyce, R., Bradley, J., Hii, J., Alexander, N., 2020. Use of pyriproxyfen in
https://doi.org/10.1007/s11356-020-10631-4, 10.1007/s11356-020-10631-4. control of Aedes mosquitoes: a systematic review. PLoS Neglected Trop. Dis. 14 (6),
Advance online publication. e0008205 https://doi.org/10.1371/journal.pntd.0008205.
Ghauch, A., Tuqan, A., 2008. Catalytic degradation of chlorothalonil in water using Hwang, J.I., Li, Z., Andreacchio, N., Ordonez Hinz, F., Wilson, P.C., 2020. Potential use
bimetallic iron-based systems. Chemosphere 73 (5), 751–759. https://doi.org/ of floating treatment wetlands established with Canna flaccida for removing organic
10.1016/j.chemosphere.2008.06.035. contaminants from surface water. Int. J. Phytoremediation 22 (12), 1304–1312.
Giddings, J., Gagne, J., Sharp, J., 2016. Synergistic effect of piperonyl butoxide on acute https://doi.org/10.1080/15226514.2020.1768511.
toxicity of pyrethrins to Hyalella azteca. Environ. Toxicol. Chem. 35 (8), 2111–2116. Hyland, C., Gunier, R.B., Metayer, C., Bates, M.N., Wesseling, C., Mora, A.M., 2018.
https://doi.org/10.1002/etc.3373. Maternal residential pesticide use and risk of childhood leukemia in Costa Rica. Int.
Glaze, W.H., Kang, J.W., Chapin, D.H., 1987. The chemistry of water treatment processes J. Canc. 143 (6), 1295–1304. https://doi.org/10.1002/ijc.31522.
involving ozone, hydrogen peroxide and ultraviolet radiation. Ozone: Sci. Eng. 9 (9), Iglesias, O., Gómez, J., Pazos, M., Sanromán, M.A., 2014. Electro-Fenton oxidation of
335_352. https://doi.org/10.1080/01919518708552148. Available from: imidacloprid by Fe alginate gel beads. Appl. Catal. B Environ. 144, 416–424. https://
Gómez-Sanabria, A., Zusman, E., Höglund-Isaksson, L., Klimont, Z., Lee, S.Y., doi.org/10.1016/j.apcatb.2013.07.046.
Akahoshi, K., Farzaneh, H., Chairunnisa, 2020. Sustainable wastewater management Imfeld, G., Meite, F., Wiegert, C., Guyot, B., Masbou, J., Payraudeau, S., 2020. Do rainfall
in Indonesia’s fish processing industry: bringing governance into scenario analysis. characteristics affect the export of copper, zinc and synthetic pesticides in surface
J. Environ. Manag. 275, 111241. https://doi.org/10.1016/j.jenvman.2020.111241. runoff from headwater catchments? Sci. Total Environ. 741, 140437. https://doi.
Gonzalez, G., Sagarzazu, A., Zoltan, T., 2013. Infuence of microstructure in drug release org/10.1016/j.scitotenv.2020.140437.
behavior of silica nanocapsules. Journal of drug delivery. https://doi.org/10.1155/ Jantunen, L.M., Wong, F., Gawor, A., Kylin, H., Helm, P.A., Stern, G.A., Strachan, W.M.,
2013/803585, 2013, 803585. Burniston, D.A., Bidleman, T.F., 2015. 20 Years of air-water gas exchange
González, T., Dominguez, J.R., Correia, S., 2020. Neonicotinoids removal by associated observations for pesticides in the western arctic ocean. Environ. Sci. Technol. 49
binary, tertiary and quaternary advanced oxidation processes: synergistic effects, (23), 13844–13852. https://doi.org/10.1021/acs.est.5b01303.
kinetics and mineralization. J. Environ. Manag. 261, 110156. https://doi.org/ Jariyal, M., Jindal, V., Mandal, K., Gupta, V.K., Singh, B., 2018. Bioremediation of
10.1016/j.jenvman.2020.110156. organophosphorus pesticide phorate in soil by microbial consortia. Ecotoxicol.
Gorito, A.M., Ribeiro, A.R., Almeida, C., Silva, A., 2017. A review on the application of Environ. Saf. 159, 310–316. https://doi.org/10.1016/j.ecoenv.2018.04.063.
constructed wetlands for the removal of priority substances and contaminants of Jeon, J.W., Kim, C.S., Kim, L., Lee, S.E., Kim, H.J., Lee, C.H., Choi, S.D., 2019.
emerging concern listed in recently launched EU legislation. Environ. Pollut. 227, Distribution and diastereoisomeric profiles of hexabromocyclododecanes in air,
428–443. https://doi.org/10.1016/j.envpol.2017.04.060. water, soil, and sediment samples in South Korea: application of an optimized
Gorito, A.M., Ribeiro, A.R., Gomes, C.R., Almeida, C., Silva, A., 2018. Constructed analytical method. Ecotoxicol. Environ. Saf. 181, 321–329. https://doi.org/
wetland microcosms for the removal of organic micropollutants from freshwater 10.1016/j.ecoenv.2019.06.015.
aquaculture effluents. Sci. Total Environ. 644, 1171–1180. https://doi.org/10.1016/ Jeong, I.S., Kwak, B.M., Ahn, J.H., Jeong, S.H., 2012. Determination of pesticide residues
j.scitotenv.2018.06.371. in milk using a QuEChERS-based method developed by response surface
Grassi, M., Kaykioglu, G., Belgiorno, V., Lofrano, G., 2012. Removal of emerging methodology. Food Chem. 133 (2), 473–481. https://doi.org/10.1016/j.
contaminants from water and wastewater by adsorption process. In: Emerging foodchem.2012.01.004.
Compounds Removal from Wastewater. Springer, Dordrecht, pp. 15–37. Jiang, W., Lin, K., Haver, D., Qin, S., Ayre, G., Spurlock, F., Gan, J., 2010. Wash-off
Gunier, R.B., Ward, M.H., Airola, M., Bell, E.M., Colt, J., Nishioka, M., Buffler, P.A., potential of urban use insecticides on concrete surfaces. Environ. Toxicol. Chem. 29
Reynolds, P., Rull, R.P., Hertz, A., Metayer, C., Nuckols, J.R., 2011. Determinants of (6), 1203–1208. https://doi.org/10.1002/etc.184.
agricultural pesticide concentrations in carpet dust. Environ. Health Perspect. 119 Jiang, W., Soeprono, A., Rust, M.K., Gan, J., 2014. Ant control efficacy of pyrethroids and
(7), 970–976. https://doi.org/10.1289/ehp.1002532. fipronil on outdoor concrete surfaces. Pest Manag. Sci. 70 (2), 271–277. https://doi.
Gupta, V.K., Ali, I., 2008. Removal of endosulfan and methoxychlor from water on org/10.1002/ps.3555.
carbon slurry. Environ. Sci. Technol. 42 (3), 766–770. https://doi.org/10.1021/ Kamel, A., Byrne, C., Vigo, C., Ferrario, J., Stafford, C., Verdin, G., Siegelman, F.,
es7025032. Knizner, S., Hetrick, J., 2009. Oxidation of selected organophosphate pesticides

26
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

during chlorination of simulated drinking water. Water Res. 43 (2), 522–534. Lamers, M., Anyusheva, M., La, N., Nguyen, V.V., Streck, T., 2011. Pesticide pollution in
https://doi.org/10.1016/j.watres.2008.10.038. surface- and groundwater by paddy rice cultivation: a case study from northern
Kapsi, M., Tsoutsi, C., Paschalidou, A., Albanis, T., 2019. Environmental monitoring and Vietnam. Clean 39, 356–361. https://doi.org/10.1002/clen.201000268.
risk assessment of pesticide residues in surface waters of the Louros River (N.W. Lehmann, E., Turrero, N., Kolia, M., Konaté, Y., de Alencastro, L.F., 2017. Dietary risk
Greece). Sci. Total Environ. 650 (Pt 2), 2188–2198. https://doi.org/10.1016/j. assessment of pesticides from vegetables and drinking water in gardening areas in
scitotenv.2018.09.185. Burkina Faso. Sci. Total Environ. 601–602, 1208–1216. https://doi.org/10.1016/j.
Karlsson, H., Muir, D.C.G., Teixiera, C.F., Burniston, D.A., Strachan, W.M.J., Hecky, R.E., scitotenv.2017.05.285.
