Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Science of the Total Environment 690 (2019) 853–866

Contents lists available at ScienceDirect

Science of the Total Environment

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

Review

1,4-Dioxane as an emerging water contaminant: State of the science and


evaluation of research needs
Krystal J. Godri Pollitt a,⁎, Jae-Hong Kim b, Jordan Peccia b, Menachem Elimelech b, Yawei Zhang a,c,
Georgia Charkoftaki a, Brenna Hodges b, Ines Zucker b, Huang Huang a, Nicole C. Deziel a, Kara Murphy d,
Momoko Ishii a, Caroline H. Johnson a, Andrea Boissevain e, Elaine O'Keefe f, Paul T. Anastas a,g, David Orlicky h,
David C. Thompson i, Vasilis Vasiliou a,⁎
a
Department of Environmental Health Sciences, School of Public Health, Yale University, New Haven, CT 06510, United States
b
Department of Chemical & Environmental Engineering, School of Engineering & Applied Science, Yale University, New Haven, CT 06520, United States
c
Department of Surgery, School of Medicine, Yale University, New Haven, CT 06520, United States
d
Northeast States for Coordinated Air Use Management (NESCAUM), Boston, MA 02111, United States
e
Stratford Health Department, Stratford, CT 06615, United States
f
Office of Public Health Practice, School of Public Health, Yale University, New Haven, CT 06510, United States
g
Center for Green Chemistry and Green Engineering, Department of Chemistry, Yale School of Forestry & Environmental Studies, New Haven, CT 06511, United States
h
Department of Pathology, University of Colorado School of Medicine, Aurora, CO 80045, United States
i
Department of Clinical Pharmacy, University of Colorado School of Pharmacy, Aurora, CO 80045, United States

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• 1,4-Dioxane is an emerging environ-


mental contaminant and a probable car-
cinogen.
• The de minimis cancer risk level is High levels

exceeded at 7% of U.S. drinking water


sites.
• The physicochemical properties chal-
lenge detection and remediation of 1,4-
dioxane.
• This review presents data needed to set
an enforceable drinking water standard.

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

Article history: 1,4-Dioxane has historically been used to stabilize chlorinated solvents and more recently has been found as a
Received 21 May 2019 contaminant of numerous consumer and food products. Once discharged into the environment, its physical
Received in revised form 26 June 2019 and chemical characteristics facilitate migration in groundwater, resulting in widespread contamination of drink-
Accepted 26 June 2019
ing water supplies. Over one-fifth of U.S. public drinking water supplies contain detectable levels of 1,4-dioxane.
Available online 28 June 2019
Remediation efforts using common adsorption and membrane filtration techniques have been ineffective,

Abbreviations: 1,1-DCA, 1,1-dichloroethane; 1,1,1-TCA, 1,1,1-trichloroethane; ATSDR, Agency for Toxic Substances and Disease Registry; EPA, Environmental Protection Agency; GC,
gas chromatography; HEAA, β-hydroxyethoxyacetic acid; MS, mass spectrometry; TCE, trichloroethylene; UCMR, unregulated contaminant monitoring rule; U.S., United States.
⁎ Corresponding authors.
E-mail addresses: krystal.pollitt@yale.edu (K.J. Godri Pollitt), vasilis.vasiliou@yale.edu (V. Vasiliou).

https://doi.org/10.1016/j.scitotenv.2019.06.443
0048-9697/© 2019 Elsevier B.V. All rights reserved.
854 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

highlighting the need for alternative removal approaches. While the data evaluating human exposure and health
Editor: Damia Barcelo effects are limited, animal studies have shown chronic exposure to cause carcinogenic responses in the liver
across multiple species and routes of exposure. Based on this experimental evidence, the U.S. Environmental Pro-
Keywords: tection Agency has listed 1,4-dioxane as a high priority chemical and classified it as a probable human carcinogen.
1,4-dioxane
Despite these health concerns, there are no federal or state maximum contaminant levels for 1,4-dioxane. Effec-
Drinking water
Health
tive public health policy for this emerging contaminant requires additional information about human health ef-
Exposure fects, chemical interactions, environmental fate, analytical detection, and treatment technologies. This review
Sensors highlights the current state of knowledge, key uncertainties, and data needs for future research on 1,4-dioxane.
Remediation © 2019 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 854
2. Environmental behavior and persistence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 855
2.1. Environmental occurrence. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 855
2.2. Co-contaminants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 855
2.3. Environmental fate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 855
3. Routes of exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
3.1. Drinking water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
3.2. Food . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
3.3. Consumer products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
3.4. Bathing and showering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
4. Toxicology and carcinogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 857
4.1. Toxicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 857
4.2. Metabolism and elimination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 857
4.3. Carcinogenic effects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 858
4.4. Mechanisms of carcinogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 858
4.5. Cancer potency factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 859
5. Exposure assessment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 859
5.1. Environmental samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 859
5.2. Biological samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 860
6. Health effects in humans. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861
7. Remediation methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861
7.1. Abiotic advanced oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861
7.2. Bioremediation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 862
8. Implications for further studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
9. Suggestions for community engagement and preparedness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
Funding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864

1. Introduction and Disease Registry (ATSDR) listed 1,4-dioxane as one of the


Superfund priority substances (214th among 275 compounds)
1,4-Dioxane (C4H8O2; CAS No. 123–91-1) is a synthetic chemical based upon its potential toxic effects and propensity to contami-
that was historically used as a stabilizer in chlorinated solvents in in- nate drinking water (ATSDR 2017). The U.S. Environmental Protec-
dustrial processes, such as 1,1,1-trichloroethane (1,1,1-TCA). More tion Agency's (EPA) unregulated contaminant monitoring rule
recently, it has been used as a solvent in the production of a variety (UCMR) also prioritized 1,4-dioxane for inclusion in the
of organic chemicals found in paints, lacquers, dyes, antifreeze, de- 2013–2015 test round. In UCMR, “contaminant of emerging con-
odorants, shampoos, and cosmetics (ATSDR 2012; EPA, 2006). 1,4- cern” is designated to compounds of which (i) the risk to human
Dioxane has also been used as a food additive and in the formulation health and the environment is not fully understood, and (ii) drink-
of pesticides and food packaging adhesives (ATSDR 2012; DHHS ing water standards have not been developed. These emerging con-
2014). In the United States (U.S.), environmental releases of 1,4-di- taminants represent key research needs especially when there is
oxane are regulated as hazardous waste only when this chemical is evidence of widespread exposure and potentially serious adverse
used as an industrial solvent. When used in other applications (i.e., effects. While substantial effort has been placed by the U.S. EPA
additive, pesticide, adhesive, etc.), its disposal is not regulated on certain emerging drinking water contaminants (e.g.,
(EPA, 2011; EPA, 2017b). Such limited regulation leads to consider- perfluorinated alkyl substances), relatively little regulatory atten-
able environmental release, e.g., approximately 617,000 pounds in tion has been given to 1,4-dioxane. As a result, drinking water guid-
2017 from solvent use alone (EPA, 2017a; EPA, 2018b). The high mo- ance values established in recent years by jurisdictions range
bility and persistence of 1,4-dioxane enables it to distribute widely widely from 0.35 to 77 μg/L due to uncertainties in the potability
in the environment. of water affected by this contaminant. This situation is further com-
Significant improvements have been made in recent decades in plicated by the difficulty of removing 1,4-dioxane from drinking
controlling microbial and chemical threats to water supplies. Fed- water by conventional treatment technologies.
eral agencies in the U.S. have recognized the threats to drinking Given these considerations, this review highlights the current state
water posed by emerging contaminants, such as 1,4-dioxane, at of knowledge, key uncertainties and data needs for future research on
Superfund sites. For example, the Agency for Toxic Substances 1,4-dioxane.
K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866 855

2. Environmental behavior and persistence dioxane is seldom found independent of chlorinated solvent contami-
nants. A high correlation between the occurrence of 1,4-dioxane and
The chemical and physical properties that facilitate the mobility of TCE was also found in studies performed in other countries such as
this contaminant through groundwater to drinking water supplies will Germany (Karges et al. 2018). These findings highlight the need to fur-
be discussed. ther study the biological effects of mixtures of 1,4-dioxane and chlori-
nated solvents, and suggest that the dominant source of 1,4 dioxane in
2.1. Environmental occurrence the environment is from its use with chlorinated solvents.

The occurrence of 1,4-dioxane in the environment has been associ- 2.3. Environmental fate
ated with contaminated groundwater in industrial sites (Adamson
et al. 2016; Lippincott et al. 2015; Postigo and Barcelo 2015). In the 1,4-Dioxane can be released into the air, water or soil. It is in general
U.S., 1,4-Dioxane is listed as number 214 on the ATSDR's Substance Pri- not a concern in the atmosphere, since it is non-volatile with very low
ority List and was found in According to U.S. EPA's UCMR-3 testing, 1,4- Henry's Law constant (4.8 × 10−6 atm·m3/mol at 25 °C) (O'Neil 2013;
dioxane was found to be the second most prevalent prior contaminant Verschueren 2008). Half-life is also relatively short at 35 h due to gas-
in public water supplies, detected at 21% of the 4864 tested sites across phase photo-oxidation (Mohr 2001). In water, 1,4-dioxane can occur
the U.S. (Fig. 1A) (Adamson et al. 2015; EPA, 2015). Exceedances of the at high concentration due to high solubility (100 mg/mL) (NTP 1992).
federal health risk guideline level for 1,4-dioxane (based upon the can- Low octanol-water partitioning coefficient (log Kow = −0.27) (ATSDR
cer de minimis 1 in a million-reference value of 0.35 μg/L) were found at 2012) and weak adsorption into mineral phases and organic matter
6.9% of these tested sites (See Section 4.5: Cancer Potency Factor). Most (Jackson and Dwarakanath 1999; Lesage et al. 1990; Priddle and
of these sites were clustered in specific geographic areas, including the Jackson 1991) allow 1,4-dioxane to efficiently leach into water systems
north Atlantic coast, the upper Midwest, and California (Adamson and transport through even subsurface environment (Zhang et al.
et al. 2017). Long Island NY is one example of a region impacted by ele- 1990); limited mobility has only been observed in regions with clay
vated 1,4-dioxane concentration; approximately 66% of its public water soils. Coupled with its chemical inertness, 1,4-dioxane is extremely per-
supplies exceed the federal guideline level (Fig. 1B). This contaminant sistent in groundwater with half-life of 2–5 years, while moderately
has also been increasingly detected at a growing number of locations in- persistent in surface water with half-life of 56 days (Adamson et al.
ternationally, including Japan (Abe 1999; Tanabe et al. 2006), Korea (An 2015; EPA, 2018a). Consequently, point-source releases of 1,4-dioxane
et al. 2014), Canada (Health Canada 2018), the United Kingdom, and can cause plumes which have the potential to contaminate both near-
Germany (Karges et al. 2018; Stepien et al. 2014; Stepien and field and distal water-supply wells. In California, for example, where
Püttmann 2013). N2000 sites contain elevated concentrations of 1,4-dioxane, N50% of
plumes exceeded 270 m in length and N10% were longer than 1000 m
2.2. Co-contaminants (Adamson et al. 2014).\
A shorter lifetime of 1,4-dioxane in surface water compared to sub-
1,4-Dioxane is commonly found in affected groundwater as a mix- surface water is likely related to more prevalent microbial degradation
ture with other chlorinated solvents, including 1,1,1-TCA, 1,1-dichloro- in the former environment. Through monooxygenase-metabolic or co-
ethane (1,1-DCA), and trichloroethylene (TCE), all of which are also metabolic pathways, it can be progressively transformed to β-
listed in the ATSDR's Substance Priority List. Together with 1,4-dioxane, hydroxyethoxyacetic acid (HEAA), ethylene glycol, glycolate, and oxa-
these contaminants have been detected at 1699 current or former U.S. late (Huang et al. 2014; Mahendra et al. 2007; Zhang et al. 2017). Min-
EPA Superfund National Priority List sites. While 1,4-dioxane was de- eralization of these intermediates to carbon dioxide and water is
tected at only 31 of these sites, testing for this contaminant only became achieved by select bacterial cultures (Grostern et al. 2012; Lan et al.
commonplace in the past 15 years. As such, its likely occurrence at some 2013; Mahendra et al. 2007). Aerobic metabolism of 1,4-dioxane has
waste sites might have not been documented (ATSDR 2016) (ATSDR been demonstrated by several bacterial and fungal strains, including
2006) (ATSDR 2015). Based on the U.S. Air Force Environmental Resto- Pseudonocardia dioxanivorans CB1190 (Mahendra and Alvarez-Cohen
ration Program Information Management System, Anderson et al. 2005; Parales et al. 1994), Mycobacterium sp. PH-06 (Kim et al. 2009),
(2012) reported that 1,4-dioxane was present in 17.4% of the ground- Rhodococcus 219 (Bernhardt and Diekmann 1991), and Cordyceps
water monitoring wells; TCE and/or 1,1,1-TCA were detected in 93.7% sinensis (Nakamiya et al. 2005). Through this metabolic transformation
of all samples containing 1,4-dioxane, confirming the notion that 1,4- process, bacteria and fungi use 1,4-dioxane as a sole energy and carbon

