54
WHO/HEP/ECH/WSH/2020.10 Trichloroethene in drinking-water Background document for development of WHO Guidelines for drinking-water quality This document replaces document reference number WHO/SDE/WSH/05.08/22

Trichloroethene in drinking-water - WHO

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Trichloroethene in drinking-water - WHO

WHO/HEP/ECH/WSH/2020.10

Trichloroethene in drinking-water

Background document for development of

WHO Guidelines for drinking-water quality

This document replaces document reference number WHO/SDE/WSH/05.08/22

Page 2: Trichloroethene in drinking-water - WHO

WHO/HEP/ECH/WSH/2020.10

© World Health Organization 2020

Some rights reserved. This work is available under the Creative Commons Attribution-

NonCommercial-ShareAlike 3.0 IGO licence (CC BY-NC-SA 3.0 IGO; https://creativecommons.org/

licenses/by-nc-sa/3.0/igo).

Under the terms of this licence, you may copy, redistribute and adapt the work for non-commercial

purposes, provided the work is appropriately cited, as indicated below. In any use of this work, there

should be no suggestion that WHO endorses any specific organization, products or services. The use of

the WHO logo is not permitted. If you adapt the work, then you must license your work under the same

or equivalent Creative Commons licence. If you create a translation of this work, you should add the

following disclaimer along with the suggested citation: “This translation was not created by the World

Health Organization (WHO). WHO is not responsible for the content or accuracy of this translation.

The original English edition shall be the binding and authentic edition”.

Any mediation relating to disputes arising under the licence shall be conducted in accordance with the

mediation rules of the World Intellectual Property Organization (http://www.wipo.int/amc/en/

mediation/rules/).

Suggested citation. Trichloroethene in drinking-water. Background document for development of

WHO Guidelines for drinking-water quality. Geneva: World Health Organization; 2020

(WHO/HEP/ECH/WSH/2020.10). Licence: CC BY-NC-SA 3.0 IGO.

Cataloguing-in-Publication (CIP) data. CIP data are available at http://apps.who.int/iris.

Sales, rights and licensing. To purchase WHO publications, see http://apps.who.int/bookorders. To

submit requests for commercial use and queries on rights and licensing, see

http://www.who.int/about/licensing.

Third-party materials. If you wish to reuse material from this work that is attributed to a third party,

such as tables, figures or images, it is your responsibility to determine whether permission is needed for

that reuse and to obtain permission from the copyright holder. The risk of claims resulting from

infringement of any third-party-owned component in the work rests solely with the user.

General disclaimers. The designations employed and the presentation of the material in this

publication do not imply the expression of any opinion whatsoever on the part of WHO concerning the

legal status of any country, territory, city or area or of its authorities, or concerning the delimitation of

its frontiers or boundaries. Dotted and dashed lines on maps represent approximate border lines for

which there may not yet be full agreement.

The mention of specific companies or of certain manufacturers’ products does not imply that they are

endorsed or recommended by WHO in preference to others of a similar nature that are not mentioned.

Errors and omissions excepted, the names of proprietary products are distinguished by initial capital

letters.

All reasonable precautions have been taken by WHO to verify the information contained in this

publication. However, the published material is being distributed without warranty of any kind, either

expressed or implied. The responsibility for the interpretation and use of the material lies with the

reader. In no event shall WHO be liable for damages arising from its use.

Page 3: Trichloroethene in drinking-water - WHO

iii

Preface

Access to safe drinking-water is essential to health, a basic human right and a component of effective

policy for health protection. A major World Health Organization (WHO) function to support access to

safe drinking-water is the responsibility “to propose ... regulations, and to make recommendations with

respect to international health matters ...”, including those related to the safety and management of

drinking-water.

The first WHO document dealing specifically with public drinking-water quality was published in 1958

as International standards for drinking-water. It was revised in 1963 and 1971 under the same title. In

1984–1985, the first edition of the WHO Guidelines for drinking-water quality (GDWQ) was published

in three volumes: Volume 1, Recommendations; Volume 2, Health criteria and other supporting

information; and Volume 3, Surveillance and control of community supplies. Second editions of these

volumes were published in 1993, 1996 and 1997, respectively. Addenda to Volumes 1 and 2 of the

second edition were published in 1998, addressing selected chemicals. An addendum on

microbiological aspects, reviewing selected microorganisms, was published in 2002. The third edition

of the GDWQ was published in 2004, the first addendum to the third edition was published in 2006,

and the second addendum to the third edition was published in 2008. The fourth edition was published

in 2011, and the first addendum to the fourth edition was published in 2017.