Mwita, J., Bootsma, H.A., Grift, N.P., Kidd, K.A., Rosenberg, B., Rosenberg, B., 2000. Li, B., Qu, C., Bi, J., 2012. Identification of trace organic pollutants in drinking water and
Persistent chlorinated pesticides in air , water , and precipitation from the Lake. the associated human health risks in Jiangsu Province, China. Bull. Environ. Contam.
Malawi area , Southern Africa. Environ. Sci. Technol. 34, 4490–4495. Toxicol. 88 (6), 880–884. https://doi.org/10.1007/s00128-012-0619-6.
Karpuzcu, M.E., Fairbairn, D., Arnold, W.A., Barber, B.L., Kaufenberg, E., Koskinen, W. Li, H., Cheng, F., Wei, Y., Lydy, M.J., You, J., 2017. Global occurrence of pyrethroid
C., Novak, P.J., Rice, P.J., Swackhamer, D.L., 2014. Identifying sources of emerging insecticides in sediment and the associated toxicological effects on benthic
organic contaminants in a mixed use watershed using principal components analysis. invertebrates: an overview. J. Hazard Mater. 324 (Pt B), 258–271. https://doi.org/
Environmental science. Processes & impacts 16 (10), 2390–2399. https://doi.org/ 10.1016/j.jhazmat.2016.10.056.
10.1039/c4em00324a. Li, H., Tyler Mehler, W., Lydy, M.J., You, J., 2011. Occurrence and distribution of
Katagi, T., 2010. Bioconcentration, bioaccumulation, and metabolism of pesticides in sediment-associated insecticides in urban waterways in the Pearl River Delta, China.
aquatic organisms. Rev. Environ. Contam. Toxicol. 204, 1–132. https://doi.org/ Chemosphere 82 (10), 1373–1379. https://doi.org/10.1016/j.
10.1007/978-1-4419-1440-8_1. chemosphere.2010.11.074.
Kaushik, C.P., Sharma, H.R., Kaushik, A., 2012. Organochlorine pesticide residues in Li, X., Tian, T., Shang, X., Zhang, R., Xie, H., Wang, X., Wang, H., Xie, Q., Chen, J.,
drinking water in the rural areas of Haryana, India. Environ. Monit. Assess. 184 (1), Kadokami, K., 2020b. Occurrence and health risks of organic micro-pollutants and
103–112. https://doi.org/10.1007/s10661-011-1950-9. metals in groundwater of Chinese rural areas. Environ. Health Perspect. 128 (10),
Kenny, J.D., Webber, B.D., Howe, E.W., Holden, R.B., 2018. Pesticide removal through 107010. https://doi.org/10.1289/EHP6483.
wastewater and advanced treatment: full-scale sampling and bench-scale testing. Li, Y., Lohmann, R., Zou, X., Wang, C., Zhang, L., 2020. Air-water exchange and
Water Sci. Technol. : a journal of the International Association on Water Pollution distribution pattern of organochlorine pesticides in the atmosphere and surface
Research 77 (3–4), 739–747. https://doi.org/10.2166/wst.2017.586. water of the open Pacific ocean. Environ. Pollut. 265 (Pt A), 114956. https://doi.
Khedr, T., Hammad, A.A., Elmarsafy, A.M., Halawa, E., Soliman, M., 2019. Degradation org/10.1016/j.envpol.2020.114956.
of some organophosphorus pesticides in aqueous solution by gamma irradiation. Li, Z., Jennings, A., 2017. Worldwide regulations of standard values of pesticides for
J. Hazard Mater. 373, 23–28. https://doi.org/10.1016/j.jhazmat.2019.03.011. human health risk control: a review. Int. J. Environ. Res. Publ. Health 14 (7), 826.
Khuman, S.N., Vinod, P.G., Bharat, G., Kumar, Y., Chakraborty, P., 2020. Spatial https://doi.org/10.3390/ijerph14070826.
distribution and compositional profiles of organochlorine pesticides in the surface Liang, Z., Wang, X., Xie, B., Zhu, Y., Wu, J., Li, S., Meng, S., Zheng, X., Ji, A., Xie, L.,
soil from the agricultural, coastal and backwater transects along the south-west coast 2016. Pesticide exposure and risk of bladder cancer: a meta-analysis. Oncotarget 7
of India. Chemosphere 254, 126699. https://doi.org/10.1016/j. (41), 66959–66969. https://doi.org/10.18632/oncotarget.11397.
chemosphere.2020.126699. Lin, K., Haver, D., Oki, L., Gan, J., 2009. Persistence and sorption of fipronil degradates
Kiefer, K., Bader, T., Minas, N., Salhi, E., Janssen, E.M., von Gunten, U., Hollender, J., in urban stream sediments. Environ. Toxicol. Chem. 28 (7), 1462–1468. https://doi.
2020. Chlorothalonil transformation products in drinking water resources: org/10.1897/08-457.1.
widespread and challenging to abate. Water Res. 183, 116066. https://doi.org/ Liu, J., Qi, S., Yao, J., Yang, D., Xing, X., Liu, H., Qu, C., 2016. Contamination
10.1016/j.watres.2020.116066. characteristics of organochlorine pesticides in multimatrix sampling of the Hanjiang
Kim, L., Jeon, J.W., Son, J.Y., Kim, C.S., Ye, J., Kim, H.J., Lee, C.H., Hwang, S.M., Choi, S. River Basin, southeast China. Chemosphere 163, 35–43. https://doi.org/10.1016/j.
D., 2020. Nationwide levels and distribution of endosulfan in air, soil, water, and chemosphere.2016.07.040.
sediment in South Korea. Environ. Pollut. 265 (Pt B), 115035. https://doi.org/ Liu, P.Y., Li, B., Liu, H.D., Tian, L., 2014. Photochemical behavior of fenpropathrin and
10.1016/j.envpol.2020.115035. λ-cyhalothrin in solution. Environ. Sci. Pollut. Res. Int. 21 (3), 1993–2001. https://
Kiss, A., Virág, D., 2009. Photostability and photodegradation pathways of distinctive doi.org/10.1007/s11356-013-2119-6.
pesticides. J. Environ. Qual. 38 (1), 157–163. https://doi.org/10.2134/ Liu, R., Alarcon, W.A., Calvert, G.M., Aubin, K.G., Beckman, J., Cummings, K.R.,
jeq2007.0504. Graham, L.S., Higgins, S.A., Mulay, P., Patel, K., Prado, J.B., Schwartz, A., Stover, D.,
Klarich, K.L., Pflug, N.C., Dewald, E.M., Hladik, M.L., Kolpin, D.W., Cwiertny, D.M., Waltz, J., 2018. Acute illnesses and injuries related to total release foggers - 10
et al., 2017. Occurrence of neonicotinoid insecticides in finished drinking water and states, 2007-2015. MMWR. Morbidity and mortality weekly report 67 (4), 125–130.
fate during drinking water treatment. Environ. Sci. Technol. 4, 168–173. https://doi.org/10.15585/mmwr.mm6704a4.
Komtchou, S., Dirany, A., Drogui, P., Robert, D., Lafrance, P., 2017. Removal of atrazine Liu, T., Xu, S., Lu, S., Qin, P., Bi, B., Ding, H., Liu, Y., Guo, X., Liu, X., 2019. A review on
and its by-products from water using electrochemical advanced oxidation processes. removal of organophosphorus pesticides in constructed wetland: performance,
Water Res. 125, 91–103. https://doi.org/10.1016/j.watres.2017.08.036. mechanism and influencing factors. Sci. Total Environ. 651 (Pt 2), 2247–2268.
Koroša, A., Auersperger, P., Mali, N., 2016. Determination of micro-organic https://doi.org/10.1016/j.scitotenv.2018.10.087.
contaminants in groundwater (Maribor, Slovenia). Sci. Total Environ. 571, Llorent-Martínez, E.J., Ortega-Barrales, P., Fernández-de Córdova, M.L., Ruiz-
1419–1431. https://doi.org/10.1016/j.scitotenv.2016.06.103. Medina, A., 2011. Trends in flow-based analytical methods applied to pesticide
Kouzayha, A., Al Ashi, A., Al Akoum, R., Al Iskandarani, M., Budzinski, H., Jaber, F., detection: a review. Anal. Chim. Acta 684 (1–2), 21–30. https://doi.org/10.1016/j.