Fig. 1. 1,4-Dioxane concentrations in public water supplies (surface and groundwater) by zip code across the U.S. (A) and Long Island, NY (B) (adapted from U.S. EPA, 2015). MRL: Minimum
reportable level.
856 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

source. Co-metabolism of 1,4-dioxane occurs in numerous bacterial Table 1


strains when grown on a carbon substrate, typically a C2 to C5 alkane Drinking water guidelines and criteria for 1,4-dioxane (An et al. 2014; EPA US, 2017b; EPA
US, 2018a, b; Health Canada 2018; Mulisch et al. 2003; WHO 2005; Yamamoto et al. 2018).
(Mahendra and Alvarez-Cohen 2006; Vainberg et al. 2006). These mi-
croorganisms include Mycobacterium austroafricanum JOB5 (Lan et al. Jurisdiction Target Type of target Year target
2013), Pseudonocardia sp. strain ENV478 (Vainberg et al. 2006), concentration was
(μg/L) introduced
Rhodococcus ruber ENV425 (Lippincott et al. 2015) and Pseudonocardia
K1 (Mahendra and Alvarez-Cohen 2006). Application of biodegradation Alaska 77 Groundwater cleanup level 2016
California 1.0 Public health protective 2011
as a potential 1,4-dioxdane remediation strategy will be further
concentration
discussed in Section 7.2: Bioremediation. Connecticut 3.0 Action level 2013
Reactive transport models have been instrumental in evaluating the Indiana 4.6 Groundwater screening level 2016
fate and transport of contaminants in subsurface. These computational Maine 4.0 Guideline 2016
models predict advective-dispersive transport and reactions that trans- Massachusetts 0.3 Guideline 2004
Michigan 7.2 Drinking water criterion 2017
form and immobilize contaminants over large temporal and spatial do- Minnesota 1.0 Health risk limit 2013
mains (Anderson et al. 2015; Benjakul 2010). Despite frequent and New Jersey 0.4 Groundwater quality criterion 2015
successful use in simulating the fate of other subsurface contaminants New York 1.0 Draft minimum criterion level 2018
including halogenated solvents, it has not yet been used for 1,4-dioxane North Carolina 3.0 Drinking water criterion 2015
Texas 9.1 Protective concentration level 2016
in groundwater due to the dearth of data required to accurately quantify
U.S. 0.35 U.S EPA, screening level for tap 2017
mass transfer and biotic/abiotic transformation in complex environ- watera
mental matrix. For example, the impact of co-contaminants and their Germany 0.1 Federal environmental agency, 2003
degradation byproducts on the potential inhibition of 1,4-dioxane re- suggested guideline
moval (Deng et al. 2018) needs to be carefully evaluated under the var- Canada 50 Health Canada, draft drinking 2018
water guideline
ious physicochemical conditions likely to be encountered in subsurface
Republic of 50 Ministry of the environment, 2014
environment. Careful validation of the model against existing field data Korea provisional standard
on 1,4-dioxane migration also needs to be performed. Development of Japan 50 Water pollution control law, 2009
these transport models will facilitate identification of drinking-water drinking water standard
World Health 50 Guideline 2005
resources vulnerable to 1,4-dioxane releases, predication of exposure
Organization
routes, and proper allocation of remedial actions.
a
U.S. EPA has not set an enforceable Minimum Criterion Level.

3. Routes of exposure
used to treat food crops, and manufactured food additives (e.g., polysor-
Exposure to 1,4-dioxane occurs through ingestion of contaminated bate 60, 65, and 80, and polyethylene glycol) (ATSDR 2012). According
water and food, dermal contact or inhalation of vapors (ATSDR 2012; to the food-exposure study by Health Canada and Environment Canada
DHHS 2014; EPA, 2010). These environmental exposure pathways to (2010), formula-fed infants (b0.6 months of age) were reported to ex-
1,4-dioxane will be discussed. perience the highest exposure to this contaminant (1.1 μg/kg body
weight/day) followed by children aged 1 through 4 years (0.3 μg/kg
3.1. Drinking water body weight day). In the U.S., 1,4-dioxane content in food additives is
regulated to not exceed 10 mg per kg of additive
Ingestion of 1,4-dioxane through drinking water is the dominant (PharmacopeialConvention, 2008).
pathway of exposure (Abe 1999; Anderson et al. 2012; Chiang et al.
2008; Karges et al. 2018; Stepien et al. 2014; Sun et al. 2016). As 3.3. Consumer products
discussed in Section 2.1: Environmental Occurrence, approximately 7%
of public drinking water supplies in the U.S. exceed the federal health Personal care products containing ethoxylated surfactants are also a
risk guideline level of 0.35 μg/L (EPA, 2013). According to a recent anal- potential source of 1,4-dioxane (Health Canada 2018). Elevated concen-
ysis by a non-government organization, when extrapolated across the trations have been reported for various cosmetic products (279 mg/kg)
U.S., these drinking water levels would expose millions of public and children's shampoo (85 mg/kg) in the U.S. (Black et al. 2001). As-
water consumers to elevated cancer risk from 1,4-dioxane contamina- sessments of consumer products available in Japan and Taiwan revealed
tion (Environmental Working Group (EWG) 2017). The World Health high levels of 1,4-dioxane in shampoos (45.5 mg/kg), body washes
Organization as well as federal agencies in Japan, Korea, and Canada (15.7 mg/kg), hand soaps (7.5 mg/kg), and hair conditioners
have all recommended a drinking water guidance level of 50 μg/L (An (0.14 mg/kg) (Fuh et al. 2005; Makino et al. 2006; Tahara et al. 2013;
et al. 2014; EPA, 2018a; Health Canada 2018; WHO 2005; WHO 2011; Tanabe and Kawata 2008). In contrast, concentrations in laundry deter-
Yamamoto et al. 2018). In Germany, 1,4-dioxane is not specifically reg- gents were found to be minimal (b5 mg/kg) (Fuh et al. 2005; Makino
ulated but it is classified as a non-regulated toxic substance by the Ger- et al. 2006; Tanabe and Kawata 2008). It is interesting to note that expo-
man Federal Environmental Agency which has suggested a guideline sure to 1,4-dioxane has been estimated to be greater in women (1.2
value of 0.1 μg/L (Mulisch et al. 2003). In contrast, there is currently μg/kg body weight/day) than men or children. The elevated exposure
no federal maximum contaminant level for 1,4-dioxane in the U.S. was primarily attributed to inhalation of the volatilized contaminant
This has resulted in different criteria and guidelines across states that from personal care products and, to a lesser extent, from dermal absorp-
range from 0.3 μg/L in Massachusetts to 77 μg/L in Alaska (Table 1). tion via topical application (Bronaugh 1982; ECB 2002; Health Canada
The wide range of acceptable concentrations of 1,4-dioxane across juris- and Environment Canada 2010; NICNAS 1998). The exposure through
dictions is partially attributable to differing exposure assumptions and the use of body wash, shampoo, and moisturizers, albeit relatively low
cancer risk targets (See Section 4.5: Cancer Potency Factor). at 0.04 μg/kg body weight/day, were also reported for infants (Health
Canada and Environment Canada 2010).
3.2. Food
3.4. Bathing and showering
1,4-Dioxane has been found in various food products (Health
Canada 2018; Nishimura et al. 2004). Sources of 1,4-dioxane contami- Volatile contaminants in drinking water such as halogenated disin-
nation include adhesive material used on food packaging, pesticides fection byproducts contribute to inhalation and dermal exposure during
K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866 857

bathing and showering (Krishnan and Carrier 2008). However, these 4.2. Metabolism and elimination
exposure routes have been estimated to be negligible compared to
other routes, although data are quite limited (Health Canada 2018). Defining the mechanism of 1,4-dioxane toxicity requires a compre-
hensive understanding of its metabolic pathways. While few studies
have described 1,4-dioxane metabolism in humans, experiments in
4. Toxicology and carcinogenesis rats have demonstrated that it is rapidly absorbed and metabolized to
HEAA and, subsequently, to diglycolic acid (Besenhofer et al. 2010;
Organ systems primarily affected by 1,4-dioxane include the liver Landry et al. 2011). 1,4-dioxane metabolism is also likely to result in
and kidneys. Adverse nasal and ocular effects have also been re- production of HEAA and diglycolic acids. The metabolism to HEAA is
ported (DeRosa et al. 1996; Dourson et al. 2014; Fujiwara et al. postulated to involve multiple pathways occurring in the liver, kidney,
2008; Kano et al. 2009; Kasai et al. 2009; Kasai et al. 2008; Stickney and nasal mucosa (Fig. 2) (Woo et al. 1977). Cytochrome P450s (specif-
et al. 2003). This section discusses studies evaluating 1,4-dioxane- ically CYP2E1) appear to be involved in two of these pathways (Nannelli
induced toxicity, as well as the cancer and non-cancerous effects of et al. 2005b). The third hypothesized pathway proposes that 1,4-diox-
this contaminant. ane is first metabolized to diethylene glycol, followed by conversion to
2-hydroxyethoxyacetaldehyde which is rapidly metabolized to HEAA
(Centers for Disease Control and Prevention 2017; Schier et al. 2013).
4.1. Toxicity Bioremediation studies have shown that monooxygenase-expressing
microorganisms can similarly metabolize 1,4-dioxane into glyoxylate
When administered chronically to rats and mice via drinking water, via the production of 2-hydroxyethoxyacetaldehyde, HEAA, and 1,4-di-
1,4-dioxane was shown to induce dose-dependent liver toxicity, mani- oxane-2-one (Grostern et al. 2012) (See Section 2.3: Environmental
festing as degeneration of hepatic cells and development of pre- Fate).
neoplastic lesions (Table 2) (EPA, 2013c; Fishbein 1981; Goldsworthy Exposure studies suggest that the average half-life of 1,4-dioxane in
et al. 1991; Kano et al. 2009; Stott et al. 1981). Other hepatic effects of human body is relatively short at approximately 1 h (Braun 1977; Braun
chronic 1,4-dioxane exposure via drinking water include increased and Young 1977) (Eckert et al. 2013; Göen et al. 2016; Young et al.
liver to body weight, centrilobular swelling, necrosis, increased DNA 1976) (Table 3). For example, a chamber exposure study found that in-
synthesis, chromosomal damage, and enzyme leakage in rats and mice halation exposure to 50 ppm (v) of 1,4-dioxane caused elevated 1,4-di-
(Kano et al. 2009; Roy et al. 2005; Stott et al. 1981). Chronic exposure oxane concentrations in plasma immediately following exposure but
to 1,4-dioxane through drinking water consumption has also been re- levels rapidly decreased to below detectable limits by six hours after
ported to impact the kidney, causing degeneration of cortical tubule the end of exposure (Young et al. 1977). Reflecting the rate of 1,4-diox-
cells, tubular necrosis, and glomerulonephritis (EPA, 2013c). ane metabolism, the half-life of HEAA was much longer at 2.7 h; its

Table 2
1,4-Dioxane-Induced Liver Toxicity (Adapted from U.S. EPA 2013).