The GDWQ are subject to a rolling revision process. Through this process, microbial, chemical and

radiological aspects of drinking-water are subject to periodic review, and documentation relating to

aspects of protection and control of drinking-water quality is accordingly prepared and updated.

Since the first edition of the GDWQ, WHO has published information on health criteria and other

information to support the GDWQ, describing the approaches used in deriving guideline values, and

presenting critical reviews and evaluations of the effects on human health of the substances or

contaminants of potential health concern in drinking-water. In the first and second editions, these

constituted Volume 2 of the GDWQ. Since publication of the third edition, they comprise a series of

free-standing monographs, including this one.

For each chemical contaminant or substance considered, a background document evaluating the risks

to human health from exposure to that chemical in drinking-water was prepared. The draft health criteria

document was submitted to a number of scientific institutions and selected experts for peer review. The

draft document was also released to the public domain for comment. Comments were carefully

considered and addressed, as appropriate, taking into consideration the processes outlined in the

Policies and procedures used in updating the WHO guidelines for drinking-water quality and the WHO

Handbook for guideline development. The revised draft was submitted for final evaluation at expert

consultations.

During preparation of background documents and at expert consultations, careful consideration was

given to information available in previous risk assessments carried out by the International Programme

on Chemical Safety, in its Environmental Health Criteria monographs and Concise International

Chemical Assessment Documents; the International Agency for Research on Cancer; the Joint Food

and Agriculture Organization of the United Nations (FAO)/WHO Meeting on Pesticide Residues; and

the Joint FAO/WHO Expert Committee on Food Additives (which evaluates contaminants such as lead,

cadmium, nitrate and nitrite, in addition to food additives).

Further up-to-date information on the GDWQ and the process of their development is available on the

WHO website and in the current edition of the GDWQ.

Page 4: Trichloroethene in drinking-water - WHO

iv

Acknowledgements

The update of this background document on trichloroethene in drinking-water for the development of

the World Health Organization (WHO) Guidelines for drinking-water quality (GDWQ) was led by

Emanuela Testai of the Istituto Superiore di Sanità of Italy. The contributions of Professor John Fawell,

of Cranfield University, United Kingdom, and France Lemieux, Health Canada, who led the update of

section 7 on practical considerations, are gratefully acknowledged.

The work of the following experts was crucial in the development of this document and others in the

second addendum to the fourth edition:

Dr M Asami, National Institute of Public Health, Japan

Dr RJ Bevan, independent consultant, United Kingdom

Mr R Carrier, Health Canada, Canada

Dr J Cotruvo, Joseph Cotruvo & Associates and NSF International WHO Collaborating Centre,

United States of America

Dr D Cunliffe, South Australia Department of Health, Australia

Dr L d’Anglada, Environmental Protection Agency, United States of America

Dr A Eckhardt, Umweltbundesamt (Federal Environment Agency), Germany

Professor JK Fawell, Cranfield University, United Kingdom

Dr A Hirose, National Institute of Health Sciences of Japan

Dr A Humpage, University of Adelaide (formerly South Australian Water Corporation), Australia

Dr P Marsden, Drinking Water Inspectorate, United Kingdom

Professor Y Matsui, Hokkaido University, Japan

Dr E Ohanian, Environmental Protection Agency, United States of America

Professor CN Ong, National University of Singapore, Singapore

Dr J Strong, Environmental Protection Agency, United States of America

Dr E Testai, National Institute of Health, Italy

The draft text was discussed at the expert consultations for the second addendum to the fourth edition

of the GDWQ, held on 28–30 March 2017 and 13–14 July 2018. The final version of the document

takes into consideration comments from both peer reviewers and the public, including N Kobayashi,

National Institute of Health Science, Japan; B Lampe, NSF International, United States of America; G

Miller, Environmental Protection Agency, United States of America; and M Templeton, Imperial

College London, United Kingdom.