2013. Occurrence of pesticide residues in Lebanon’s water resources. Bull. Environ. aca.2010.10.036.
Contam. Toxicol. 91 (5), 503–509. https://doi.org/10.1007/s00128-013-1071-y. Lockridge, O., Verdier, L., Schopfer, L.M., 2019. Half-life of chlorpyrifos oxon and other
Kreuzig, R., Hartmann, C., Teigeler, J., Höltge, S., Cvetković, B., Schlag, P., 2010. organophosphorus esters in aqueous solution. Chem. Biol. Interact. 311, 108788.
Development of a novel concept for fate monitoring of biocides in liquid manure and https://doi.org/10.1016/j.cbi.2019.108788.
manured soil taking 14C-imazalil as an example. Chemosphere 79 (11), 1089–1094. Loffredo, E., Taskin, E., 2017. Adsorptive removal of ascertained and suspected
https://doi.org/10.1016/j.chemosphere.2010.03.014. endocrine disruptors from aqueous solution using plant-derived materials. Environ.
Kromer, T., Ophoff, H., Stork, A., Führ, F., 2004. Photodegradation and volatility of Sci. Pollut. Res. Int. 24 (23), 19159–19166. https://doi.org/10.1007/s11356-017-
pesticides: chamber experiments. Environ. Sci. Pollut. Res. Int. 11 (2), 107–120. 9595-z.
https://doi.org/10.1007/BF02979710. Lu, C., Chang, C.H., Palmer, C., Zhao, M., Zhang, Q., 2018. Neonicotinoid residues in
Kudom, A.A., Anane, L.N., Afoakwah, R., Adokoh, C.K., 2018. Relating high insecticide fruits and vegetables: an integrated dietary exposure assessment approach. Environ.
residues in larval breeding habitats in urban residential areas to the selection of Sci. Technol. 52 (5), 3175–3184. https://doi.org/10.1021/acs.est.7b05596.
pyrethroid resistance in Anopheles gambiae s.l. (Diptera: Culicidae) in akim oda, Lu, Y., Li, S., Sha, M., Wang, B., Cheng, G., Guo, Y., Zhu, J., 2020. Cascading effects
Ghana. J. Med. Entomol. 55 (2), 490–495. https://doi.org/10.1093/jme/tjx223. caused by fenoxycarb in freshwater systems dominated by Daphnia carinata and
Kunstadter, P., Prapamontol, T., Sirirojn, B.O., Sontirat, A., Tansuhaj, A., Dolerocypris sinensis. Ecotoxicol. Environ. Saf. 203, 111022. https://doi.org/
Khamboonruang, C., 2001. Pesticide exposures among Hmong farmers in Thailand. 10.1016/j.ecoenv.2020.111022.
Int. J. Occup. Environ. Health 7 (4), 313–325. https://doi.org/10.1179/ Lu, Z., Fang, N., Liu, Y., Zhang, Z., Pan, H., Hou, Z., Li, Y., Lu, Z., 2017. Dissipation and
107735201800339227. residues of the diamide insecticide chlorantraniliprole in ginseng ecosystems under
Kuster, M., Díaz-Cruz, S., Rosell, M., López de Alda, M., Barceló, D., 2010. Fate of different cultivation environments. Environ. Monit. Assess. 189 (11), 534. https://
selected pesticides, estrogens, progestogens and volatile organic compounds during doi.org/10.1007/s10661-017-6241-7.
artificial aquifer recharge using surface waters. Chemosphere 79 (8), 880–886. Lucas, N., Bienaime, C., Belloy, C., Queneudec, M., Silvestre, F., Nava-Saucedo, J.E.,
https://doi.org/10.1016/j.chemosphere.2010.02.026. 2008. Polymer biodegradation: mechanisms and estimation techniques.
Kusvuran, E., Yildirim, D., Mavruk, F., Ceyhan, M., 2012. Removal of chloropyrifos ethyl, Chemosphere 73 (4), 429–442. https://doi.org/10.1016/j.
tetradifon and chlorothalonil pesticide residues from citrus by using ozone. chemosphere.2008.06.064.
J. Hazard Mater. 241–242, 287–300. https://doi.org/10.1016/j. Lv, T., Zhang, Y., Zhang, L., Carvalho, P.N., Arias, C.A., Brix, H., 2016. Removal of the
jhazmat.2012.09.043. pesticides imazalil and tebuconazole in saturated constructed wetland mesocosms.
Lacayo-Romero, M., van Bavel, B., Mattiasson, B., 2006. Degradation of toxaphene in Water Res. 91, 126–136. https://doi.org/10.1016/j.watres.2016.01.007.
aged and freshly contaminated soil. Chemosphere 63 (4), 609–615. https://doi.org/ Ma, Y.S., Kumar, M., Lin, J.G., 2009. Degradation of carbofuran-contaminated water by
10.1016/j.chemosphere.2005.08.019. the Fenton process. J. Environ. Sci. Health - Part A Toxic/Hazard. Subst. Environ.
Lafi, W.K., Al-Qodah, Z., 2006. Combined advanced oxidation and biological treatment Eng. 44 (9), 914–920. https://doi.org/10.1080/10934520902958807.
processes for the removal of pesticides from aqueous solutions. J. Hazard Mater. 137 Mahai, G., Wan, Y., Xia, W., Wang, A., Shi, L., Qian, X., He, Z., Xu, S., 2021. A nationwide
(1), 489–497. https://doi.org/10.1016/j.jhazmat.2006.02.027. study of occurrence and exposure assessment of neonicotinoid insecticides and their

27
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

metabolites in drinking water of China. Water Res. 189, 116630. https://doi.org/ sundays river estuaries, eastern cape province, South Africa. Environ. Geochem.
10.1016/j.watres.2020.116630. Health 41 (6), 2777–2801. https://doi.org/10.1007/s10653-019-00336-0.
Mahler, B.J., Van Metre, P.C., Wilson, J.T., Musgrove, M., Zaugg, S.D., Burkhardt, M.R., Olisah, C., Okoh, O.O., Okoh, A.I., 2020. Occurrence of organochlorine pesticide residues
2009. Fipronil and its degradates in indoor and outdoor dust. Environ. Sci. Technol. in biological and environmental matrices in Africa: a two-decade review. Heliyon 6
43 (15), 5665–5670. https://doi.org/10.1021/es901292a. (3), e03518. https://doi.org/10.1016/j.heliyon.2020.e03518.
Malato, S., Blanco, J., Cáceres, J., Fernández-Alba, A.R., Agüera, A., Rodríguez, A., 2002. Ormad, M.P., Miguel, N., Claver, A., Matesanz, J.M., Ovelleiro, J.L., 2008. Pesticides
Photocatalytic treatment of water-soluble pesticides by photo-Fenton and TiO2 using removal in the process of drinking water production. Chemosphere 71 (1), 97–106.
solar energy. Catal. Today 76 (2–4), 209–220. https://doi.org/10.1016/S0920-5861 https://doi.org/10.1016/j.chemosphere.2007.10.006.
(02)00220-1. Page, D., Miotliński, K., Gonzalez, D., Barry, K., Dillon, P., Gallen, C., 2014.
Margot, J., Kienle, C., Magnet, A., Weil, M., Rossi, L., de Alencastro, L.F., Abegglen, C., Environmental monitoring of selected pesticides and organic chemicals in urban
Thonney, D., Chèvre, N., Schärer, M., Barry, D.A., 2013. Treatment of stormwater recycling systems using passive sampling techniques. J. Contam. Hydrol.
micropollutants in municipal wastewater: ozone or powdered activated carbon? Sci. 158, 65–77. https://doi.org/10.1016/j.jconhyd.2014.01.004.
Total Environ. 461–462, 480–498. https://doi.org/10.1016/j. Palma, P., Köck-Schulmeyer, M., Alvarenga, P., Ledo, L., Barbosa, I.R., López de Alda, M.,
scitotenv.2013.05.034. Barceló, D., 2014. Risk assessment of pesticides detected in surface water of the
Marican, A., Durán-Lara, E.F., 2018. A review on pesticide removal through different Alqueva reservoir (Guadiana basin, southern of Portugal). Sci. Total Environ.
processes. Environ. Sci. Pollut. Res. Int. 25 (3), 2051–2064. https://doi.org/ 488–489, 208–219. https://doi.org/10.1016/j.scitotenv.2014.04.088.