Study Designa Doseb (mg/kg/d) Cytotoxicityc Proliferation/hyperplasiac Liver tumorsd

Kociba et al. (1974) Rats 0 − − 2/106


Male and females 14 − − 0/110
104-week exposure 121 + + 1/106
1307 + + 12/66
NCI (1978) Rats 0 − − 2/31
Males 240 − − 2/32
110-week exposure 530 − − 1/33
NCI (1978) Rats 0 − − 0/31
Females 350 − − 10/33
110-week exposure 640 − − 11/32
NCI (1978) Mice 0 − − 8/49
Males 720 − − 19/50
90-week exposure 830 − − 28/47
NCI (1978) Mice 0 − − 0/50
Females 380 − − 21/48
90-week exposure 860 − − 35/37
Kano et al. (2009) Rats 0 − − 3/50
Males 11 − − 4/50
104-week exposure 55 − − 7/50
274 + − 39/50
Kano et al. (2009) Rats 0 − − 3/50
Females 18 − − 1/50
104-week exposure 83 − − 6/50
429 − − 48/50
Kano et al. (2009) Mice 0 − − 23/50
Males 49 − − 31/50
104-week exposure 191 − − 37/50
677 + − 40/50
Kano et al. (2009) Mice 0 − − 5/50
Females 66 − − 35/50
104-week exposure 278 − − 41/50
964 + − 46/50
a
Design includes the species under investigation, animal genders used and exposure duration.
b
Animals were exposed to 1,4-dioxane in drinking water.
c
“+”: increased cytotoxicity; “-”: no evidence of cytotoxicity.
d
Numerator: incidence of liver tumor; Denominator: effect number of animals.
858 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

Fig. 2. Postulated metabolic pathways of 1,4-dioxane in humans.


(Adapted from Woo et al. 1977)

concentrations lagged behind 1,4-dioxane concentrations by an hour several isoforms of CYP450 in various tissues following acute oral ad-
and persisted over 18 h after the exposure (Sweeney et al. 2008; ministration by gavage or drinking water (Nannelli et al. 2005a). Thus,
Young et al. 1977). During an eight-hour inhalation exposure, 1,4-diox- it is plausible the toxic metabolites produced, such as diethylene glycol
ane and HEAA concentrations reached steady state levels in blood and diglycolic acid in combination with the induction of the CYP iso-
within the first few hours of exposure, and were detected in urine for forms (i.e. CYP2E1) in the liver may have contributed to the hepatic tu-
up to four hours and 24 h post-exposure, respectively (Göen et al. morigenesis responses observed.
2016) (Table 3). Urinary concentrations of 1,4-dioxane measured across It is noteworthy that all the aforementioned co-contaminants of 1,4-
the eight-hour exposure period corresponded to approximately 0.3% of dioxane, 1,1,1-TCA, 1,1-DCA, and TCE, have been shown to cause carci-
the theoretically inhaled total 1,4-dioxane amount, while HEAA ex- nogenic responses in the liver and kidneys of animals via the formation
creted in urine represented 53% of the estimated exposure. of reactive metabolites (ATSDR 2006). To the best of our knowledge, no
studies have examined the effects of mixtures involving 1,4-dioxane
4.3. Carcinogenic effects and its chlorinated solvent co-contaminants. Additive toxic effects on
the liver and kidney are plausible for binary and higher-order combina-
To date, in vivo studies have documented 1,4-dioxane-induced tions of these chlorinated compounds with 1,4-dioxane, as has been re-
tumor formation in the liver and nasal cavity following chronic expo- ported for the chlorinated compounds themselves in interaction studies
sure through inhalation or ingestion (Table 2) (Dourson et al. 2017; (Goldsworthy and Popp 1987; Inoue et al. 1989).
Wilbur et al. 2012) (ATSDR 2012) (Kano et al. 2009; Kasai et al. 2009;
Kociba et al. 1974). The formation of tumors in the liver has been further 4.4. Mechanisms of carcinogenesis
demonstrated to be dose-dependent, with enhanced effects in female
mice compared to males over a two-year period (Kano et al. 2009). In Despite the evidence of a carcinogenic effect, the mechanism by
contrast, in vitro studies have failed to demonstrate any genotoxic ef- which 1,4-dioxane causes cancer is currently unclear (EPA, 2013). The
fects induced by 1,4-dioxane (EPA, 2013). It is unclear whether 1,4-di- data summarized in Table 2 do not support a cytotoxicity-based mech-
oxane or its metabolites are responsible for the formation of anism as six out of ten studies showed no cytotoxic effect. All studies,
hepatocellular tumors. While the primary metabolite found in urine, however, demonstrated the presence of liver tumors. In addition, tu-
HEAA, has not been reported to be highly toxic or carcinogenic, other mors have been found at exposure doses that were without overt cyto-
metabolites, including diethylene glycol and diglycolic acid (Fig. 2), toxicity in the liver (Kano et al. 2009). Interestingly, studies have shown
have been shown to exhibit toxicity in the liver and kidney (Conrad that 1,4-dioxane appears to be a tumor promoter rather than initiator
et al. 2016; Landry et al. 2015; Landry et al. 2011; Schier et al. 2013). (Dourson et al. 2014; EPA, 2013; Kitchin and Brown 1990; Lundberg
Furthermore, the metabolism of 1,4-dioxane has been shown to induce et al. 1987). The data showed that 1,4-dioxane promoted the

Table 3
Summary of Human 1,4-Dioxane Toxicokinetic and Biomonitoring Studies.

Study Exposure concentration, route, Study populationa Sample Measurement time, post- exposure Average concentration
duration type (hrs)b (mg/L)

1,4-Dioxane HEAA

Chamber exposure 50 ppm Healthy Adult volunteers Urine 0 4.3 422


Young et al. (1977) Inhalation N=4 4 0.22 473
Sweeney et al. (2008) 6h 18 b 0.07 3.3
0 13 5.0
Plasma 4 1.5 2.0
18 b 0.07 b 0.5
Chamber exposure 20 ppm Healthy Adult volunteers Urine −2.5 0.85 160
Eckert et al. (2013) Inhalation N = 18 1.5 Not reported 204
Göen et al. (2016) 8h −4.0 0.98 Not measured
Whole blood
0 1.1 Not measured
Occupational exposure 1.6 ppm Adult Dioxane plant workers Urine 0 0.31 49.7
Young et al. (1976) Inhalation N=5
7.5 h
NHANES, 2013–2014 Not measured General Adult population Whole blood Spot samples b 0.0015 Not measured
CDC (2017) N = 3125
a
N indicates the number of participants in each study.
b
Negative number indicate samples that were collected during the exposure period.
K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866 859

carcinogenic potential of diethylnitrosamine in rats (Lundberg et al. potency factor and ultimately setting a regulatory standard for drinking
1987), and significant mortality was observed when it was adminis- water.
tered for 60 weeks as a promoter (4 male and 5 female mice survived,
n = 30/sex) (King et al. 1973). 5. Exposure assessment
A mechanistic transcriptomic analysis conducted in rats exposed to
1,4-dioxane in drinking water (440 mg/kg-day) for four weeks suggested Quantifying exposure to 1,4-dioxane presents analytical challenges
that 1,4-dioxane exposure elicited a unique pattern of altered gene ex- due to its high affinity for water, lack of readily detectable functionality,
pression in rat liver. These marker genes had previously been used to seg- and low concentration in environmental and biological samples. Nu-
regate genotoxic (3,3-dimethylbenzidine, diethylnitrosamine) from non- merous analytical techniques have been suggested for 1,4-dioxane de-
genotoxic (diethylhexylphthalate) carcinogen expression profiles tection over a wide concentration range in a broad spectrum of
(Furihata et al. 2018). More specifically, gene expression profiles treated aqueous matrices. Analytical approaches for measuring 1,4-dioxane
with 1,4-dioxane were different from those with typical genotoxic usually include a preparation stage (e.g., pre-concentration or extrac-
hepatocarcinogen for five genes (Bax, Btg2, Cdkn1a, Lrp1, and Plk2) and tion) to improve sensitivity, which is then followed by separation
substantially different from treatment with a typical non-genotoxic using gas chromatography (GC) and detection by mass spectrometry
hepatocarcinogen for nine genes (Aen, Bax, Btg2, Ccnf, Ccng1, Cdkn1a, (MS) or flame ionization (Fig. 3). These methods for exposure assess-
Mbd1, Phlda3, and Tubb4b). A more recent in vivo study concluded that ment are discussed below.
1,4-dioxane was clastogenic (i.e., induced chromosomal breakages) in
the liver but not genotoxic in the bone marrow (Itoh and Hattori 2019). 5.1. Environmental samples
Indirect genotoxic mechanisms related to inhibition of DNA repair is
one of several plausible mechanisms to explain the divergent Several pre-concentration techniques have been developed over the
genotoxicity responses observed by in vitro and in vivo studies investigat- past four decades to concentrate 1,4-dioxane from water samples, in-
ing 1,4-dioxane. cluding purge-and-trap, headspace, and phase extraction. These
Based on this evidence, it is reasonable to conclude that the mecha- methods have been reported to achieve detection limits ≈ 10 μg/L
nism by which 1,4-dioxane promotes cancer is not easily explained by (Table 4) (EPA, 1984; EPA, 1996a; EPA, 1996b; EPA, 1996c; EPA,
classic schemes. Further studies are needed to understand the cellular 1996d; Friant and Suffet 1979). Given the miscibility of 1,4-dioxane in
mechanism of 1,4-dioxane-induced carcinogenesis using a systems bi- water, its low volatility from water, and low partition coefficient, vari-
ology approach employing advanced transcriptomic and metabolomic ous modifications of these preparation techniques have been suggested
techniques. to attain the sensitivity required for drinking water monitoring (1 μg/L)
(Epstein et al. 1987; Park et al. 2005). For example, modified purge-and-
trap procedures with extended purge time and temperature as well as
4.5. Cancer potency factor increased addition of salt improve detection limits, i.e., down to 0.15
μg/L (Sun et al. 2016). Frozen microextraction and continuous liquid-
The exposure dose-dependent formation of liver tumors by 1,4-di- liquid extraction of water samples further improve detection sensitivity
oxane reported by Kano et al. (2009) was used as a basis for dose- to 0.20 μg/L (Draper et al. 2000).
response modeling and human extrapolation by the U.S. EPA's Integrate The heightened regulatory attention to 1,4-dioxane prompted the
Risk Information System to derive a cancer potency factor of 0.1 development of advanced analytical methods for detection of trace
mg/kg/day (EPA, 2013). This cancer potency factor was subsequently quantities (b0.1 μg/L). Using solid phase extraction of 1,4-dioxane as a
used to develop the federal health risk guideline of 0.35 μg/L for 1,4-di- pre-concentration approach for environmental samples minimizes in-
oxane in drinking water based on a cancer de minimis one-in-a-million- terferences and lowers detection limits to 0.05 μg/L (Carrera et al.
reference value. 2017; Isaacson et al. 2006; Song and Zhang 1997). A decade ago, the
The Kano et al. (2009) study enabled calculation of the cancer po- U.S. EPA released Method 522 for determination of 1,4-dioxane in
tency factor because it tested a broad range of doses in male and female drinking water by solid phase extraction using coconut charcoal or syn-
rats and mice. However, there was a large gap between the lowest dose thetic carbon cartridges and dichloromethane elution prior to GC–MS
in the most sensitive test group (66 mg/kg/day; female mice), the calcu- analysis with selected ion monitoring (EPA, 2008). This resulted in a de-
lated human equivalent benchmark concentration (7 mg/kg/day, lower tection limit between 0.020 and 0.026 μg/L in drinking water, suitable to
limit) (EPA, 2013) and the environmental exposure levels relevant for detect 1,4-dioxane below the one-in-a-million cancer risk level
consideration of drinking water criteria (0.35 to 10 μg/L). Extrapolating established by the U.S. EPA (0.35 μg/L).
responses from animal to humans across multiple orders of magnitude While the laboratory-based methods discussed above accurately
is typical in cancer risk assessment; however, it requires low-dose characterize trace levels of 1,4-dioxane in environmental samples,
modeling assumptions to be made to predict the risk at the much such analytical measurements present challenges when monitoring dy-
lower doses to which humans are likely to be exposed. Given the lack namic systems, such as public water supplies. Portable sensors that pro-
of definitive mechanistic information, the U.S. EPA has assumed a linear vide real-time information about the temporal variability of 1,4-dioxane
low-dose approach from the calculated human equivalent benchmark concentrations in the field are lacking but would be of high value to
concentration. This approach has been disputed based on evidence water systems operators. Development of sensors for 1,4-dioxane for
that hepatotoxicity and regenerative hyperplasia co-occurred in some deployment across public water distribution systems involves several
1,4-dioxane dose-response studies, suggesting a plausible basis for a aspects, such as a recognition event wherein 1,4-dioxane is specifically
cytotoxicity-based threshold response (Dourson et al. 2014, 2017). and selectively detected through interaction with the sensor receptors,
However, as stated above and shown in Table 2, this was not always and a corresponding signal response to report on the interaction, e.g.,
the case, since liver tumorigenesis was observed in absence of chemical, optical, or electrical response. The sensing components
cytotoxicity. should then be translated into practical devices able to perform onsite
Additional mechanistic studies are needed to understand how 1,4- under environmental conditions. Finally, connection of the sensing de-
dioxane may contribute to liver carcinogenesis in humans at doses vice to a wireless sensor network would provide an additional advan-
below the levels tested by Kano and coworkers and that are more rep- tage for real-time monitoring of 1,4-dioxane. Ideally, sensitivity to 1,4-
resentative of actual low, environmentally-relevant levels of exposure. dioxane would be achieved through the design of specific sensing ele-
This information is critical to allow quantitative risk assessments for ments that recognize 1,4-dioxane as a single target. In that case, a recog-
1,4-dioxane that can form the basis of deriving a more accurate cancer nition event might occur through selective binding of substances or
860 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