The coordinator was Ms J De France, WHO, with support from Dr V Bhat, formerly of NSF

International, United States of America. Strategic direction was provided by Mr B Gordon, WHO. Dr

A Tritscher, formerly of WHO, and Dr P Verger, WHO, provided liaisons with the Joint FAO/WHO

Expert Committee on Food Additives and the Joint FAO/WHO Meeting on Pesticide Residues. Dr R

Brown and Ms C Vickers, WHO, provided liaisons with the International Programme on Chemical

Safety. Dr M Perez contributed on behalf of the WHO Radiation Programme. Dr Andina Faragher,

Biotext, Australia, was responsible for the scientific editing of the document.

Many individuals from various countries contributed to the development of the GDWQ. The efforts of

all who contributed to the preparation of this document are greatly appreciated.

Page 5: Trichloroethene in drinking-water - WHO

v

Acronyms and abbreviations

BMD benchmark dose

BMDL lower 95% confidence limit of the benchmark dose

BMDL01 lower 95% confidence limit on the benchmark dose for a 1% response

bw body weight

CAS Chemical Abstracts Service

CH chloral hydrate

CI confidence interval

CNS central nervous system

CYP cytochrome P450

DCA dichloroacetic acid

DCVC S-dichlorovinyl-L-cysteine

DCVG S-dichlorovinyl glutathione

DNA deoxyribonucleic acid

FAO Food and Agriculture Organization of the United Nations

GAC granular activated carbon

GD gestation day

GDWQ Guidelines for drinking-water quality

GSH glutathione

GST glutathione-S-transferase

GV guideline value

HED human equivalent dose

Leq litre-equivalent

LOAEL lowest-observed-adverse-effect level

NOAEL no-observed-adverse-effect level

OR odds ratio

PBPK physiologically based pharmacokinetic (modelling)

PCE perchloroethylene (tetrachloroethene)

POD point of departure

PPAR peroxisome proliferator activated receptor

RR relative risk

TCA trichloroacetic acid

TCE trichloroethene

TCOG trichloroethanol glucuronide

TCOH trichloroethanol

TDI tolerable daily intake

USA United States of America

US EPA United States Environmental Protection Agency

VHL Von Hippel–Lindau

VOC volatile organic compound

WHO World Health Organization

Page 6: Trichloroethene in drinking-water - WHO
Page 7: Trichloroethene in drinking-water - WHO

vii

Contents

Executive summary .................................................................................................................. 1

1 General description ...................................................................................................... 2

1.1 Identity ............................................................................................................... 2

1.2 Physicochemical properties ............................................................................... 2

1.3 Organoleptic properties ...................................................................................... 2

1.4 Major uses and sources ...................................................................................... 3

2 Environmental levels and human exposure ............................................................... 3

2.1 Water .................................................................................................................. 3

2.2 Food ................................................................................................................... 4

2.3 Air ...................................................................................................................... 5

2.4 Bioaccumulation ................................................................................................ 5

2.5 Occupational exposure ....................................................................................... 5

2.6 Estimated total exposure, biomonitoring studies and relative

contribution of drinking-water ........................................................................... 6

3 Toxicokinetics and metabolism in animals and humans .......................................... 7

3.1 Absorption.......................................................................................................... 7

3.2 Distribution ........................................................................................................ 8

3.3 Metabolism ........................................................................................................ 8

3.4 Elimination ....................................................................................................... 10

3.5 Physiologically based pharmacokinetic modelling .......................................... 10

4 Effects on humans ...................................................................................................... 11

4.1 Acute exposure................................................................................................. 11

4.2 Short-term exposure ......................................................................................... 12

4.3 Long-term exposure ......................................................................................... 12

4.3.1 Systemic effects ................................................................................... 12

4.3.2 Neurological effects ............................................................................. 12

4.3.3 Reproductive and developmental effects ............................................. 12

4.3.4 Immunological effects ......................................................................... 13

4.3.5 Genotoxicity and carcinogenicity ........................................................ 13

5 Effects on experimental animals and in vitro test systems ..................................... 16

5.1 Acute exposure................................................................................................. 16

5.2 Short-term exposure ......................................................................................... 17

Page 8: Trichloroethene in drinking-water - WHO

viii

5.3 Long-term exposure ......................................................................................... 18

5.3.1 Systemic effects ................................................................................... 18

5.3.2 Neurological effects ............................................................................. 18

5.3.3 Reproductive and developmental effects ............................................. 18

5.3.4 Immunological effects ......................................................................... 20

5.3.5 Genotoxicity and carcinogenicity ........................................................ 21

5.4 Mode of action ................................................................................................. 23

6 Overall database and quality of evidence ................................................................ 25

6.1 Summary of health effects ............................................................................... 25

6.2 Quality of evidence .......................................................................................... 27

7 Practical considerations............................................................................................. 27