10.1007/s11356-017-0796-2. Palma, P., Kuster, M., Alvarenga, P., Palma, V.L., Fernandes, R.M., Soares, A.M., López de
Markowitz, S.B., 1992. Poisoning of an urban family due to misapplication of household Alda, M.J., Barceló, D., Barbosa, I.R., 2009. Risk assessment of representative and
organophosphate and carbamate pesticides. J. Toxicol. Clin. Toxicol. 30 (2), priority pesticides, in surface water of the Alqueva reservoir (South of Portugal)
295–303. https://doi.org/10.3109/15563659209038640. using on-line solid phase extraction-liquid chromatography-tandem mass
Marvin, C., Painter, S., Williams, D., Richardson, V., Rossmann, R., Van Hoof, P., 2004. spectrometry. Environ. Int. 35 (3), 545–551. https://doi.org/10.1016/j.
Spatial and temporal trends in surface water and sediment contamination in the envint.2008.09.015.
Laurentian Great Lakes. Environ. Pollut. 129 (1), 131–144. https://doi.org/10.1016/ Pan, X., Lin, D., Zheng, Y., Zhang, Q., Yin, Y., Cai, L., Fang, H., Yu, Y., 2016.
j.envpol.2003.09.029. Biodegradation of DDT by Stenotrophomonas sp. DDT-1: characterization and
Mawussi, G., Sanda, K., Merlina, G., Pinelli, E., 2009. Assessment of average exposure to genome functional analysis. Sci. Rep. 6, 21332. https://doi.org/10.1038/srep21332.
organochlorine pesticides in southern Togo from water, maize (Zea mays) and Panseri, S., Catalano, A., Giorgi, A., Arioli, F., Procopio, A., Britti, D., Chiesa, L., 2014.
cowpea (Vigna unguiculata). Food Addit. Contam. Part A, Chemistry, analysis, Occurrence of pesticide residues in Italian honey from different areas in relation to
control, exposure & risk assessment 26 (3), 348–354. https://doi.org/10.1080/ its potential contamination sources. Food Contr. 38, 150–156. https://doi.org/
02652030802528343. 10.1016/j.foodcont.2013.10.024.
McConnell, R., Pacheco, F., Wahlberg, K., Klein, W., Malespin, O., Magnotti, R., Pardo, L.A., Beane Freeman, L.E., Lerro, C.C., Andreotti, G., Hofmann, J.N., Parks, C.G.,
Akerblom, M., Murray, D., 1999. Subclinical health effects of environmental Sandler, D.P., Lubin, J.H., Blair, A., Koutros, S., 2020. Pesticide exposure and risk of
pesticide contamination in a developing country: cholinesterase depression in aggressive prostate cancer among private pesticide applicators. Environ. Health : a
children. Environ. Res. 81 (2), 87–91. https://doi.org/10.1006/enrs.1999.3958. global access science source 19 (1), 30. https://doi.org/10.1186/s12940-020-00583-
McManus, S.L., Coxon, C.E., Richards, K.G., Danaher, M., 2013. Quantitative solid phase 0.
microextraction - gas chromatography mass spectrometry analysis of the pesticides Pazi, I., Kucuksezgin, F., Gonul, L.T., 2011. Distribution and sources of
lindane, heptachlor and two heptachlor transformation products in groundwater. organochlorinated contaminants in sediments from Izmir Bay (eastern aegean sea).
J. Chromatogr., A 1284, 1–7. https://doi.org/10.1016/j.chroma.2013.01.099. Mar. Pollut. Bull. 62 (5), 1115–1119. https://doi.org/10.1016/j.
Mekonen, S., Argaw, R., Simanesew, A., Houbraken, M., Senaeve, D., Ambelu, A., marpolbul.2011.03.016.
Spanoghe, P., 2016. Pesticide residues in drinking water and associated risk to Pinasseau, L., Wiest, L., Volatier, L., Mermillod-Blondin, F., Vulliet, E., 2020. Emerging
consumers in Ethiopia. Chemosphere 162, 252–260. https://doi.org/10.1016/j. polar pollutants in groundwater: potential impact of urban stormwater infiltration
chemosphere.2016.07.096. practices. Environ. Pollut. 266 (Pt 2), 115387. https://doi.org/10.1016/j.
Milhome, M., de Lima, L.K., de A Nobre, C., de A F Lima, F., do Nascimento, R.F., 2018. envpol.2020.115387.
Effect of ozonization in degradation of trifluralin residues in aqueous and food Pirnie, E.F., Talley, J.W., Hundal, L.S., 2006. Abiotic transformation of DDT in aqueous
matrices. J. Environ. Sci. Health - Part B Pesticides, Food Contam. Agric. Wastes 53 solutions. Chemosphere 65 (9), 1576–1582. https://doi.org/10.1016/j.
(12), 786–792. https://doi.org/10.1080/03601234.2018.1505074. chemosphere.2006.03.055.
Miranda-Contreras, L., Gómez-Pérez, R., Rojas, G., Cruz, I., Berrueta, L., Salmen, S., et al., Poole, K.G., Elkin, B.T., Bethke, R.W., 1998. Organochlorine and heavy metal
2013. Occupational exposure to organophosphate and carbamate pesticides affects contaminants in wild mink in western Northwest Territories, Canada. Arch. Environ.
sperm chromatin integrity and reproductive hormone levels among Venezuelan farm Contam. Toxicol. 34 (4), 406–413. https://doi.org/10.1007/s002449900337.
workers. J. Occup. Health 55, 195–203. Posthuma, L., Zijp, M.C., De Zwart, D., Van de Meent, D., Globevnik, L., Koprivsek, M.,
Mohammed, S., Lamoree, M., Ansa-Asare, O.D., de Boer, J., 2019. Review of the analysis Focks, A., Van Gils, J., Birk, S., 2020. Chemical pollution imposes limitations to the
of insecticide residues and their levels in different matrices in Ghana. Ecotoxicol. ecological status of European surface waters. Sci. Rep. 10 (1), 14825. https://doi.
Environ. Saf. 171, 361–372. https://doi.org/10.1016/j.ecoenv.2018.12.049. org/10.1038/s41598-020-71537-2.
Mohapatra, S.P., Kumar, M., Gajbhiye, V.T., Agnihotri, N.P., 1995. Ground water Postle, J.K., Rheineck, B.D., Allen, P.E., Baldock, J.O., Cook, C.J., Zogbaum, R.,
contamination by organochlorine insecticide residues in a rural area in the indo- Vandenbrook, J.P., 2004. Chloroacetanilide herbicide metabolites in Wisconsin
gangetic plain. Environ. Monit. Assess. 35 (2), 155–164. https://doi.org/10.1007/ groundwater: 2001 survey results. Environ. Sci. Technol. 38 (20), 5339–5343.
BF00633712. https://doi.org/10.1021/es040399h.
Mondal, R., Mukherjee, A., Biswas, S., Kole, R.K., 2018. GC-MS/MS determination and Pouchieu, C., Piel, C., Carles, C., Gruber, A., Helmer, C., Tual, S., Marcotullio, E.,
ecological risk assessment of pesticides in aquatic system: a case study in Hooghly Lebailly, P., Baldi, I., 2018. Pesticide use in agriculture and Parkinson’s disease in
River basin in West Bengal, India. Chemosphere 206, 217–230. https://doi.org/ the AGRICAN cohort study. Int. J. Epidemiol. 47 (1), 299–310. https://doi.org/
10.1016/j.chemosphere.2018.04.168. 10.1093/ije/dyx225.
Morrissey, C.A., Mineau, P., Devries, J.H., Sanchez-Bayo, F., Liess, M., Cavallaro, M.C., PPDB A to Z Index - University of Hertfordshire, 2007. A to Z index for the PPDB -
Liber, K., 2015. Neonicotinoid contamination of global surface waters and associated pesticides properties DataBase. In: THE PPDB: Last Updated: 01/08/2020 A to Z List
risk to aquatic invertebrates: a review. Environ. Int. 74, 291–303. https://doi.org/ of Pesticide Active Ingredients.
10.1016/j.envint.2014.10.024. Prosser, R.S., Hoekstra, P.F., Gene, S., Truman, C., White, M., Hanson, M.L., 2020.
Mount, G.A., 1998. A critical review of ultralow-volume aerosols of insecticide applied A review of the effectiveness of vegetated buffers to mitigate pesticide and nutrient
with vehicle-mounted generators for adult mosquito control. J. Am. Mosq. Contr. transport into surface waters from agricultural areas. J. Environ. Manag. 261,
Assoc. 14 (3), 305–334. 110210. https://doi.org/10.1016/j.jenvman.2020.110210.