Fig. 3. Analytical methods commonly used for measuring 1,4-dioxane in environmental and biological samples.

chemical reaction. Several studies have described various sensor tech- Development of these highly sensitive sensors for real-time monitoring
nologies for remote monitoring of 1,4-dioxane in aqueous samples. of 1,4-dioxane in the field would enable detection of dynamic contami-
Kumar et al. (2015) developed a silver nanoparticle-based nation events and thereby facilitate rapid notification of the public of
solvatochromic sensor that selectively quantifies 1,4-dioxane in various exceedances in public water supplies.
solvents with a 1 mg/L detection limit. Sensors based on infrared and
photoacoustic spectroscopy have also been investigated but exhibit 5.2. Biological samples
poor 1,4-dioxane sensitivity (20 mg/L detection limit) (De Melas et al.
2003; Holthoff et al. 2015). Luminescent porous organic polymers Biomonitoring is a popular approach for personal exposure assess-
have also been created that quantify 1,4-dioxane in glycol medium ment of a contaminant in a biological sample. Human biomonitoring
(Heuvelsland 1988; Ma et al. 2016; Zhou et al. 2015). The approaches studies use GC–MS for separation and detection and follow the same
developed thus far offer low selectivity and detection limits insufficient methods used for detection of 1,4-dioxane in water samples
to detect U.S. EPA-based action levels of 1,4-dioxane. Sensors that use (Section 5.1: Environmental Samples). However, extraction techniques
DNA aptamers, nanobodies, or antibodies have shown promise for used for biological samples (e.g., urine, plasma, whole blood) differ from
their high selectivity, and their affinity for target molecules (commonly those used for water samples. Headspace and liquid-liquid extraction
in the 0.1 to 10 nM range) (Hamaguchi et al. 2001; Liss et al. 2002). protocols have been reported for 1,4-dioxane with detections limits of
When these biomolecules have been coupled to electrochemical dis- 100 and 137 μg/L, respectively (Table 4) (Eckert et al. 2013; Göen
placement sensors, exceptionally low detection levels (pM to aM et al. 2016; Sweeney et al. 2008). An additional method using direct in-
range) have been realized (Hansen et al. 2006; Zhang et al. 2018). jection of urine and plasma sample reported a detection limit of 70 μg/L

Table 4
Summary of Analytical Techniques and Water Methods for Measurement of 1,4-Dioxane in Water and Biological Fluids.

Study Limit of detection (μg/L) Sample matrix Sample volume (mL) Extraction and analysis techniquea

Friant and Suffet (1979) 740 Water 5 Headspace GC–MS


EPA US (1996a, b, c, d) 12 Water 5–25 Liquid-liquid extraction; GC–MS
Methods 5031, 5032
EPA US (1996a, b, c, d) 12 Water 5–25 Liquid-liquid extraction; GC–MS
Methods 8260B, 8015B
U.S. EPA (1984) 10 Water 5 Purge-and-trap; GC–MS
Method 1624
Park et al. (2005) 6 Water 200 Liquid-liquid extraction; GC–MS
Epstein et al. (1987) 1 Water 4000 Liquid-liquid extraction; GC-FID
Isaacson et al. (2006) 0.31 Water 80 Solid-phase extraction; GC–MS
Draper et al. (2000) 0.20 Water 1000 Liquid-liquid extraction; GC–MS
Sun et al. (2016) 0.15 Water 5 Purge-and-trap; GC–MS
Song and Zhang (1997) 0.05 Water 1 Solid-phase extraction; GC–MS
Carrera et al. (2017)
U.S. EPA (2008) 0.02 Water 100 Solid-phase extraction; GC–MS
Method 522
Sweeney et al. (2008) 137 Blood 0.1 Liquid-liquid extraction; GC–MS
Göen et al. (2016) 100 Urine, blood 2 Headspace GC–MS
Braun (1977) 70 Urine, plasma 0.5 Direct injection; GC–MS
Braun and Young (1977)
Chambers (2017) 0.5 Blood 3 Headspace solid-phase microextraction, GC–MS
a
GC: gas chromatography; MS: mass spectrometry; FID: flame ionization detector.
K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866 861

for 1,4-dioxane (Braun 1977; Braun and Young 1977). While these and cause-specific (N = 1.7) deaths, the difference was not statistically
methods were able to detect 1,4-dioxane in blood and urine following significant. Kramer et al. (1978) recruited 151 matched pairs of exposed
moderate to high exposure levels (1.6–50 ppm), their sensitivity was and control workers from two adjacent textile plants in Michigan be-
reduced compared to most techniques used for monitoring environ- tween 1969 and 1975; one plant used 1,1,1-TCA as a cleaning solvent
mental water samples (0.02–12 μg/L) (Table 4). More recently, an ex- which contained 4% 1,4-dioxane. This study did not find any alterations
traction technique using headspace solid-phase microextraction in cardiovascular or hepatic parameters to be associated with occupa-
developed at the U.S. Centers for Disease Control and Prevention tional exposure to 1,1,1-TCA. Thiess et al. (1976) conducted a cohort
(CDC) achieved the most sensitive levels of detection than other bio- study in Germany between 1964 and 1974 that included 74 workers
monitoring approaches for 1,4-dioxane detection (0.5 μg/L) (Table 4) who were exposed to 1,4-dioxane. The number of overall deaths (N
(Chambers 2017). This method was used to test whole blood samples = 12) and cancer-related deaths (N = 2, one lamellar epithelial carci-
from the U.S. general population (NHANES 2013–2014). All 3125 sam- noma and one myelofibrotic leukemia) were not significantly increased
ples analyzed were found to be below the level of detection (Centers over the expected number of cancer deaths in age- and gender-match
for Disease Control and Prevention 2017). It is clear that evaluation of patients in the general population (N = 14.5). Stickney et al. (2003)
the general population to low dose exposure to 1,4-dioxane requires reviewed an unpublished study conducted in 1976 which reported
improved analytical methods with increased sensitivity for 1,4-dioxane four cancer cases (a colon cancer, a pulmonary cancer, a lymphosar-
detection. coma, and a glioblastoma) among 80 workers who were exposed to
As previously discussed, 1,4-dioxane is rapidly metabolized via cyto- 1,4-dioxane. The observed number of cancer cases was similar to the ex-
chrome P450 monooxygenases to 1,4-dioxane-2-one and HEAA (See pected number of cancer deaths in the general population. While this
Section 4.2: Metabolism and Elimination). Human chamber studies small number of available epidemiologic studies do not support an asso-
have demonstrated that HEAA can be used as a short-term surrogate ciation between 1,4-dioxane and cancer, only limited conclusions can
marker of moderate-to-high inhalation exposure to 1,4-dioxane be drawn from these studies because their sample sizes were too
(Table 3). While HEAA has not been evaluated as a biomarker of envi- small to allow detection of a low-level elevated cancer risk. In addition,
ronmental exposure to 1,4-dioxane, the sensitivity of analytical tech- the reported cancers had sites of origin that were not in line with those
niques to detect low levels of HEAA supports the feasibility of seen in animal models (which mainly manifested in the liver or nasal
measuring this compound in the general population. However, an im- cavity; See Section 4.4: Carcinogenic Effects) (Stickney et al. 2003).
portant caveat of using HEAA for such exposure assessment is the Available exposure assessment methods have limited the feasibility
non-specificity of this metabolite to 1,4-dioxane. For example, HEAA is of epidemiological studies to evaluate 1,4-dioxane across general popu-
also the major metabolite produced from diethylene glycol, diethylene lations (See Section 5: Exposure Assessment). 1,4-Dioxane concentra-
glycol ethers, and diethylene glycol monoethyl ether, all of which are tions can be measured in public drinking water supplies to infer
used in antifreeze preparations, petroleum solvent extraction, produc- exposure in large populations; however, a single residential estimate
tion of polyurethanes and unsaturated polyester resins, and dehydra- of 1,4-dioxane in water is subject to exposure misclassification depend-
tion of natural gas (Göen et al. 2016; Larranaga et al. 2016; Schier ing on an individual's consumption of water sourced from locations
et al. 2013). Nevertheless, environmental exposure of the general pop- other than home, e.g., from work or bottled water. Exposure misclassifi-
ulation to these chemicals is limited (Schriks et al. 2010; Stefania et al. cation can be reduced by using biological samples to assess 1,4-dioxane
2009). exposure. Unfortunately, the low sensitivity of current analytical tech-
As an alternative to measuring 1,4-dioxane directly or focusing niques for biomonitoring 1,4-dioxane inhibits exposure assessment in
solely on the major metabolite of this contaminant, MS-based metabo- the general population. Only through the development of new, more
lomics may serve as a possible screening tool for risk assessment. This sensitive detection methods will meaningful epidemiological studies
approach could be used to characterize an endogenous signature of me- be possible in humans.
tabolite biomarkers for the assessment of 1,4-dioxane exposure. In addi-
tion, the metabolomic signature of 1,4-dioxane could be used as a basis 7. Remediation methods
to investigate the potential toxic effects associated with biomarker indi-
ces of exposure. This can be performed through evaluation of biochem- The unique physical and chemical characteristics of 1,4-dioxane
ical pathways and networks. (Rattray et al. 2018). present challenges in remediation from contaminated groundwater
(Stepien et al. 2014). 1,4-dioxane favors partitioning into the aqueous
6. Health effects in humans phase where it is completely miscible, rendering common physical sep-
aration techniques, such as air stripping or adsorption by activated car-
1,4-Dioxane has been classified as a probable human carcinogen bon, unsuitable. Removal by membrane filtration is also ineffective;
based on evidence from experimental animal studies but there is a pau- even reverse osmosis fails to reach higher than 70% rejection due to its
city of data related to the impact of 1,4-dioxane exposure on human small size and neutrality (Kegel et al. 2010; Linares et al. 2011). These
health. A small study in 2006 evaluated responses induced by controlled physical removal processes, even if they are effective, simply transfer
1,4-dioxane exposure. This chamber study included twelve healthy vol- 1,4-dioxane from one phase to another, wherein it is more concentrated
unteer participants and concluded that exposure to 20 ppm 1,4-dioxane and requires subsequent treatment. Abiotic and biotic oxidation that
for 2 h did not increase the frequency of symptoms such as blinking, mineralizes or otherwise partially degrades 1,4-dioxane to fewer toxic
nasal swelling, inflammation, or difficulty with pulmonary function products are desired. These remediation strategies of contaminated
(Ernstgård et al. 2006). groundwater are discussed in this section.
All other human studies have exclusively focused on elevated occu-
pational exposures in small cohorts (N 〈200) with most evaluating 7.1. Abiotic advanced oxidation
cancer-related mortality. Buffler et al. (1978) conducted a retrospective
cohort study to estimate the association between occupational expo- 1,4-Dioxane degradation by conventional oxidants such as chlorine
sure to 1,4-dioxane and cancer mortality among 165 workers at a is not effective (Klečka and Gonsior 1986; Li et al. 2018) and can lead
major chemical plant in Texas. No estimate was given by authors of to the formation of the chlorinated byproducts that are more toxic
workers' exposure levels to the 1,4-dioxane. Between April 1, 1954 than the parent 1,4-dioxane (Li et al. 2018; Woo et al. 1980). Alterna-
and June 30, 1975, 12 deaths occurred in these plant workers, of tively, past studies have demonstrated mineralization of 1,4-dioxane
which three were cancer-related. Although the number of observed by reactive radicals such as hydroxyl (•OH) and sulfate radicals
deaths was higher than the expected Texas age-sex-race (N = 4.9) (SO4•−) in advanced oxidation processes (Fig. 4) (Campos-Martin
862 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