7.1 Analytical methods and achievability .............................................................. 27

7.2 Source control .................................................................................................. 28

7.3 Treatment methods and performance ............................................................... 28

8 Conclusion .................................................................................................................. 29

8.1 Derivation of the guideline value ..................................................................... 29

8.1.1 Noncancer effects................................................................................. 31

8.1.2 Guideline value .................................................................................... 32

8.2 Considerations in applying the guideline value ............................................... 32

References ............................................................................................................................... 34

Page 9: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

1

Executive summary

Trichloroethene (TCE) is primarily, if not exclusively, a groundwater contaminant, because it

volatilizes to the atmosphere from surface waters. The primary cause of groundwater

contamination is poor handling and disposal practices, which result in soil contamination; in

vulnerable aquifers, soil contamination can result in groundwater contamination. Commercial

utility of TCE is decreasing as a result of increasing regulations.

TCE is a data-rich compound, with many high-quality studies available on kinetics and toxicity

in humans and laboratory animals. These include studies on TCE-induced neurological effects

in humans and animals; effects on kidney, liver and body weight in animals; immunological

effects in animals; reproductive effects in humans and animals; and developmental effects in

animals.

Selection of multiple critical effects, rather than the lowest point of departure, as the basis of

the guideline value (GV) of 8 g/L helped overcome possible limitations of individual studies.

The GV is achievable using currently available treatment technologies. Source control

measures should include improved handling and disposal practices.

TCE monitoring requirements in drinking-water regulations and standards should be limited to

groundwater sources where a catchment risk assessment indicates the possibility of presence

of TCE. Source control should be the primary mitigation measure; however, this is not feasible

where there is historical contamination. Effective treatment techniques include aeration,

including packed tower aeration, and granular activated carbon. Ozone and advanced oxidation

processes with ozone may also be effective. Surface water sources do not need to be monitored

or treated, since TCE volatilizes to the atmosphere.

Page 10: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

2

1 General description

1.1 Identity

Trichloroethene (TCE) is also known as trichloroethylene, acetylene trichloride, 1-chloro-2,2-

dichloroethylene, 1,1-dichloro-2-chloroethylene, ethylene trichloride or 1,1,2-

trichloroethylene.

CAS No.: 79-01-6

Molecular formula: C2HCl3

Chemical structure:

1.2 Physicochemical properties

Table 1.1. Physicochemical properties of trichloroethene

Property Value

Molecular weight 131.39

Boiling point 87.2 °C

Melting point –84.7 °C

Density at 20 °C 1.4642 g/cm3

Vapour density (air = 1) 4.53

Vapour pressure at 25 °C 69 mm Hg

Solubility

Water at 25 °C 1280 mg/L

Organic solvents Soluble in ethanol, diethyl ether, acetone and

chloroform

Partition coefficients

Log Kow 2.61

Log Koc 49–460

Henry’s law constant at 25 °C 9.85 × 10–3 atm-m3/mol

Note: Conversion factors: 1 ppm = 5.46 mg/m3; 1 mg/m3 = 0.18 ppm (ATSDR, 2019)

Source: ATSDR (2019)

1.3 Organoleptic properties

TCE is a liquid with a sweet ether-like and chloroform-like odour. The odour thresholds for

TCE are 546–1092 mg/m3 in air and 0.31 mg/L in water (Amoore & Hautala, 1983; Ruth,

1986).