Munter, R., 2001. Advanced oxidation processes-current status and prospects. Proc. Est. Quirós-Alcalá, L., Bradman, A., Nishioka, M., Harnly, M.E., Hubbard, A., McKone, T.E.,
Acad. Sci. 50 (2), 59_80. Ferber, J., Eskenazi, B., 2011. Pesticides in house dust from urban and farmworker
Natasha, Shahid M., Khalid, S., Murtaza, B., Anwar, H., Shah, A.H., Sardar, A., households in California: an observational measurement study. Environ. Health : a
Shabbir, Z., Niazi, N.K., 2020. A critical analysis of wastewater use in agriculture and global access science source 10, 19. https://doi.org/10.1186/1476-069X-10-19.
associated health risks in Pakistan. Environ. Geochem. Health. https://doi.org/ Rafique, N., Tariq, S.R., 2015. Photodegradation of α-cypermethrin in soil in the presence
10.1007/s10653-020-00702-3, 10.1007/s10653-020-00702-3. Advance online of trace metals (Cu2+, Cd2+, Fe2+ and Zn2+). Environmental science. Processes &
publication. impacts 17 (1), 166–176. https://doi.org/10.1039/c4em00439f.
Navarrete, I.A., Tee, K., Unson, J., Hallare, A.V., 2018. Organochlorine pesticide residues Ramos-Delgado, N.A., Gracia-Pinilla, M.A., Maya-Treviño, L., Hinojosa-Reyes, L.,
in surface water and groundwater along Pampanga River, Philippines. Environ. Guzman-Mar, J.L., Hernández-Ramírez, A., 2013. Solar photocatalytic activity of
Monit. Assess. 190 (5), 289. https://doi.org/10.1007/s10661-018-6680-9. TiO2 modified with WO3 on the degradation of an organophosphorus pesticide.
Nir, S., Undabeytia, T., Yaron, D., El-Nahhal, Y., Polubesova, T., Serban, S., Rytwo, G., J. Hazard Mater. 263 (Pt 1), 36–44. https://doi.org/10.1016/j.
Lagaly, G., Rubin, B., 2000. Optimization of adsorption of hydrophobic herbicides on jhazmat.2013.07.058.
montmorillonite preadsorbed by monovalent organic cations: interaction between Rasoulifard, M.H., Akrami, M., Eskandarian, M.R., 2015. Degradation of
phenyl rings. Environ. Sci. Technol. 34, 1269–1274. organophosphorus pesticide diazinon using activated persulfate: optimization of
Nurul Amin, M., Kaneco, S., Kato, T., Katsumata, H., Suzuki, T., Ohta, K., 2008. Removal operational parameters and comparative study by Taguchi’s method. Journal of the
of thiobencarb in aqueous solution by zero valent iron. Chemosphere 70 (3), Taiwan Institute of Chemical Engineers 57, 77–90. https://doi.org/10.1016/j.
511–515. https://doi.org/10.1016/j.chemosphere.2007.09.017. jtice.2015.05.014.
Olisah, C., Adeniji, A.O., Okoh, O.O., Okoh, A.I., 2019. Occurrence and risk evaluation of Rawn, D.F., Lockhart, W.L., Wilkinson, P., Savoie, D.A., Rosenberg, G.B., Muir, D.C.,
organochlorine contaminants in surface water along the course of swartkops and 2001. Historical contamination of Yukon Lake sediments by PCBs and

28
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

organochlorine pesticides: influence of local sources and watershed characteristics. North China. Sci. Total Environ. 741, 140110. https://doi.org/10.1016/j.
Sci. Total Environ. 280 (1–3), 17–37. https://doi.org/10.1016/s0048-9697(01) scitotenv.2020.140110.
00798-7. Snow, D.D., Chakraborty, P., Uralbekov, B., Satybaldiev, B., Sallach, J.B., Thornton
Regnery, J., Parrhysius, P., Schulz, R.S., Möhlenkamp, C., Buchmeier, G., Hampton, L.M., Jeffries, M., Kolok, A.S., Bartelt-Hunt, S.B., 2020. Legacy and
Reifferscheid, G., Brinke, M., 2019. Wastewater-borne exposure of limnic fish to current pesticide residues in Syr Darya, Kazakhstan: contamination status, seasonal
anticoagulant rodenticides. Water Res. 167, 115090. https://doi.org/10.1016/j. variation and preliminary ecological risk assessment. Water Res. 184, 116141.
watres.2019.115090. https://doi.org/10.1016/j.watres.2020.116141.
Regnery, J., Schulz, R.S., Parrhysius, P., Bachtin, J., Brinke, M., Schäfer, S., Snyder, S.A., Adham, S., Redding, A.M., Cannon, F.S., Decarolis, J., Oppenheimer, J.,
Reifferscheid, G., Friesen, A., 2020. Heavy rainfall provokes anticoagulant et al., 2007. Role of membranes and activated carbon in the removal of endocrine
rodenticides’ release from baited sewer systems and outdoor surfaces into receiving disruptors and pharmaceuticals. Desalination 202, 156_181. https://doi.org/
streams. Sci. Total Environ. 740, 139905. https://doi.org/10.1016/j. 10.1016/j.desal.2005.12.052. Available from:
scitotenv.2020.139905. Stefan, I.M., 2017. Advanced Oxidation Processes for Water Treatment: Fundamentals
Reichenauer, T.G., Germida, J.J., 2008. Phytoremediation of organic contaminants in and Applications. IWA Publishing. https://doi.org/10.2166/9781780407197.
soil and groundwater. ChemSusChem 1 (8–9), 708–717. https://doi.org/10.1002/ Stehle, S., Bline, A., Bub, S., Petschick, L.L., Wolfram, J., Schulz, R., 2019. Aquatic
cssc.200800125. pesticide exposure in the U.S. as a result of non-agricultural uses. Environ. Int. 133
Rendon-von Osten, J., Dzul-Caamal, R., 2017. Glyphosate residues in groundwater, (Pt B), 105234. https://doi.org/10.1016/j.envint.2019.105234.
drinking water and urine of subsistence farmers from intensive agriculture localities: Su, X.N., Zhang, J.J., Liu, J.T., Zhang, N., Ma, L.Y., Lu, F.F., Chen, Z.J., Shi, Z., Si, W.J.,
a survey in hopelchén, campeche, Mexico. Int. J. Environ. Res. Publ. Health 14 (6), Liu, C., Yang, H., 2019. Biodegrading two pesticide residues in paddy plants and the
595. https://doi.org/10.3390/ijerph14060595. environment by a genetically engineered approach. J. Agric. Food Chem. 67 (17),
Rheinheimer Dos Santos, D., Monteiro de Castro Lima, J.A., Paranhos Rosa de Vargas, J., 4947–4957. https://doi.org/10.1021/acs.jafc.8b07251.
Camotti Bastos, M., Santanna Dos Santos, M.A., Mondamert, L., Labanowski, J., Suarez-Lopez, J.R., Nazeeh, N., Kayser, G., Suárez-Torres, J., Checkoway, H., López-
2020. Pesticide bioaccumulation in epilithic biofilms as a biomarker of agricultural Paredes, D., Jacobs Jr., D.R., Cruz, F., 2020. Residential proximity to greenhouse
activities in a representative watershed. Environ. Monit. Assess. 192 (6), 381. crops and pesticide exposure (via acetylcholinesterase activity) assessed from
https://doi.org/10.1007/s10661-020-08264-8. childhood through adolescence. Environ. Res. 188, 109728. https://doi.org/
Richards, J., Reif, R., Luo, Y., Gan, J., 2016. Distribution of pesticides in dust particles in 10.1016/j.envres.2020.109728.
urban environments. Environ. Pollut. 214, 290–298. https://doi.org/10.1016/j. Sultana, T., Murray, C., Kleywegt, S., Metcalfe, C.D., 2018. Neonicotinoid Pesticides in
envpol.2016.04.025. Drinking Water in Agricultural Regions of Southern Ontario. Canada, Chemosphere.
Roberts, J.R., Karr, C.J., Council On Environmental Health, 2012. Pesticide exposure in https://doi.org/10.1016/j.chemosphere.2018.02.108.
children. Pediatrics 130 (6), e1765–e1788. https://doi.org/10.1542/peds.2012- Tanabe, A., Mitobe, H., Kawata, K., Yasuhara, A., Shibamoto, T., 2001. Seasonal and
2758. spatial studies on pesticide residues in surface waters of the Shinano river in Japan.