Fig. 4. Pathways of abiotic advanced oxidation processes (adapted from Zhang et al. 2017). Compounds shown in a dotted box are biodegradable intermediates.

et al. 2006; Wacławek et al. 2017). However, practical implantation re- and the high cost of typical boron-doped diamond electrodes (com-
quires effective pretreatment that removes other water constituents monly used in such electrochemical cells) remain unresolved obstacles
such as natural organic matter which can reduce the efficiency of 1,4-di- (Choi et al. 2010; Jasmann et al. 2016). Photocatalytic advanced oxida-
oxane degradation by scavenging radicals. Engineering advanced oxida- tion processes in a slurry reactor coupled with membrane filtration
tion is also complicated by several challenges, including the need for are also appealing options due to effectiveness in completely degrading
continual dosing of precursor chemicals and provision of high energy 1,4-dioxane. However, the high energy consumption to provide light
(e.g., heat or ultraviolet (UV) irradiation), since short-lived radicals (often UV) and pressure for membrane filtration requires careful cost
need to be generated in situ by “activating” stable precursor oxidants analysis (Hwangbo et al. 2019). The integration of diverse unit opera-
such as ozone (O3), hydrogen peroxide (H2O2), and persulfate (HSO− 5 tions, such as electrochemical treatment with membrane filtration or
and S2O2− 8 ) (Barndõk et al. 2014; Coleman et al. 2007; Feng et al. ozonation, have yielded promising results that warrant further investi-
2017; Lee et al. 2015; Son et al. 2009; Zhong et al. 2015). The lack of a gation (Kishimoto et al. 2008; Mameda et al. 2017).
facile method to quantify trace levels of 1,4-dioxane also makes it diffi-
cult for advanced oxidation process to be responsive to spatial and tem- 7.2. Bioremediation
poral variations in water quality (See Section 5.1: Environmental
Samples). Biological degradation of 1,4-dioxane has been explored as an al-
Alternative oxidative treatments continue to emerge but with both ternative approach to remediate contaminated water. As described
longstanding and novel technical barriers to overcome. Direct oxidation in Section 2.3: Environmental Fate, select bacterial and fungal strains
of 1,4-dioxane at the anode of an electrochemical cell has been shown to are capable of degrading 1,4-dioxane through metabolic and co-
proceed under conditions of low conductivity which may be relevant to metabolic pathways under aerobic conditions (Gedalanga et al.
groundwater. However, a lack of selectivity (which can lead to the for- 2014; Lan et al. 2013; Mahendra and Alvarez-Cohen 2006; Zhang
mation of toxic halogenated byproducts in the presence of chloride) et al. 2017). Field testing at contaminated sites has shown the
K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866 863

promise of bioremediation under specific conditions (e.g., the pres- spatially dispersed nature of private wells, the potential for changing
ence of nutrients and dissolved oxygen), while the inhibitory effects water supply characteristics, such as groundwater levels and plume
of certain metals and chlorinated solvents are limiting factors (Shen movement. Research to improve detection capabilities will involve
et al. 2008; Zhang et al. 2017). Further enhancement of 1,4-dioxane the development of highly selective binding and sensing compo-
removal efficiency has been achieved through a bioaugmented ap- nents and will require field testing under various scenarios.
proach wherein 1,4-dioxane-degrading bacterial strains are grown 4. Research to improve 1,4-dioxane treatment. Advanced oxidation
on granular activated carbon adsorbents (Myers et al. 2018). Field in- processes are the promising option for centralized large-scale treat-
vestigations evaluating the feasibility and utility of these ment of drinking water which is distributed to consumers through
bioaugmented adsorbents for surface groundwater remediation are a network of water distribution pipes. However, application for a
still necessary. While biological degradation can be a viable option decentralized treatment such as smaller public supplies and point-
for site remediation, slow degradation kinetics is likely to limit its of-use household treatment is challenging. Considering the prevalent
application for potable water treatment (Li et al. 2010). contamination in groundwater that serves as a source for drinking
water in many contaminated sites, advancing advanced oxidation
8. Implications for further studies process technologies for smaller, modular applications are needed.
Immediate research needs include (i) making smaller advanced oxi-
1,4-Dioxane contamination in groundwater illustrates key scientific dation process systems cost efficient, (ii) better control of reaction
and policy challenges that can occur with emerging contaminants. A oxidation byproducts, and (iii) effectively integrating with other
current lack of scientific consensus on low dose health effects of 1,4-di- treatment options to enhance the efficiency of advanced oxidation
oxane has inhibited the establishment of enforceable federal standards process (e.g., pretreatment to remove organic matter that reduce
on safe drinking water targets. Consequently, major gaps remain in the treatment efficiency).
water testing, exposure assessment, and effective treatment. Millions 5. Research to investigate co-occurring exposures. Past epidemio-
of drinking water consumers in the U.S. are estimated to be ingesting logic study results may have been confounded by exposure of
1,4-dioxane at levels that confer an elevated cancer risk. While current human subjects to other solvents (and/or other risk factors) since
research provides an indication of the health risks and informs policy few studies isolated exposure to common co-occurring 1,4-dioxane
options, there are major scientific gaps that need to be filled to improve contaminants, such as TCE or 1,1,1-TCA. Therefore, it is important
health protection from this contaminant. to consider co-contaminants when studying 1,4-dioxane.
1. Research to establish drinking water targets. The variations in
Successful completion of these research activities will address the
state guidance values for 1,4-dioxane in water demonstrates the
key questions still outstanding for 1,4-dioxane as an environmental
lack of consensus on the methodology to derive a health-based tar-
contaminant, and enable improvements in community engagement
get. The major discrepancy in these guidance values is rooted in the
and public health protection
poor understanding of the 1,4-dioxane cancer mechanism and how
to apply the high dose animal data to risk associated with chronic ex-
9. Suggestions for community engagement and preparedness
posure to low, environmentally-relevant levels. Early theories that
1,4-dioxane was a completely non-genotoxic agent, causing liver
Engagement with communities about 1,4-dioxane in the water sup-
cancer only at high cytotoxic doses, appear to have less credibility
ply will uncover a multitude of concerns. Likely areas of concern include
today in light of the evidence for in vivo genotoxicity and carcinoge-
proximity of 1,4-dioxane sources to drinking water supplies, responsi-
nicity at moderate, non-cytotoxic doses (Gi et al. 2018; Kano et al.
ble parties, health effects, regulatory levels in water, vulnerable popula-
2009). Given the lack of in vitro mutagenicity of 1,4-dioxane (or its
tions, availability and sufficiency of current treatment systems, and
metabolites), alternative hypotheses need to be examined. Thus, re-
recommendations for medical screening and for lowering health risks
search has focused on how 1,4-dioxane produces genetic damage
in the future. Addressing these concerns requires the participation of
in vivo but not in vitro (Morita and Hayashi 1998). Inhibition of
numerous stakeholders and agencies involved in environmental inves-
DNA repair is a plausible hypothesis given that one difference be-
tigation, risk assessment, and site remediation processes, and who com-
tween in vivo and in vitro systems is the greater likelihood of back-
monly span local, state, and federal levels.
ground DNA damage in need of repair in vivo. Inhibition of DNA
Identifying and responding to community concerns about 1,4-diox-
repair could have promotional and co-carcinogenic effects rather
ane in an open and transparent way is essential for meaningful engage-
than initiating activity which fits with the experimental evidence re-
ment and broad participation as remedial options are evaluated and site
lating to 1,4-dioxane. Future studies should examine this and other
decisions are made. Technical information regarding the potential
mechanisms for genotoxic effects that do not involve direct interac-
health risks associated with 1,4-dioxane exposure (e.g., liver cancer)
tion with DNA, such as topoisomerase II inhibition. Sensitive
must be distilled into clear messages that can be understood by all com-
transcriptomic, metabolomic and stem cell studies are needed to un-
munity members. It is also essential to inform the public that (i) health
derstand 1,4-dioxane-induced early effects in the liver that can lead
risks are multifactorial processes, and (ii) 1,4-dioxane exposure is only
to genotoxicity and carcinogenesis.
one of potentially many environmental risk factors associated with po-
2. Research to understand 1,4-dioxane exposure and effects in
tential adverse health outcomes. Community members should be made
human populations. Sufficiently powered epidemiology studies in-
aware that there are modifiable lifestyle factors that can affect these
vestigating the general population have yet to be conducted. Such
health risks, and that are within their control. Action at the individual
studies would be enhanced by the development of sensitive bio-
stakeholder level can complement the short- and long-term solutions
markers of exposure and effect. These could examine the relative
developed by responsible parties and governmental agencies to the
contribution of exposures to 1,4-dioxane in food and consumer prod-
contamination issue.
ucts as well as contaminated water supplies. Transcriptomic and
Situations involving emerging drinking water contaminants, such as
metabolomic indicators of perturbation uncovered in animal testing
1,4-dioxane, require intensive and sustained community engagement,
could potentially be used in epidemiology studies. Such studies are
with opportunities for stakeholders to participate in the continuum of
urgently needed to better understand human exposure and the pub-
effort from investigation through remediation. Providing unfettered ac-
lic health implications of 1,4-dioxane exposure.
cess to site-related and health information, and various platforms to un-
3. Research to develop rapid, in situ detection of 1,4-dioxane. Detec-
derstand the health risks, uncertainties and options for moving towards
tion of 1,4-dioxane on-site and in real-time is critical given the
solutions allows impacted communities to actively participate in the
864 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