Page 11: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

3

1.4 Major uses and sources

TCE is used primarily in metal degreasing operations. It is also used as a solvent for greases,

oils, fats and tars; in paint removers, coatings and vinyl resins; and by the textile processing

industry to scour cotton, wool and other fabrics. Historically, the most important use of TCE

has been vapour degreasing of metal parts in the automotive and metals industries. This use

has been declining since the 1990s, as a result of increased environmental regulations

governing TCE emissions (ATSDR, 2019). For example, the use of TCE as a solvent in Europe

dropped by 85% from 1984 to 2006, and by a further 60% from 2006 to 2010 (ECSA, 2012;

IARC, 2014).

Currently, the main use of TCE is as a feedstock material to produce other chemicals, such as

fluorinated hydrocarbons and fluorinated polymers, which are being phased out under the

Montreal Protocol on Substances that Deplete the Ozone Layer. About 80% of current

production in the European Union is used for this purpose (ECSA, 2012). TCE may be used as

a chemical intermediate in the production of polyvinyl chloride, flame-retardant chemicals and

insecticides.

Most of the TCE used for degreasing is believed to be emitted to the atmosphere (US EPA,

1985a). TCE may also be introduced into surface water and groundwater in industrial effluents

(IPCS, 1985). Poor handling, and improper disposal of TCE in landfills, have been the main

causes of groundwater contamination. Biodegradation of another volatile organic pollutant,

tetrachloroethene (also called perchloroethylene, PCE), in groundwater may also lead to the

formation of TCE (Major, Hodgins & Butler, 1991).

2 Environmental levels and human exposure

TCE is widely distributed in the environment as a result of industrial emissions.

Potential environmental exposure to TCE in the air, rainwater, surface waters and drinking-

water has been reviewed (IARC, 2014; ATSDR, 2019). The partitioning tendency of TCE in

the environment has been estimated as follows: air, 97.7%; water, 0.3%; soil, 0.004%; sediment,

0.004% (Boutonnet et al., 1998).

TCE in the atmosphere is highly reactive and persists for an estimated half-life of 6.8 days. It

is transformed in the atmosphere by reaction with photochemically produced hydroxyl radicals

(ATSDR, 2019).

In surface water, volatilization is the principal route of degradation; photodegradation and

hydrolysis play minor roles. In groundwater, TCE is degraded slowly by microorganisms.

2.1 Water

TCE has been detected frequently in natural water and drinking-water in various countries

(IARC, 2014). Because of the high volatility of TCE, it is normally present at low or

undetectable concentrations in surface water (≤1 μg/L; Health Canada, 2005). A 2000 survey

of 68 First Nations community water supplies (groundwater and surface water) in Manitoba,

Canada, found that TCE concentrations were undetectable (<0.5 μg/L) (Yuen & Zimmer, 2001).

The United States Environmental Protection Agency (US EPA, 2014) noted that very low

concentrations of TCE are anticipated in surface water, based on TCE releases to water and

wastewater treatment reported to the Toxics Release Inventory, as well as the fate of TCE in

wastewater treatment.

Page 12: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

4

The major route of removal of TCE from water is volatilization. The estimated volatilization

half-life is 1.2 hours from a model river (1 m deep, flowing at 1 m/second, with a wind velocity

of 5 m/second) and 4.6 days from a model lake (1 m deep, flowing at 0.05 m/second, with a

wind velocity of 0.5 m/second) (US EPA, 2010). However, in groundwater systems where

volatilization and biodegradation are limited, concentrations are higher if contamination has

occurred in the vicinity and leaching has taken place.

Data from Canada – New Brunswick (1994–2001), Alberta (1998–2001), Yukon (2002),

Ontario (1996–2001) and Quebec (1985–2002) – for raw water (surface water and

groundwater), and for treated and distributed water indicated that more than 99% of samples

contained TCE at concentrations ≤1.0 μg/L. Most samples with detectable TCE concentrations

were from groundwater, with the highest concentration being 81 μg/L (Alberta Department of

Environmental Protection, New Brunswick Department of Health and Wellness, Ontario

Ministry of Environment and Energy, Yukon Department of Health and Social Services and

Quebec Ministry of the Environment, personal communications, 2002). In England and Wales,

for about 5000 raw water (surface water and groundwater) samples taken in 2017 from about

800 abstraction points, the mean concentration of TCE was 0.55 µg/L and the maximum was

17.4 µg/L (P. Marsden, UK Drinking Water Inspectorate, personal communication, 28 August

2018). Also in England and Wales, for more than 11 000 drinking-water samples analysed in

2003, the mean concentration of TCE was 0.39 µg/L and the maximum was 21.8 µg/L (P.