Rodriguez, E., Campinas, M., Acero, J.L., Rosa, M.J., 2016. Investigating PPCP removal J. Agric. Food Chem. 49 (8), 3847–3852. https://doi.org/10.1021/jf010025x.
from wastewater by powdered activated carbon/ultrafiltration. Water Air Soil Tao, Y., Phung, D., Dong, F., Xu, J., Liu, X., Wu, X., et al., 2019. Urinary monitoring of
Pollut. 227 (6) https://doi.org/10.1007/s11270-016-2870-7. neonicotinoid imidacloprid exposure to pesticide applicators. Sci. Total Environ.
Roe, B.A., Lemley, A.T., 1997. Treatment of two insecticides in an electrochemical 669, 721–728.
Fenton system. J. Environ. Sci. Health - Part B Pesticides, Food Contam. Agric. Taylor, A.R., Li, J., Wang, J., Schlenk, D., Gan, J., 2019. Occurrence and probable sources
Wastes 32 (2), 261–281. https://doi.org/10.1080/03601239709373085. of urban-use insecticides in marine sediments off the coast of los angeles. Environ.
Rohani, A., Fakhriy, H.A., Suzilah, I., Zurainee, M.N., Najdah, W., Ariffin, M.M., Sci. Technol. 53 (16), 9584–9593. https://doi.org/10.1021/acs.est.9b02825.
Shakirudin, N.M., Afiq, M., Jenarun, J., Tanrang, Y., Lee, H.L., 2020. Indoor and Teklu, B.M., Adriaanse, P.I., Ter Horst, M.M., Deneer, J.W., Van den Brink, P.J., 2015.
outdoor residual spraying of a novel formulation of deltamethrin K-Othrine® Surface water risk assessment of pesticides in Ethiopia. Sci. Total Environ. 508,
(Polyzone) for the control of simian malaria in Sabah, Malaysia. PloS One 15 (5), 566–574. https://doi.org/10.1016/j.scitotenv.2014.11.049.
e0230860. https://doi.org/10.1371/journal.pone.0230860. Terzaghi, E., Vitale, C.M., Di Guardo, A., 2020. Modelling peak exposure of pesticides in
Rojas, R., Vanderlinden, E., Morillo, J., Usero, J., El Bakouri, H., 2014. Characterization terrestrial and aquatic ecosystems: importance of dissolved organic carbon and
of sorption processes for the development of low-cost pesticide decontamination vertical particle movement in soil. SAR QSAR Environ. Res. 31 (1), 19–32. https://
techniques. Sci. Total Environ. 488–489, 124–135. https://doi.org/10.1016/j. doi.org/10.1080/1062936X.2019.1686715.
scitotenv.2014.04.079. Thiombane, M., Petrik, A., Di Bonito, M., Albanese, S., Zuzolo, D., Cicchella, D., Lima, A.,
Rubin, C., Esteban, E., Kieszak, S., Hill Jr., R.H., Dunlop, B., Yacovac, R., Trottier, J., Qu, C., Qi, S., De Vivo, B., 2018. Status, sources and contamination levels of
Boylan, K., Tomasewski, T., Pearce, K., 2002. Assessment of human exposure and organochlorine pesticide residues in urban and agricultural areas: a preliminary
human health effects after indoor application of methyl parathion in Lorain County, review in central-southern Italian soils. Environ. Sci. Pollut. Res. Int. 25 (26),
Ohio, 1995-1996. Environ. Health Perspect. 110, 1047–1051. https://doi.org/ 26361–26382. https://doi.org/10.1007/s11356-018-2688-5.
10.1289/ehp.02110s61047. Suppl 6(Suppl 6). Thomas, J.E., Ou, L.T., Al-Agely, A., 2008. All-Agely A. DDE remediation and
Ruggieri, F., D’Archivio, A.A., Fanelli, M., Mazzeo, P., Paoletti, E., 2008. A multi- degradation. Rev. Environ. Contam. Toxicol. 194, 55–69. https://doi.org/10.1007/
lysimeter investigation on the mobility and persistence of pesticides in the loam soil 978-0-387-74816-0_3.
of the Fucino Plain (Italy). J. Environ. Monit. : JEM 10 (6), 747–752. https://doi. Tinwell, H., Ashby, J., 2004. Sensitivity of the immature rat uterotrophic assay to
org/10.1039/b800559a. mixtures of estrogens. Environ. Health Perspect. 112, 575–582.
Salis, S., Testa, C., Roncada, P., Armorini, S., Rubattu, N., Ferrari, A., et al., 2017. Toan, P.V., Sebesvari, Z., Bläsing, M., Rosendahl, I., Renaud, F.G., 2013. Pesticide
Occurrence of imidacloprid, carbendazim, and other biocides in Italian house dust: management and their residues in sediments and surface and drinking water in the
potential relevance for intakes in children and pets. J. Environ. Sci. Health B 52, Mekong Delta, Vietnam. Sci. Total Environ. 452–453, 28–39. https://doi.org/
699–709. 10.1016/j.scitotenv.2013.02.026.
Sangami, S., Manu, B., 2017. Fenton’s treatment of actual agriculture runoff water Tosi, S., Costa, C., Vesco, U., Quaglia, G., Guido, G., 2018. A 3-year survey of Italian
containing herbicides. Water Sci. Technol. : a journal of the International Association honey bee-collected pollen reveals widespread contamination by agricultural
on Water Pollution Research 75 (2), 451–461. https://doi.org/10.2166/ pesticides. Sci. Total Environ. 615, 208–218. https://doi.org/10.1016/j.
wst.2016.538. scitotenv.2017.09.226.
Sayles, G., You, G., Wang, M., Kupferle, M., 1997. DDT, DDD, and DDE dechlorination by Tröger, R., Ren, H., Yin, D., Postigo, C., Nguyen, P.D., Baduel, C., Golovko, O., Been, F.,
zero-valent iron. Environ. Sci. Technol. 31, 3448–3454. ES9701669. Joerss, H., Boleda, M.R., Polesello, S., Roncoroni, M., Taniyasu, S., Menger, F.,
Segura, C., Zaror, C., Mansilla, H.D., Mondaca, M.A., 2008. Imidacloprid oxidation by Ahrens, L., Yin Lai, F., Wiberg, K., 2021. What’s in the water? - target and suspect
photo-Fenton reaction. J. Hazard Mater. 150 (3), 679–686. https://doi.org/ screening of contaminants of emerging concern in raw water and drinking water
10.1016/j.jhazmat.2007.05.018. from Europe and Asia. Water Res. 198, 117099. https://doi.org/10.1016/j.
Shah, N.S., He, X., Khan, H.M., Khan, J.A., O’Shea, K.E., Boccelli, D.L., Dionysiou, D.D., watres.2021.117099.
2013. Efficient removal of endosulfan from aqueous solution by UV-C/peroxides: a Turabik, M., Oturan, N., Gözmen, B., Oturan, M.A., 2014. Efficient removal of insecticide
comparative study. J. Hazard Mater. 263 (Pt 2), 584–592. https://doi.org/10.1016/ "imidacloprid" from water by electrochemical advanced oxidation processes.
j.jhazmat.2013.10.019. Environ. Sci. Pollut. Res. Int. 21 (14), 8387–8397. https://doi.org/10.1007/s11356-
Sharma, R.K., Kumar, A., Joseph, P.E., 2008. Removal of atrazine from water by low cost 014-2788-9.
adsorbents derived from agricultural and industrial wastes. Bull. Environ. Contam. Turgut, C., Gokbulut, C., Cutright, T.J., 2009. Contents and sources of DDT impurities in
Toxicol. 80 (5), 461–464. https://doi.org/10.1007/s00128-008-9389-6. dicofol formulations in Turkey. Environ. Sci. Pollut. Res. Int. 16 (2), 214–217.
Shayeghi, M., Khoobdel, M., Vatandoost, H., 2007. Determination of organophosphorus https://doi.org/10.1007/s11356-008-0083-3.
insecticides (malathion and diazinon) residue in the drinking water. Pakistan J. Biol. Ukalska-Jaruga, A., Lewińska, K., Mammadov, E., Karczewska, A., Smreczak, B.,
Sci. : PJBS 10 (17), 2900–2904. https://doi.org/10.3923/pjbs.2007.2900.2904. Medyńska-Juraszek, A., 2020a. Residues of persistent organic pollutants (POPs) in
Shea, P.J., Machacek, T.A., Comfort, S.D., 2004. Accelerated remediation of pesticide- agricultural soils adjacent to historical sources of their storage and distribution-the
contaminated soil with zerovalent iron. Environ. Pollut. 132 (2), 183–188. https:// case study of Azerbaijan. Molecules 25 (8), 1815. https://doi.org/10.3390/
doi.org/10.1016/j.envpol.2004.05.003. molecules25081815.