process and following precautionary principles. Further, it is important Centers for Disease Control and Prevention, 2017. Nationa Health and Nutrition Examina-
tion Survey: Volatie Organic Compounds (VOCs) and Trihalomethanes/MTBE-Blood-
for researchers to recognize the key role that community members Special Sample (VOCWBS_H). 2013–2014 Data Documentation, Codebook, and
can play in advocating for the interests of those most impacted, holding Frequencies.
local, state and federal agencies accountable, and supporting the devel- Chambers, D., 2017. Laboratory Procedure Manual-volatile Organic Compounds (VOCs) &
Trihalomethanes/MTBE. Centers for Disease Control, Atlanta, GA.
opment and adoption of science-based remedies. Chiang, D.S., Glover Jr., E.W., Peterman, J., Harrigan, J., DiGuiseppi, B., Woodward, D.S.,
2008. Evaluation of natural attenuation at a 1,4-dioxane-contaminated site.
Remediat. J. 19, 19–37.
Funding Choi, J.Y., Lee, Y.J., Shin, J., Yang, J.W., 2010. Anodic oxidation of 1,4-dioxane on
boron-doped diamond electrodes for wastewater treatment. J. Hazard. Mater.
179, 762–768.
This research did not receive any specific grant from funding agen-
Coleman, H.M., Vimonses, V., Leslie, G., Amal, R., 2007. Degradation of 1,4-dioxane in
cies in the public, commercial, or not-for-profit sectors. water using TiO2 based photocatalytic and H2O2/UV processes. J. Hazard. Mater.
146, 496–501.
Conrad, T., Landry, G.M., Aw, T.Y., Nichols, R., McMartin, K.E., 2016. Diglycolic acid, the
References toxic metabolite of diethylene glycol, chelates calcium and produces renal mitochon-
drial dysfunction in vitro. Clin. Toxicol. (Phila.) 54, 501–511.
Abe, A., 1999. Distribution of 1,4-dioxane in relation to possible sources in the water en- De Melas, F., Pustogov, V.V., Croitoru, N., Mizaikoff, B., 2003. Development and optimiza-
vironment. Sci. Total Environ. 227, 41–47. tion of a mid-infrared hollow waveguide gas sensor combined with a supported cap-
Adamson, D.T., Mahendra, S., Walker Jr., K.L., Rauch, S.R., Sengupta, S., Newell, C.J., 2014. A illary membrane sampler. Appl. Spectrosc. 57, 600–606.
multisite survey to identify the scale of the 1,4-dioxane problem at contaminated Deng, D., Li, F., Wu, C., Li, M., 2018. Synchronic biotransformation of 1,4-dioxane and 1,1-
groundwater sites. Environ. Sci. Technol. 1, 254–258. dichloroethylene by a gram-negative propanotroph Azoarcus sp. DD4. Environ. Sci.
Adamson, D.T., Anderson, R.H., Mahendra, S., Newell, C.J., 2015. Evidence of 1,4-dioxane Technol. Lett. 5, 526–532.
attenuation at groundwater sites contaminated with chlorinated solvents and 1,4-di- DeRosa, C.T., Wilbur, S., Holler, J., Richter, P., Stevens, Y.W., 1996. Health evaluation of 1,4-
oxane. Environ Sci Technol 49, 6510–6518. dioxane. Toxicol. Ind. Health 12, 1–43.
Adamson, D.T., de Blanc, P.C., Farhat, S.K., Newell, C.J., 2016. Implications of matrix diffu- DHHS, 2014. Report on Carcinogens. Twelfth edition. U.S. Department of Health and
sion on 1,4-dioxane persistence at contaminated groundwater sites. Sci. Total Envi- Human Services.
ron. 562, 98–107. Dourson, M., Reichard, J., Nance, P., Burleigh-Flayer, H., Parker, A., Vincent, M., et al., 2014.
Adamson, D.T., Pina, E.A., Cartwright, A.E., Rauch, S.R., Hunter Anderson, R., Mohr, T., et al., Mode of action analysis for liver tumors from oral 1,4-dioxane exposures and
2017. 1,4-Dioxane drinking water occurrence data from the third unregulated con- evidence-based dose response assessment. Regul. Toxicol. Pharmacol. 68, 387–401.
taminant monitoring rule. Sci. Total Environ. 596-597, 236–245. Dourson, M.L., Higginbotham, J., Crum, J., Burleigh-Flayer, H., Nance, P., Forsberg, N.D., et
An, Y.J., Kwak, J., Nam, S.H., Jung, M.S., 2014. Development and implementation of surface al., 2017. Update: mode of action (MOA) for liver tumors induced by oral exposure
water quality standards for protection of human health in Korea. Environ. Sci. Pollut. to 1,4-dioxane. Regul. Toxicol. Pharmacol. 88, 45–55.
Res. Int. 21, 77–85. Draper, W.M., Dhoot, J.S., Remoy, J.W., Perera, S.K., 2000. Trace-level determination of 1,4-
Anderson, R.H., Anderson, J.K., Bower, P.A., 2012. Co-occurrence of 1,4-dioxane with tri- dioxane in water by isotopic dilution GC and GC-MS. Analyst 125, 1403–1408.
chloroethylene in chlorinated solvent groundwater plumes at US Air Force installa- ECB, 2002. European Union Risk Assessment Report: CAS 123-91-1: 1,4-Dioxane.
tions: fact or fiction. Integr. Environ. Assess. Manag. 8, 731–737. European Chemical Bureau, Institute for Health and Consumer Protection,
Anderson, M.P., Woessner, W.W., Hunt, R.J., 2015. Applied Groundwater Modeling: Sim- Luxembourg.
ulation of Flow and Advective Transport. Elsevier, San Diego, CA. Eckert, E., Gries, W., Göen, T., Leng, G., 2013. Reliable quantitation of beta-
ATSDR, 2006. Toxicological Profile for 1,1,1-Trichloroethane. Agency for Toxic Substances hydroxyethoxyacetic acid in human urine by an isotope-dilution GC-MS procedure.
and Disease Registry, Division of Toxicology and Human Health Sciences, Atlanta, GA. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 935, 80–84.
ATSDR, 2012. Toxicological Profile for 1,4 Dioxane. Agency for Toxic Substances and Dis- Environmental Working Group (EWG), 2017. Hidden Carcinogen Tains Tap Water, Con-
ease Registry. sumer Products Nationwide.
ATSDR, 2015. Public Health Statement for 1,1-Dichloroethane. Agency for Toxic Sub- EPA US, 1984. Method 1624, Revision B: Volatile Organic Compounds by Isotope Dilution
stances and Disease Registry, Division of Toxicology and Human Health Sciences, At- GC/MS. United States Environmental Protection Agency (U.S. EPA).
lanta, GA. EPA US, 1996a. Method 5031: Volatile, Nonpurgeable, Water-Soluble Compounds by
ATSDR, 2016. Public Health Statement for Trichloroethylene. Agency for Toxic Substances Azeotropic Distillation. United States Environmental Protection Agency.
and Disease Registry, Division of Toxicology and Human Health Sciences, Atlanta, GA. EPA US, 1996b. Method 5032: Volatile Organic Compounds by Vacuum Distillation.
ATSDR, 2017. Support Document to the 2017 Substance Priority List (Candidates for Tox- United States Environmental Protection Agency.
icological Profiles). Division of Toxicology and Human Health Sciences, Atlanta, GA. EPA US, 1996c. Method 8015B: Nonhalogenated Organics Using GC/FID. United States En-
Barndõk, H., Cortijo, L., Hermosilla, D., Negro, C., Blanco, A., 2014. Removal of 1,4-di- vironmental Protection Agency.
oxane from industrial wastewaters: routes of decomposition under different op- EPA US, 1996d. Method 8260B: Volatile Organic Compounds by Gas Chromatogra-
erational conditions to determine the ozone oxidation capacity. J. Hazard. Mater. phy/Mass Spectrometry (GC/MS). United States Environmental Protection
280, 340–347. Agency.
Benjakul, R., 2010. Stimulating Dioxane Transport in a Heterogeneous Glacial Aquifer Sys- EPA US, 2006. Treatment Technologies for 1,4-Dioxane: Fundamentals and Field Applica-
tem (Washtenaw County, Michigan) Using Publicly Available Models and Data. Geo- tions. United States Environmental Protection Agency.
logical and Mining Engineering and Sciences M.S. Michigan Technological University, EPA US, 2008. Method 522: Determination of 1,4-Dioxane in Drinking Water by Solid
Houghton, Michigan. Phase Extraction (SPE) and Gas Chromatography Mass Spectrometry (GC/MS) with
Bernhardt, D., Diekmann, H.D., 1991. Degradation of dioxane, tetrahydrofuran and other Selected Ion Monitoring (SIM). United States Environmental Protection Agency
cyclic ethers by an environmental Rhodococcus strain. Appl. Microbiol. Biotechnol. (U.S. EPA).
36, 120–123. EPA US, 2010. IRIS Toxicological Review of 1,4-Dioxane (Final Report). United States En-
Besenhofer, L.M., Adegboyega, P.A., Bartels, M., Filary, M.J., Perala, A.W., McLaren, M.C., et vironmental Protection Agency, Washington, DC.
al., 2010. Inhibition of metabolism of diethylene glycol prevents target organ toxicity EPA US, 2011. Reportable Quantities of Hazardous Substances Designated Pursuant to
in rats. Toxicol. Sci. 117, 25–35. Section 311 of the Clean Water Act. Code of Federal Regulations. 40 CFR 302.4.
Black, R.E., Hurley, F.J., Havery, D.C., 2001. Occurrence of 1,4-dioxane in cosmetic raw ma- Washington, DC. .
terials and finished cosmetic products. J. AOAC Int. 84, 660–670. EPA US, 2013. Toxicological Review of 1,4-Dioxane (With Inhalation Update). United
Braun, W.H., 1977. Rapid method for the simultaneous determination of 1,4-dioxan and States Environmental Protection Agency, Washington, DC.
its major metabolite, beta-hydroxyethoxyacetic acid, concentrations in plasma and EPA US, 2013c. EPA Contract Laboratory Program Statement of Work for Organic
urine. J. Chromatogr. 133, 263–266. Superfund Methods SOM02.3. United States Environmental Protection Agency.
Braun, W.H., Young, J.D., 1977. Identification of beta-hydroxyethoxyacetic acid as the EPA US, 2017a. Scope of Risk Evaluation for 1,4-Dioxane.
major urinary metabolite of 1,4-dioxane in the rat. Toxicol. Appl. Pharmacol. 39, EPA US, 2017b. Technical Fact Sheet-1,4-Dioxane. United States Environmental Protection
33–38. Agency, Washington, DC.
Bronaugh, R., 1982. Percutaneous absorption of cosmetic ingredients. In: Frost, P., Horvitz, EPA US, 2018a. Problem Formulation of the Risk Evaluation for 1,4-Dioxane. United States
N. (Eds.), Principles of Cosmetics for the Dermatologist. CV Mosby, St. Louis, MO, Environmental Protection Agency, Office of Chemical Safety and Pollution Prevention,
pp. 277–284. Washington, DC.
Buffler, P.A., Wood, S.M., Suarez, L., Kilian, D.J., 1978. Mortality follow-up of workers ex- EPA US, 2018b. Toxics Release Inventory (TRI), Reporting Year 2017, Washington, DC.
posed to 1,4-dioxane. J. Occup. Med. 20, 255–259. Epstein, P.S., Mauer, T., Wagner, M., Chase, S., Giles, B., 1987. Determination of parts-per-
Campos-Martin, J.M., Blanco-Brieva, G., Fierro, J.L., 2006. Hydrogen peroxide synthesis: an billion concentrations of dioxane in water and soil by purge and trap gas chromatog-
outlook beyond the anthraquinone process. Angew. Chem. Int. Ed. Engl. 45, raphy/mass spectrometry or charcoal tube enrichment gas chromatography. Anal.
6962–6984. Chem. 59, 1987–1990.
Carrera, G., Vegue, L., Boleda, M.R., Ventura, F., 2017. Simultaneous determination Ernstgård, L., Iregren, A., Sjögren, B., Johanson, G., 2006. Acute effects of exposure to va-
of the potential carcinogen 1,4-dioxane and malodorous alkyl-1,3-dioxanes pours of dioxane in humans. Hum. Exp. Toxicol. 25, 723–729.
and alkyl-1,3-dioxolanes in environmental waters by solid-phase extraction Feng, Y., Lee, P.H., Wu, D., Shih, K., 2017. Surface-bound sulfate radical-dominated degra-
and gas chromatography tandem mass spectrometry. J. Chromatogr. A 1487, dation of 1,4-dioxane by alumina-supported palladium (Pd/Al2O3) catalyzed peroxy-
1–13. monosulfate. Water Res. 120, 12–21.
K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866 865