Marsden, personal communication, 28 Aug 2018).

Contamination of drinking-water supplies with TCE varies with location and with the drinking-

water source:

• Contamination is more likely in locations with relevant industrial activities, and improper

handling and disposal.

• Generally higher levels of TCE are expected in groundwater because of the lack of

volatilization that occurs compared with surface water.

Because analytical methods have improved since TCE was first assayed, concentrations that

were once considered “nondetectable” are now quantifiable. This confounds the use of

historical TCE data, because the values for “nondetectable” have changed over time. Since the

use of TCE continues to decrease, more recent data on the concentration of TCE in drinking-

water is required to provide an accurate assessment of human exposure to TCE via drinking-

water and its contribution to the total body burden.

2.2 Food

The daily intakes of TCE in food for Canadian adults (20–70 years old) and children (5–

11 years old) were estimated to range from 0.004 to 0.01 μg/kg body weight (bw)/day and from

0.01 to 0.04 μg/kg bw/day, respectively (Canadian Department of National Health and Welfare,

1993). These numbers were based on TCE concentrations from food surveys in the United

States of America from the mid- to late 1980s, as well as Canadian food consumption data. In

recent decades, the severe restrictions on the use of TCE in North America and Europe suggest

that levels in food have been decreasing.

As part of the Total Diet Study in the USA, TCE was found in 30 out of 70 (43%) food items

purchased in supermarkets or restaurants in 1996–2000 at concentrations in the low

microgram-per-kilogram range (Fleming-Jones & Smith, 2003). Food was sampled four times

per year on a regional basis over a 5-year period. Based on 20 samples of each food item, PCE

Page 13: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

5

was most frequently detected in raw avocado (n = 6; 2–75 μg/kg). Potato chips (n = 4; 4–

140 μg/kg) had the highest level of TCE detection, and beef frankfurters (n = 5; 2–105 μg/kg)

had the second most frequent and second highest level of detection. Potential sources of the

contamination were not investigated (Fleming-Jones & Smith, 2003).

Among 17 samples of brown grease from grease traps in food preparation facilities, TCE was

detected in three of the samples, with a mean TCE concentration of 321.3 μg/L (range 146–

600 μg/L (Ward, 2012).

2.3 Air

TCE has been detected worldwide in outdoor and indoor air. In the USA, the results of

1200 measurements in 25 states suggest a general downward trend in mean concentrations of

TCE in air, from about 1.5 μg/m3 in the late 1980s to 0.8 μg/m3 in the late 1990s (IARC, 2014).

TCE concentrations in air have continued to decrease steadily in the USA: an analysis by

McCarthy et al. (2007) of Air Quality System data over three trend periods (1990–2005, 1995–

2005, and 2000–2005) suggested a decrease of about 4–7% for median trichloroethylene levels

annually. Data available on ambient air measurements obtained from EPA’s Air Quality

System database, as reported by ATSDR (2019), indicate that, during the period 2010–2018,

annual mean 50th percentile TCE airborne concentrations from various sampling sites across

the USA ranged from 0 to 0.021 μg/m3. The mean 95th percentile TCE airborne concentrations

across all sampling sites in 2002 and 2018 were 0.25 μg/m3 and 0.0128 μg/m3, respectively

(ATSDR, 2019).

Data on TCE concentrations in air measured in different remote, rural, suburban and urban sites

indicate a similar decreasing trend (IARC, 2014). Concentrations in urban air and in

commercial/industrial areas were about three times higher than in rural areas (Wu & Schaum,

2000).

Modelling suggests that concentrations of TCE in indoor air can increase when TCE-

contaminated water is used domestically – for example, during showering (Ömür-Özbek,

Gallagher & Dietrich, 2011).

Brenner (2010) measured median and maximum TCE concentrations of 0.895 and 1.69 μg/m3

(0.16 and 0.31 ppb), respectively, for 541 indoor air samples from four large buildings at the

southern end of San Francisco Bay. The levels were attributed to vapour intrusion from

underlying contaminated groundwater and soil (US EPA, 2011c; Burk & Zarus, 2013).