Sherrard, R.M., Bearr, J.S., Murray-Gulde, C.L., Rodgers Jr., J.H., Shah, Y.T., 2004. Ukalska-Jaruga, A., Smreczak, B., Siebielec, G., 2020b. Assessment of pesticide residue
Feasibility of constructed wetlands for removing chlorothalonil and chlorpyrifos content in polish agricultural soils. Molecules 25 (3), 587. https://doi.org/10.3390/
from aqueous mixtures. Environ. Pollut. 127 (3), 385–394. https://doi.org/10.1016/ molecules25030587.
j.envpol.2003.08.017. Umulisa, V., Kalisa, D., Skutlarek, D., Reichert, B., 2020. First evaluation of DDT
Shi, L., Jiang, Y., Wan, Y., Huang, J., Meng, Q., He, Z., Xu, S., Xia, W., 2020. Occurrence (dichlorodiphenyltrichloroethane) residues and other Persistence Organic Pollutants
of the insecticide fipronil and its degradates in indoor dust from South, Central, and

29
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

in soils of Rwanda: nyabarongo urban versus rural wetlands. Ecotoxicol. Environ. Total Environ. 470–471, 1047–1055. https://doi.org/10.1016/j.
Saf. 197, 110574. https://doi.org/10.1016/j.ecoenv.2020.110574. scitotenv.2013.10.056.
US EPA, 2000. Supplementary guidance for conducting health risk assessment of Wu, C., Shemer, H., Linden, K.G., 2007. Photodegradation of metolachlor applying UV
chemical mixtures. Risk Assessment Forum Technical Panel. In: United States and UV/H2O2. J. Agric. Food Chem. 55 (10), 4059–4065. https://doi.org/10.1021/
Environmental Protection Agency Office of EPA/630/R-00/002. jf0635762.
van der Kooij, D. (Ed.), 2014. Microbial Growth in Drinking-Water Supplies: Problems, Wu, H., Shen, J., Jiang, X., Liu, X., Sun, X., Li, J., Han, W., Mu, Y., Wang, L., 2018.
Causes, Control and Research Needs; [International BTO-Conference “Microbial Bioaugmentation potential of a newly isolated strain Sphingomonas sp. NJUST37 for
Growth in Drinking Water Distribution Systems and Tap Water Installations” on 11 the treatment of wastewater containing highly toxic and recalcitrant tricyclazole.
and 12 May 2011 at Santpoort in the Netherlands]. IWA Publ, London. Bioresour. Technol. 264, 98–105. https://doi.org/10.1016/j.biortech.2018.05.071.
Van Maele-Fabry, G., Libotte, V., Willems, J., Lison, D., 2006. Review and meta-analysis Wu, P., Chen, Z., Zhang, Y., Wang, Y., Zhu, F., Cao, B., Jin, L., Hou, Y., Wu, Y., Li, N.,
of risk estimates for prostate cancer in pesticide manufacturing workers. Cancer 2019a. Carbaryl waste-water treatment by Rhodopseudomonas sphaeroides.
Causes Control : CCC (Cancer Causes Control) 17 (4), 353–373. https://doi.org/ Chemosphere 233, 597–602. https://doi.org/10.1016/j.chemosphere.2019.05.237
10.1007/s10552-005-0443-y. (Retraction published Chemosphere. 2020 Oct;257:127394).
Vanraes, P., Wardenier, N., Surmont, P., Lynen, F., Nikiforov, A., Van Hulle, S., Leys, C., Wu, P., Xie, L., Mo, W., Wang, B., Ge, H., Sun, X., Tian, Y., Zhao, R., Zhu, F., Zhang, Y.,
Bogaerts, A., 2018. Removal of alachlor, diuron and isoproturon in water in a falling Wang, Y., 2019b. The biodegradation of carbaryl in soil with Rhodopseudomonas
film dielectric barrier discharge (DBD) reactor combined with adsorption on capsulata in wastewater treatment effluent. J. Environ. Manag. 249, 109226.
activated carbon textile: reaction mechanisms and oxidation by-products. J. Hazard https://doi.org/10.1016/j.jenvman.2019.06.127.
Mater. 354, 180–190. https://doi.org/10.1016/j.jhazmat.2018.05.007. Wu, X., Chen, A., Yuan, Z., Kang, H., Xie, Z., 2020. Atmospheric organochlorine
Varca, L.M., 2012. Pesticide residues in surface waters of Pagsanjan-Lumban catchment pesticides (OCPs) and polychlorinated biphenyls (PCBs) in the Antarctic marginal
of Laguna de Bay, Philippines. Agric. Water Manag. 106, 35–41. https://doi.org/ seas: distribution, sources and transportation. Chemosphere 258, 127359. https://
10.1016/j.agwat.2011.08.006. doi.org/10.1016/j.chemosphere.2020.127359.
Varloud, M., Fourie, J.J., Blagburn, B.L., Deflandre, A., 2015. Expellency, anti-feeding Xia, H., Ma, X., 2006. Phytoremediation of ethion by water hyacinth (Eichhornia
and speed of kill of a dinotefuran-permethrin-pyriproxyfen spot-on (Vectra®3D) in crassipes) from water. Bioresour. Technol. 97 (8), 1050–1054. https://doi.org/
dogs weekly challenged with adult fleas (Ctenocephalides felis) for 1 month- 10.1016/j.biortech.2005.04.039.
comparison to a spinosad tablet (Comfortis®). Parasitol. Res. 114 (7), 2649–2657. Xie, J., Wang, P., Liu, J., Lv, X., Jiang, D., Sun, C., 2011. Photodegradation of lambda-
https://doi.org/10.1007/s00436-015-4470-7. cyhalothrin and cypermethrin in aqueous solution as affected by humic acid and/or
Vogel, J.R., Majewski, M.S., Capel, P.D., 2008. Pesticides in rain in four agricultural copper: intermediates and degradation pathways. Environ. Toxicol. Chem. 30 (11),
watersheds in the United States. J. Environ. Qual. 37 (3), 1101–1115. https://doi. 2440–2448. https://doi.org/10.1002/etc.655.
org/10.2134/jeq2007.0079. Xu, L., Granger, C., Dong, H., Mao, Y., Duan, S., Li, J., Qiang, Z., 2020. Occurrences of 29
Vymazal, J., Březinová, T., 2015. The use of constructed wetlands for removal of pesticides in the Huangpu River, China: highest ecological risk identified in Shanghai
pesticides from agricultural runoff and drainage: a review. Environ. Int. 75, 11–20. metropolitan area. Chemosphere 251, 126411. https://doi.org/10.1016/j.
https://doi.org/10.1016/j.envint.2014.10.026. chemosphere.2020.126411.
Wahaab, R.A., Badawy, M.I., 2004. Water quality assessment of the River Nile system: an Yabalak, E., Gizir, A.M., 2020. Treatment of agrochemical wastewater by subcritical
overview. Biomed. Environ. Sci. : BES (Biomed. Environ. Sci.) 17 (1), 87–100. water oxidation method: chemical composition and ion analysis of treated and
Wan, Y., Wang, Y., Xia, W., He, Z., Xu, S., 2019. Neonicotinoids in raw, finished, and tap untreated samples. J. Environ. Sci. Health - Part A Toxic/Hazard. Subst. Environ.
water from Wuhan, Central China: assessment of human exposure potential. Sci. Eng. 55 (12), 1424–1435. https://doi.org/10.1080/10934529.2020.1805249.
Total Environ. 675, 513–519. https://doi.org/10.1016/j.scitotenv.2019.04.267. Yadav, I.C., Devi, N.L., Syed, J.H., Cheng, Z., Li, J., Zhang, G., Jones, K.C., 2015. Current
Wan, Y., Tran, T.M., Nguyen, V.T., Wang, A., Wang, J., Kannan, K., 2021. status of persistent organic pesticides residues in air, water, and soil, and their
Neonicotinoids, fipronil, chlorpyrifos, carbendazim, chlorotriazines, chlorophenoxy possible effect on neighboring countries: a comprehensive review of India. Sci. Total
herbicides, bentazon, and selected pesticide transformation products in surface Environ. 511, 123–137. https://doi.org/10.1016/j.scitotenv.2014.12.041.
water and drinking water from northern Vietnam. Sci. Total Environ. 750, 141507. Yang, F., Gao, Y., Sun, L., Zhang, S., Li, J., Zhang, Y., 2018. Effective sorption of atrazine
https://doi.org/10.1016/j.scitotenv.2020.141507. by biochar colloids and residues derived from different pyrolysis temperatures.