Fishbein, L., 1981. Carcinogenicity and mutagenicity of solvents. I. Glycidyl ethers, diox- Kim, Y.M., Jeon, J.R., Murugesan, K., Kim, E.J., Chang, Y.S., 2009. Biodegradation of 1,4-di-
ane, nitroalkanes, dimethylformamide and allyl derivatives. Sci. Total Environ. 17, oxane and transformation of related cyclic compounds by a newly isolated Mycobac-
97–110. terium sp. PH-06. Biodegradation 20, 511–519.
Friant, S.L., Suffet, I.H., 1979. Interactive effects of temperature, salt concentration, and pH King, M.E., Shefner, A.M., Bates, R.R., 1973. Carcinogenesis bioassay of chlorinated
on head space analysis for isolating volatile trace organics in aqueous environmental dibenzodioxins and related chemicals. Environ. Health Perspect. 5, 163–170.
samples. Anal. Chem. 51, 2167–2172. Kishimoto, N., Nakagawa, T., Asano, M., Abe, M., Yamada, M., Ono, Y., 2008. Ozonation
Fuh, C.B., Lai, M., Tsai, H.Y., Chang, C.M., 2005. Impurity analysis of 1,4-dioxane in nonionic combined with electrolysis of 1,4-dioxane using a two-compartment electrolytic
surfactants and cosmetics using headspace solid-phase microextraction coupled with flow cell with solid electrolyte. Water Res. 42, 379–385.
gas chromatography and gas chromatography–mass spectrometry. J. Chromatogr. A Kitchin, K.T., Brown, J.L., 1990. Is 1,4-dioxane a genotoxic carcinogen? Cancer Lett. 53,
1071, 141–145. 67–71.
Fujiwara, T., Tamada, T., Kurata, Y., Ono, Y., Kose, T., Ono, Y., et al., 2008. Investigation of Klečka, G.M., Gonsior, S.J., 1986. Removal of 1,4-dioxane from wastewater. J. Hazard.
1,4-dioxane originating from incineration residues produced by incineration of mu- Mater. 13, 161–168.
nicipal solid waste. Chemosphere 71, 894–901. Kociba, R.J., McCollister, S.B., Park, C., Torkelson, T.R., Gehring, P.J., 1974. 1,4-Dioxane. I. Re-
Furihata, C., Toyoda, T., Ogawa, K., Suzuki, T., 2018. Using RNA-Seq with 11 marker genes sults of a 2-year ingestion study in rats. Toxicol. Appl. Pharmacol. 30, 275–286.
to evaluate 1,4-dioxane compared with typical genotoxic and non-genotoxic rat Kramer, C.G., Imbus, H.R., Ott, M.G., Fulkerson, J.E., Hicks, N., 1978. Health of workers ex-
hepatocarcinogens. Mutat. Res. 834, 51–55. posed to 1, 1, 1,-trichloroethane: a matched-pair study. Arch. Environ. Health 33,
Gedalanga, P.B., Pornwongthong, P., Mora, R., Chiang, S., Baldwin, B., Ogles, D., et al., 2014. 331–342.
Identification of biomarker genes to predict biodegradation of 1,4-dioxane. Appl. En- Krishnan, K., Carrier, R., 2008. Approaches for evaluating the relevance of multiroute ex-
viron. Microbiol. 80, 3209–3218. posures in establishing guideline values for drinking water contaminants AU-
Gi, M., Fujioka, M., Kakehashi, A., Okuno, T., Masumura, K., Nohmi, T., et al., 2018. In vivo Krishnan, Kannan. J. Environ. Sci. Health C 26, 300–316.
positive mutagenicity of 1,4-dioxane and quantitative analysis of its mutagenicity Kumar, A., Vyas, G., Bhatt, M., Bhatt, S., Paul, P., 2015. Silver nanoparticle based highly se-
and carcinogenicity in rats. Arch. Toxicol. 92, 3207–3221. lective and sensitive solvatochromatic sensor for colorimetric detection of 1,4-diox-
Göen, T., von Helden, F., Eckert, E., Knecht, U., Drexler, H., Walter, D., 2016. Metabolism ane in aqueous media. Chem. Commun. 51, 15936–15939.
and toxicokinetics of 1,4-dioxane in humans after inhalational exposure at rest and Lan, R.S., Smith, C.A., Hyman, M.R., 2013. Oxidation of cyclic ethers by alkane-grown my-
under physical stress. Arch. Toxicol. 90, 1315–1324. cobacterium vaccae JOB5. Remediat. J. 23, 23–42.
Goldsworthy, T.L., Popp, J.A., 1987. Chlorinated hydrocarbon-induced peroxisomal en- Landry, G.M., Martin, S., McMartin, K.E., 2011. Diglycolic acid is the nephrotoxic metabo-
zyme activity in relation to species and organ carcinogenicity. Toxicol. Appl. lite in diethylene glycol poisoning inducing necrosis in human proximal tubule cells
Pharmacol. 88, 225–233. in vitro. Toxicol. Sci. 124, 35–44.
Goldsworthy, T.L., Monticello, T.M., Morgan, K.T., Bermudez, E., Wilson, D.M., Jackh, R., et Landry, G.M., Dunning, C.L., Abreo, F., Latimer, B., Orchard, E., McMartin, K.E., 2015.
al., 1991. Examination of potential mechanisms of carcinogenicity of 1,4-dioxane in Diethylene glycol-induced toxicities show marked threshold dose response in rats.
rat nasal epithelial cells and hepatocytes. Arch. Toxicol. 65, 1–9. Toxicol. Appl. Pharmacol. 282, 244–251.
Grostern, A., Sales, C.M., Zhuang, W.Q., Erbilgin, O., Alvarez-Cohen, L., 2012. Glyoxylate Larranaga, M.D., Lewis, R.J., Lewis, R.A., 2016. Hawley's Condensed Chemical Dictionary.
metabolism is a key feature of the metabolic degradation of 1,4-dioxane by John Wiley & Sons, Hoboken, NJ.
Pseudonocardia dioxanivorans strain CB1190. Appl. Environ. Microbiol. 78, Lee, K.C., Beak, H.J., Choo, K.H., 2015. Membrane photoreactor treatment of 1,4-dioxane-
3298–3308. containing textile wastewater effluent: performance, modeling, and fouling control.
Hamaguchi, N., Ellington, A., Stanton, M., 2001. Aptamer beacons for the direct detection Water Res. 86, 58–65.
of proteins. Anal. Biochem. 294, 126–131. Lesage, S., Jackson, R.E., Priddle, M.W., Riemann, P.G., 1990. Occurrence and fate of organic
Hansen, J.A., Wang, J., Kawde, A.N., Xiang, Y., Gothelf, K.V., Collins, G., 2006. Quantum-dot/ solvent residues in anoxic groundwater at the Gloucester landfill, Canada. Environ.
aptamer-based ultrasensitive multi-analyte electrochemical biosensor. J. Am. Chem. Sci. Technol. 24, 559–566.
Soc. 128, 2228–2229. Li, M., Fiorenza, S., Chatham, J.R., Mahendra, S., Alvarez, P.J., 2010. 1,4-Dioxane biodegra-
Health Canada, 2018. 1,4-Dioxane in Drinking Water. Guideline Technical Document for dation at low temperatures in Arctic groundwater samples. Water Res. 44,
Public Consultation. Government of Canada, Federal-Provincial-Territorial Committee 2894–2900.
on Drinking Water, Ottawa, Ontario. Li, W., Xu, E., Schlenk, D., Liu, H., 2018. Cyto- and geno-toxicity of 1,4-dioxane and its
Health Canada and Environment Canada, 2010. Screening Assessment for the Challenge transformation products during ultraviolet-driven advanced oxidation processes. En-
-1,4-Dioxane. Environment Canada, Health Canada, Ottawa, Ontario. viron Sci Water Res Technol 4, 1213–1218.
Heuvelsland, A.J., 1988. Method for Producing 1,4-Dioxane. US4764626A. Dow Chemical Linares, R.V., Yangali-Quintanilla, V., Li, Z., Amy, G., 2011. Rejection of micropollutants by
Nederland BV, Dow Chemical Co, United States of America. clean and fouled forward osmosis membrane. Water Res. 45, 6737–6744.
Holthoff, E.L., Marcus, L.S., Pellegrino, P.M., 2015. Toward the realization of a compact Lippincott, D., Streger, S.H., Schaefer, C.E., Hinkle, J., Stormo, J., Steffan, R.J., 2015. Bioaug-
chemical sensor platform using quantum cascade lasers. Proceedings of the SPIE. mentation and propane biosparging for in situ biodegradation of 1,4-dioxane. Ground
9467. U.S. Army Research Lab. Water Monit Remediat 35, 81–92.
Huang, H., Shen, D., Li, N., Shan, D., Shentu, J., Zhou, Y., 2014. Biodegradation of 1,4-diox- Liss, M., Petersen, B., Wolf, H., Prohaska, E., 2002. An aptamer-based quartz crystal protein
ane by a novel strain and its biodegradation pathway. Water Air Soil Pollut. 225, biosensor. Anal. Chem. 74, 4488–4495.
2135–2146. Lundberg, I., Hogberg, J., Kronevi, T., Holmberg, B., 1987. Three industrial solvents investi-
Hwangbo, M., Claycomb, E.C., Liu, Y., Alivio, T.E.G., Banerjee, S., Chu, K.H., 2019. Effec- gated for tumor promoting activity in the rat liver. Cancer Lett. 36, 29–33.
tiveness of zinc oxide-assisted photocatalysis for concerned constituents in Ma, D., Li, B., Cui, Z., Liu, K., Chen, C., Li, G., et al., 2016. Multifunctional luminescent porous
reclaimed wastewater: 1,4-dioxane, trihalomethanes, antibiotics, antibiotic re- organic polymer for selectively detecting iron ions and 1,4-dioxane via luminescent
sistant bacteria (ARB), and antibiotic resistance genes (ARGs). Sci. Total Environ. turn-off and turn-on sensing. ACS Appl. Mater. Interfaces 8, 24097–24103.
649, 1189–1197. Mahendra, S., Alvarez-Cohen, L., 2005. Pseudonocardia dioxanivorans sp nov., a novel ac-
Inoue, O., Seiji, K., Kawai, T., Jin, C., Liu, Y.T., Chen, Z., et al., 1989. Relationship between tinomycete that grows on 1,4-dioxane. Int. J. Syst. Evol. Microbiol. 55, 593–598.
vapor exposure and urinary metabolite excretion among workers exposed to trichlo- Mahendra, S., Alvarez-Cohen, L., 2006. Kinetics of 1,4-dioxane biodegradation by
roethylene. Am. J. Ind. Med. 15, 103–110. monooxygenase-expressing bacteria. Environ Sci Technol 40, 5435–5442.
Isaacson, C., Mohr, T.K.G., Field, J.A., 2006. Quantitative determination of 1,4-dioxane and Mahendra, S., Petzold, C.J., Baidoo, E.E., Keasling, J.D., Alvarez-Cohen, L., 2007. Identifica-
tetrahydrofuran in groundwater by solid phase extraction GC/MS/MS. Environ. Sci. tion of the intermediates of in vivo oxidation of 1,4-Dioxane by monooxygenase-
Technol. 40, 7305–7311. containing Bacteria. Environ. Sci. Technol. 41, 7330–7336.
Itoh, S., Hattori, C., 2019. In vivo genotoxicity of 1,4-dioxane evaluated by liver and bone Makino, R., Kawasaki, H., Kishimoto, A., Gamo, M., Nakanishi, J., 2006. Estimating health
marrow micronucleus tests and Pig-a assay in rats. Mutat. Res. 837, 8–14. risk from exposure to 1,4-dioxane in Japan. Environ. Sci. 13, 43–58.
Jackson, R.E., Dwarakanath, V., 1999. Chlorinated decreasing solvents: physical-chemical Mameda, N., Park, H.J., Choo, K.H., 2017. Membrane electro-oxidizer: a new hybrid mem-
properties affecting aquifer contamination and remediation. Ground Water Monit. brane system with electrochemical oxidation for enhanced organics and fouling con-
Rem. 19, 102–110. trol. Water Res. 126, 40–49.
Jasmann, J.R., Borch, T., Sale, T.C., Blotevogel, J., 2016. Advanced electrochemical oxidation Mohr, T., 2001. 1,4-Dioxane and Other Solvent Stabilizers. Santa Clara Valley Water Dis-
of 1,4-dioxane via dark catalysis by novel titanium dioxide (TiO2) pellets. Environ. trict, Santa Jose, CA.
Sci. Technol. 50, 8817–8826. Morita, T., Hayashi, M., 1998. 1,4-Dioxane is not mutagenic in five in vitro assays and
Kano, H., Umeda, Y., Kasai, T., Sasaki, T., Matsumoto, M., Yamazaki, K., et al., 2009. Carcino- mouse peripheral blood micronucleus assay, but is in mouse liver micronucleus
genicity studies of 1,4-dioxane administered in drinking-water to rats and mice for assay. Environ. Mol. Mutagen. 32, 269–280.
2 years. Food Chem. Toxicol. 47, 2776–2784. Mulisch, H.-M., Winter, W., Dieter, H.H., 2003. Modular system for total evaluation of en-
Karges, U., Becker, J., Puttmann, W., 2018. 1,4-Dioxane pollution at contaminated ground- vironmental contaminants in soil and water. (Modulares system zur Bewertung von
water sites in western Germany and its distribution within a TCE plume. Sci. Total En- Umweltkontaminanten in Böden und Gewässern). 46. Bundesgesundheitsbl.
viron. 619-620, 712–720. Gesundheitsforsch. Gesundheitsschutz, pp. 668–676.
Kasai, T., Saito, M., Senoh, H., Umeda, Y., Aiso, S., Ohbayashi, H., et al., 2008. Thirteen-week Myers, M.A., Johnson, N.W., Marin, E.Z., Pornwongthong, P., Liu, Y., Gedalanga, P.B., et al.,
inhalation toxicity of 1,4-dioxane in rats. Inhal. Toxicol. 20, 961–971. 2018. Abiotic and bioaugmented granular activated carbon for the treatment of 1,4-
Kasai, T., Kano, H., Umeda, Y., Sasaki, T., Ikawa, N., Nishizawa, T., et al., 2009. Two-year in- dioxane-contaminated water. Environ. Pollut. 240, 916–924.
halation study of carcinogenicity and chronic toxicity of 1,4-dioxane in male rats. Nakamiya, K., Hashimoto, S., Ito, H., Edmonds, J.S., Morita, M., 2005. Degradation of 1,4-di-
Inhal. Toxicol. 21, 889–897. oxane and cyclic ethers by an isolated fungus. Appl. Environ. Microbiol. 71, 1254.
Kegel, F.S., Rietman, B.M., Verliefde, A.R., 2010. Reverse osmosis followed by activated car- Nannelli, A., De Rubertis, A., Longo, V., Gervasi, P.G., 2005a. Effects of dioxane on cyto-
bon filtration for efficient removal of organic micropollutants from river bank filtrate. chrome P450 enzymes in liver, kidney, lung and nasal mucosa of rat. Arch. Toxicol.
Water Sci. Technol. 61, 2603–2610. 79, 74–82.
866 K.J. Godri Pollitt et al. / Science of the Total Environment 690 (2019) 853–866