2.4 Bioaccumulation

Bioconcentration of TCE in aquatic species is low, with bioconcentration factor values ranging

between 3 and 100 in aquatic organisms (ATSDR, 2019) and some plants (Schroll et al., 1994).

2.5 Occupational exposure

The great majority of data regarding worker exposure to TCE were obtained from degreasing

operations, which is the primary industrial use of TCE (ATSDR, 2019).

Worker exposure is likely to vary, although in most workplaces TCE concentration is regulated

by time-weighted averages (TWA). The United States Occupational Safety and Health

Administration allows an 8-hour TWA permissible exposure limit of 100 ppm and a 15-minute

TWA exposure of 300 ppm, which should not be exceeded at any time during a work day

(OSHA, 1993; Rosa, 2003).

Page 14: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

6

Worker exposure in the dry-cleaning industry is a notable route for exposure to TCE. This is

generally evaluated using the relationship between concentrations of TCE in urine and

concentrations in air collected in the breathing zone of workers in the workplace. In one study

comparing exposed and non-exposed workers in a dry-cleaning centre, the mean values for

exposure to TCE in the breathing zone were 1.56, 1.75 and 2.40 mg/m3 (0.28, 0.32 and

0.43 ppm based on the conversation factor in table 1) for sites with dry-cleaning machine

capacities of 8, 12 and 18 kg, respectively. The mean value for exposure to TCE in the

breathing zone for the occupationally non-exposed participants was 0.98 mg/m3 (0.18 ppm).

Mean urinary concentrations before and after work shifts were measured. Levels before work

were 2.38, 5.53 and 8.18 μg/L (ppb), and levels after work were 4.46, 11.31 and 4.46 μg/L (ppb)

at sites with dry-cleaning machine capacities of 8, 12 and 18 kg, respectively. For

occupationally non-exposed participants, levels were 0.31 μg/L (ppb) before work and

0.29 μg/L (ppb) after work (Rastkari, Yunesian & Ahmadkhaniha, 2011).

2.6 Estimated total exposure, biomonitoring studies and relative contribution of

drinking-water

Most people are exposed to TCE through drinking-water or air. Exposure is likely to have

decreased in North America and Europe as a result of restricted use of TCE in the past several

decades in these regions. TCE has been detected in human body fluids such as blood (Brugnone

et al., 1994; Skender et al., 1994) and breast milk (Pellizzari, Hartwell & Harris, 1982). Several

studies have examined blood concentrations of TCE in the general population. The number of

individuals with measurable concentrations of TCE is generally low and has declined in recent

years (IARC, 2014).

In the United States National Health and Nutrition Examination Survey 1999–2000, blood

samples were taken from 290 subjects; 88% of samples were below the limit of detection of

TCE, and the mean TCE concentration in the positive samples was 0.013 μg/L. In an update of

this survey, for 2001–2014, blood concentrations were usually below the limit of detection of

0.012 ng/mL for 17,419 subjects from the USA general population, including different ethnic

groups and age groups. The most recent data reported include results from 923 cigarette

smokers and 2,054 nonsmokers within the USA general population surveyed during 2013 and

2014: again, the levels were below the limit of detection (ATSDR, 2019).

Exposure of the general population from air, water and food was several orders of magnitude

lower than occupational exposure.

As a result of the volatility and lipid solubility of TCE, exposure can also occur dermally and

through inhalation, especially through bathing and showering (Krishnan & Carrier, 2008).

These indirect exposures are evaluated in terms of litre-equivalents per day (Leq/day). For

example, an inhalation exposure of 1.7 Leq/day means that the daily exposure to TCE via

inhalation is equivalent to a person drinking an extra 1.7 L of water per day. The use of Leq as

a metric of exposure is the most appropriate approach for systemically acting contaminants that

do not exhibit portal-of-entry effects but are likely to induce the same adverse effect by various

exposure routes (Krishnan & Carrier, 2008).