Wandan, E.N., Zabik, M.J., 1996. Assessment of the contamination of surface water and Environ. Sci. Pollut. Res. Int. 25 (19), 18528–18539. https://doi.org/10.1007/
fish from Côte d’Ivoire. J. Environ. Sci. Health - Part B Pesticides, Food Contam. s11356-018-2077-0.
Agric. Wastes 31 (2), 225–240. https://doi.org/10.1080/03601239609372984. Yang, K.J., Lee, J., Park, H.L., 2020. Organophosphate pesticide exposure and breast
Wang, G., Han, Y., Li, X., Andaloro, J., Chen, P., Hoffmann, W.C., Han, X., Chen, S., cancer risk: a rapid review of human, animal, and cell-based studies. Int. J. Environ.
Lan, Y., 2020. Field evaluation of spray drift and environmental impact using an Res. Publ. Health 17 (14), 5030. https://doi.org/10.3390/ijerph17145030.
agricultural unmanned aerial vehicle (UAV) sprayer. Sci. Total Environ. 737, Yang, L., Zhang, Z., 2019. Degradation of six typical pesticides in water by VUV/UV/
139793. https://doi.org/10.1016/j.scitotenv.2020.139793. chlorine process: evaluation of the synergistic effect. Water Res. 161, 439–447.
Wardenier, N., Vanraes, P., Nikiforov, A., Van Hulle, S., Leys, C., 2019. Removal of https://doi.org/10.1016/j.watres.2019.06.021.
micropollutants from water in a continuous-flow electrical discharge reactor. Yang, X., Wang, S., Bian, Y., Chen, F., Yu, G., Gu, C., Jiang, X., 2008. Dicofol application
J. Hazard Mater. 362, 238–245. https://doi.org/10.1016/j.jhazmat.2018.08.095. resulted in high DDTs residue in cotton fields from northern Jiangsu province, China.
Weber, J., Kurková, R., Klánová, J., Klán, P., Halsall, C.J., 2009. Photolytic degradation J. Hazard Mater. 150 (1), 92–98. https://doi.org/10.1016/j.jhazmat.2007.04.076.
of methyl-parathion and fenitrothion in ice and water: implications for cold Yao, F., Yu, G., Bian, Y., Yang, X., Wang, F., Jiang, X., 2007. Bioavailability to grains of
environments. Environ. Pollut. 157 (12), 3308–3313. https://doi.org/10.1016/j. rice of aged and fresh DDD and DDE in soils. Chemosphere 68 (1), 78–84. https://
envpol.2009.05.045. doi.org/10.1016/j.chemosphere.2006.12.035.
Westlund, P., Isazadeh, S., Therrien, A., Yargeau, V., 2018. Endocrine activities of Yoon, Y., Westerhoff, P., Snyder, S.A., Wert, E.C., Yoon, J., 2007. Removal of endocrine
pesticides during ozonation of waters. Bull. Environ. Contam. Toxicol. 100 (1), disrupting compounds and pharmaceuticals by nanofiltration and ultrafiltration
112–119. https://doi.org/10.1007/s00128-017-2254-8. membranes. Desalination 202 (1_3), 16_23. https://doi.org/10.1016/j.
Whelan, M.J., Ramos, A., Villa, R., Guymer, I., Jefferson, B., Rayner, M., 2020. A new desal.2005.12.033. Available from:
conceptual model of pesticide transfers from agricultural land to surface waters with Yu, J.-J., 2002. Removal of organophosphate pesticides from wastewater by supercritical
a specific focus on metaldehyde. Environmental science. Processes & impacts 22 (4), carbon dioxide extraction. Water Res. 36 (4), 1095–1101. https://doi.org/10.1016/
956–972. https://doi.org/10.1039/c9em00492k. s0043-1354(01)00293-7.
WHO, World Health Organization, 2009. The WHO Recommended Classification of Zahn, L.K., Cox, D.L., Gerry, A.C., 2019. Mortality rate of house flies (Diptera: muscidae)
Pesticides by Hazard and Guidelines to Classification: 2009. ISBN 978 92 4 154796 exposed to insecticidal granular fly baits containing indoxacarb, dinotefuran, or
3; ISSN 1684-1042. cyantraniliprole. J. Econ. Entomol. 112 (5), 2474–2481. https://doi.org/10.1093/
WHO World Health Organization, 2017. Guidelines for Drinking-Water Quality: Fourth jee/toz170.
Edition Incorporating the First Addendum. Geneva: Licence: CC BY-NC-SA 3.0 IGO. Zhang, H., Dorr, G.J., Hewitt, A.J., 2015. Retention and efficacy of ultra-low volume
http://apps.who.int/iris. pesticide applications on Culex quinquefasciatus (Diptera: Culicidae). Environ. Sci.
Willemin, M.E., Kadar, A., de Sousa, G., Leclerc, E., Rahmani, R., Brochot, C., 2015. In Pollut. Res. Int. 22 (21), 16492–16501. https://doi.org/10.1007/s11356-015-5480-
vitro human metabolism of permethrin isomers alone or as a mixture and the 9.
formation of the major metabolites in cryopreserved primary hepatocytes. Toxicol. Zhang, W., Lin, Z., Pang, S., Bhatt, P., Chen, S., 2020. Insights into the biodegradation of
Vitro : an international journal published in association with BIBRA 29 (4), 803–812. lindane (γ-Hexachlorocyclohexane) using a microbial system. Front. Microbiol. 11,
https://doi.org/10.1016/j.tiv.2015.03.003. 522. https://doi.org/10.3389/fmicb.2020.00522.
Willis, G.H., Smith, S., McDowell, L.L., Southwick, L.M., 1996. Carbaryl washoff from Zhang, Y., Xu, Z., Chen, Z., Wang, G., 2020. Simultaneous degradation of triazophos,
soybean plants. Arch. Environ. Contam. Toxicol. 31 (2), 239–243. https://doi.org/ methamidophos and carbofuran pesticides in wastewater using an Enterobacter
10.1007/BF00212372. bacterial bioreactor and analysis of toxicity and biosafety. Chemosphere 261,
Wilson, P.C., Lu, H., Lin, Y., 2011. Norflurazon and simazine removal from surface water 128054. https://doi.org/10.1016/j.chemosphere.2020.128054.
using a constructed wetland. Bull. Environ. Contam. Toxicol. 87 (4), 426–430. Zhi, S., Shen, S., Zhou, J., Ding, G., Zhang, K., 2020. Systematic analysis of occurrence,
https://doi.org/10.1007/s00128-011-0380-2. density and ecological risks of 45 veterinary antibiotics: focused on family livestock
Woudneh, M.B., Ou, Z., Sekela, M., Tuominen, T., Gledhill, M., 2009. Pesticide farms in Erhai Lake basin, Yunnan, China. Environ. Pollut. 267, 115539. https://doi.
multiresidues in waters of the lower fraser valley, British columbia, Canada. Part II. org/10.1016/j.envpol.2020.115539.
Groundwater. J. Environ. Qual. 38 (3), 948–954. https://doi.org/10.2134/
jeq2007.0523.
Wu, C., Luo, Y., Gui, T., Huang, Y., 2014. Concentrations and potential health hazards of
organochlorine pesticides in (shallow) groundwater of Taihu Lake region, China. Sci.

30
I. El-Nahhal and Y. El-Nahhal Journal of Environmental Management 299 (2021) 113611

Zhou, R., Zhu, L., Chen, Y., 2008. Levels and source of organochlorine pesticides in in urban and rural areas of China. Environ. Int. 142, 105822. https://doi.org/
surface waters of Qiantang River, China. Environ. Monit. Assess. 136 (1–3), 10.1016/j.envint.2020.105822.
277–287. https://doi.org/10.1007/s10661-007-9683-5. Zhu, C., Fang, G., Dionysiou, D.D., Liu, C., Gao, J., Qin, W., Zhou, D., 2016. Efficient
Zhou, Y., Guo, J., Wang, Z., Zhang, B., Sun, Z., Yun, X., Zhang, J., 2020. Levels and transformation of DDTs with persulfate activation by zero-valent iron nanoparticles:
inhalation health risk of neonicotinoid insecticides in fine particulate matter (PM2.5) a mechanistic study. J. Hazard Mater. 316, 232–241. https://doi.org/10.1016/j.
jhazmat.2016.05.040.

31

You might also like