Nannelli, A., De Rubertis, A., Longo, V., Gervasi, P.G., 2005b. Effects of dioxane on cyto- Sweeney, L.M., Thrall, K.D., Poet, T.S., Corley, R.A., Weber, T.J., Locey, B.J., et al., 2008. Phys-
chrome P450 enzymes in liver, kidney, lung and nasal mucosa of rat. Arch. Toxicol. iologically based pharmacokinetic modeling of 1,4-dioxane in rats, mice, and
79, 74–82. humans. Toxicol. Sci. 101, 32–50.
National Cancer Institute, 1978. Bioassay of 1, 4-Dioxane for Possible Carcinogenicity Be- Tahara, M., Obama, T., Ikarashi, Y., 2013. Development of analytical method for determi-
thesda, MD. nation of 1,4-dioxane in cleansing products. Int. J. Cosmet. Sci. 35, 575–580.
NICNAS, 1998. 1,4-Dioxane. Priority Existing Chemical Assessment Report No. 7. National Tanabe, A., Kawata, K., 2008. Determination of 1,4-dioxane in household detergents and
Occupational Health and Safety Commission, Commonwealth of Australia, Canberra, cleaners. J. AOAC Int. 91, 439–444.
ACT. Tanabe, A., Tsuchida, Y., Ibaraki, T., Kawata, K., 2006. Impact of 1,4-dioxane from domestic
Nishimura, T., Iizuka, S., Kibune, N., Ando, M., 2004. Study of 1,4-dioxane intake in the effluent on the Agano and Shinano Rivers, Japan. Bull. Environ. Contam. Toxicol. 76,
total diet using the market-basket method. J. Health Sci. 50, 101–107. 44–51.
NTP, 1992. National Toxicology Program Chemical Repository Database. Institute of Envi- Thiess, A.M., T, E., F, I., 1976. Arbeitsmedizinische Untersuchungsergebuisse von Dioxan-
ronmental Health Sciences, National Institutes of Health, Research Triangle Park, NC. exponierten Mitarbeitern. Arbeitsmedizin Sozialmedizin Praventivmedizin,
O'Neil, M.J., 2013. The Merck Index: An Encyclopedia of Chemicals, Drugs, and Biologicals. pp. 36–46.
Royal Society of Chemistry, Cambridge, UK. US EPA, 2015. The Third Unregulated Contaminant Monitoring Rule (UCMR 3): Searching
Parales, R.E., Adamus, J.E., White, N., May, H.D., 1994. Degradation of 1,4-dioxane by an ac- for Emerging Contaminants in Drinking Water. https://www.epa.gov/sites/produc-
tinomycete in pure culture. Appl. Environ. Microbiol. 60, 4527–4530. tion/files/2015-10/documents/ucmr3_factsheet_general.pdf.
Park, Y.M., Pyo, H., Park, S.J., Park, S.K., 2005. Development of the analytical method for Vainberg, S., McClay, K., Masuda, H., Root, D., Condee, C., Zylstra, G.J., et al., 2006. Biodeg-
1,4-dioxane in water by liquid–liquid extraction. Anal. Chim. Acta 548, 109–115. radation of ether pollutants by Pseudonocardia sp. strain ENV478. Appl. Environ.
Pharmacopeial Convention US, 2008. Food Chemicals Codex. Rockville, Maryland. . Microbiol. 72, 5218–5224.
Postigo, C., Barcelo, D., 2015. Synthetic organic compounds and their transformation Verschueren, K., 2008. Handbook of Environmental Data on Organic Chemicals. John
products in groundwater: occurrence, fate and mitigation. Sci. Total Environ. 503- Wiley & Sons, New York, NY.
504, 32–47. Wacławek, S., Lutze, H.V., Grübel, K., Padil, V.V.T., Černík, M., Dionysiou, D.D., 2017. Chem-
Priddle, M.W., Jackson, R.E., 1991. Laboratory column measurement of VOC retardation istry of persulfates in water and wastewater treatment: a review. Chem. Eng. J. 330,
factors and comparison with field values. Groundwater 29, 260–266. 44–62.
Rattray, N.J.W., Deziel, N.C., Wallach, J.D., Khan, S.A., Vasiliou, V., Ioannidis, J.P.A., et al., WHO, 2005. 1,4-Dioxane in Drinking-water. Background Document for Development of
2018. Beyond genomics: understanding exposotypes through metabolomics. Hum. WHO Guidelines for Drinking-water Quality. World Health Organization, Geneva,
Genomics 12, 4. Switzerland.
Roy, S.K., Thilagar, A.K., Eastmond, D.A., 2005. Chromosome breakage is primarily respon- WHO, 2011. Guidelines for Drinking-water Quality. World Health Organization, Geneva,
sible for the micronuclei induced by 1,4-dioxane in the bone marrow and liver of Switzerland.
young CD-1 mice. Mutat. Res. 586, 28–37. Wilbur, S., Jones, D., Risher, J.F., Crawford, J., Tencza, B., Llados, F., et al., 2012. Toxicological
Schier, J.G., Hunt, D.R., Perala, A., McMartin, K.E., Bartels, M.J., Lewis, L.S., et al., 2013. Char- Profile for 1,4-Dioxane, Atlanta (GA).
acterizing concentrations of diethylene glycol and suspected metabolites in human Woo, Y.-t., Arcos, J.C., Argus, M.F., Griffin, G.W., Nishiyama, K., 1977. Metabolism in vivo of
serum, urine, and cerebrospinal fluid samples from the Panama DEG mass poisoning. dioxane: identification of p-dioxane-2-one as the major urinary metabolite. Biochem.
Clin Toxicol (Phila) 51, 923–929. Pharmacol. 26, 1535–1538.
Schriks, M., Heringa, M.B., van der Kooi, M.M.E., de Voogt, P., van Wezel, A.P., 2010. Tox- Woo, Y.T., Neuburger, B.J., Arcos, J.C., Argus, M.F., Nishiyama, K., Griffin, G.W., 1980. En-
icological relevance of emerging contaminants for drinking water quality. Water Res. hancement of toxicity and enzyme-repressing activity of p-dioxane by chlorination:
44, 461–476. stereoselective effects. Toxicol. Lett. 5, 69–75.
Shen, W., Chen, H., Pan, S., 2008. Anaerobic biodegradation of 1,4-dioxane by sludge Yamamoto, N., Inoue, D., Sei, K., Saito, Y., Ike, M., 2018. Field test of on-site treatment of
enriched with iron-reducing microorganisms. Bioresour. Technol. 99, 2483–2487. 1,4-dioxane-contaminated groundwater using Pseudonocardia sp. D17. J. Water En-
Son, H.S., Im, J.K., Zoh, K.D., 2009. A Fenton-like degradation mechanism for 1,4-dioxane viron. Technol. 16, 256–268.
using zero-valent iron (Fe0) and UV light. Water Res. 43, 1457–1463. Young, J.D., Braun, W.H., Gehring, P.J., Horvath, B.S., Daniel, R.L., 1976. 1,4-Dioxane and
Song, D., Zhang, S., 1997. Rapid determination of 1,4-dioxane in water by solid-phase ex- beta-hydroxyethoxyacetic acid excretion in urine of humans exposed to dioxane va-
traction and gas chromatography-mass spectrometry. J. Chromatogr. A 787, 283–287. pors. Toxicol. Appl. Pharmacol. 38, 643–646.
Stefania, G., Maura, B., Claudia, V.L., Barbara, P., Ginevra, M., Francesco, R., 2009. Biological Young, J.D., Braun, W.H., Rampy, L.W., Chenoweth, M.B., Blau, G.E., 1977. Pharmacokinet-
effects of diethylene glycol (DEG) and produced waters (PWs) released from offshore ics of 1,4-dioxane in humans. J. Toxicol. Environ. Health 3, 507–520.
activities: a multi-biomarker approach with the sea bass Dicentrarchus labrax. Envi- Zhang, Z.Z., Low, P.F., Cushman, J.H., Roth, C.B., 1990. Adsorption and heat of adsorption of
ron. Pollut. 157, 3166–3173. organic compounds on montmorillonite from aqueous solutions. Soil Sci. Soc. Am. J.
Stepien, D.K., Püttmann, W., 2013. Simultaneous determination of six hydrophilic ethers 54, 59–66.
at trace levels using coconut charcoal adsorbent and gas chromatography/mass spec- Zhang, S., Gedalanga, P.B., Mahendra, S., 2017. Advances in bioremediation of 1,4-diox-
trometry. Anal. Bioanal. Chem. 405, 1743–1751. ane-contaminated waters. J. Environ. Manag. 204, 765–774.
Stepien, D.K., Diehl, P., Helm, J., Thoms, A., Puttmann, W., 2014. Fate of 1,4-dioxane in the Zhang, M., Li, G., Zhou, Q., Pan, D., Zhu, M., Xiao, R., et al., 2018. Boosted electrochemical
aquatic environment: from sewage to drinking water. Water Res. 48, 406–419. immunosensing of genetically modified crop markers using nanobody and mesopo-
Stickney, J.A., Sager, S.L., Clarkson, J.R., Smith, L.A., Locey, B.J., Bock, M.J., et al., 2003. An up- rous carbon. ACS Sens 3, 684–691.
dated evaluation of the carcinogenic potential of 1,4-dioxane. Regul. Toxicol. Zhong, H., Brusseau, M.L., Wang, Y., Yan, N., Quig, L., Johnson, G.R., 2015. In-situ activation
Pharmacol. 38, 183–195. of persulfate by iron filings and degradation of 1,4-dioxane. Water Res. 83, 104–111.
Stott, W.T., Quast, J.F., Watanabe, P.G., 1981. Differentiation of the mechanisms of oncoge- Zhou, J., Li, H., Zhang, H., Li, H., Shi, W., Cheng, P., 2015. A bimetallic lanthanide metal-
nicity of 1,4-dioxane and 1,3-hexachlorobutadiene in the rat. Toxicol. Appl. organic material as a self-calibrating color-gradient luminescent sensor. Adv. Mater.
Pharmacol. 60, 287–300. 27, 7072–7077.
Sun, M., Lopez-Velandia, C., Knappe, D.R.U., 2016. Determination of 1,4-dioxane in the
cape fear river watershed by heated purge-and-trap preconcentration and gas
chromatography–mass spectrometry. Environ. Sci. Technol 50, 2246–2254.

You might also like