McKone (1987) has estimated that the indoor-air exposure attributable to tap water is 1.5–

6 times the exposure attributable to the consumption of 2 L/day of tap water. Bogen et al. (1988)

proposed lifetime Leq/day values for 70 kg adults of 2.2 (ingestion), 2.9 (inhalation) and 2

(dermal). Weisel & Jo (1996) have reported that approximately equivalent amounts can enter

the body by inhalation, dermal absorption and ingestion. Lindstrom & Pleil (1996) calculated,

Page 15: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

7

using a TCE concentration of 4.4 μg/L in water, that the ingested dose was more important

than the inhaled dose for a 10-minute shower, which, in turn, was greater than the dermal dose.

Krishnan (2003) determined Leq/day values for dermal and inhalation exposures of adults and

children to TCE (5 μg/L) in drinking-water on the basis of the methodological approach of

Lindstrom & Pleil (1996), the use of physiologically based pharmacokinetic models and

consideration of the fraction absorbed (Laparé, Tardif & Brodeur, 1995; Lindstrom & Pleil,

1996; Poet et al., 2000). Bathing in water for 30 minutes resulted in 2.39 Leq (1.67 inhalation

and 0.72 dermal) for adults. In Japan, the median indoor-air exposure to TCE attributable to

tap water for a Japanese lifestyle was estimated to be 3.1 Leq/day (Akiyama et al., 2018). Thus,

the indirect exposure rates depend on exposure scenarios, such as the duration and frequency

of showering. These scenarios are associated with local lifestyles. Since most people do not

take a daily 30-minute bath, the values here are considered to be conservative. Overall, indirect

exposure attributable to tap water may equal direct exposure from water intake. However,

estimates could be improved by considering bioavailability, target tissue dose, and extent of

absorption via all routes and media (Krishnan & Carrier, 2013).

3 Toxicokinetics and metabolism in animals and humans

3.1 Absorption

TCE is readily absorbed following both oral and inhalation exposure. Dermal absorption is also

possible, but information on this route of exposure is limited. Significant variability between

and within species in TCE absorption following all routes of exposure has been well

documented.

In animals, TCE is rapidly and extensively absorbed from the gastrointestinal tract into the

systemic circulation. Mass balance studies using radiolabelled TCE indicated that mice and

rats metabolized TCE at 38–100% and 15–100%, respectively, following oral administration

in corn oil vehicle. For both species, the lower values were obtained following treatment with

large doses, in excess of 1000 mg/kg bw, implying that the rate of absorption was higher at low

doses than at high doses (Daniel, 1963; Parchman & Magee, 1982; Dekant & Henschler, 1983;

Dekant, Metzler & Henschler, 1984; Buben & O’Flaherty, 1985; Mitoma et al., 1985; Prout,

Provan & Green, 1985; Rouisse & Chakrabarti, 1986). Different vehicles affect the rate of

absorption: the rate is almost 15 times greater following dosing in water than following dosing

in corn oil. Overall, absorption of TCE through the gastrointestinal tract is considerable and, at

very low concentration, nearly complete.

Although human exposure studies investigating oral absorption of TCE were not identified,

numerous case studies of accidental or intentional ingestion of TCE suggest that absorption

from the gastrointestinal tract in humans is likely to be extensive (Kleinfeld & Tabershaw,

1954; DeFalque, 1961; Brüning et al., 1998). Following ingestion accidents, TCE and its

metabolites were reported in blood and/or urine at the first sampling times after exposure, the

earliest of which was 13 hours, with peak amounts in blood within the first 24 hours (Brüning

et al., 1998; Perbellini et al., 1991; Yoshida et al., 1996).

Pulmonary uptake of TCE into the systemic circulation is rapid in animals, after both

administration through the nose only and exposure of the whole body to TCE vapour (IARC,

2014). Blood:gas partition coefficients in rodents vary between species, strains and sexes (Lash,

Parker & Scott, 2000). After inhalation exposure to radiolabelled TCE at 54 or 3200 mg/m3

over a 6-hour period, net pulmonary uptake was 10 times greater at the higher concentration

than at the lower concentration in rats, whereas it was similar at both exposure concentrations

Page 16: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

8

in mice (Stott, Quast & Watanabe, 1982). In humans, TCE is rapidly and extensively absorbed

by the lungs and into the alveolar capillaries. The blood:air partition coefficient of TCE has

been estimated to be approximately 1.5- to 2.5-fold lower in humans than in rodents (Sato et

al., 1977; Monster, 1979; Clewell et al., 1995, Simmons et al., 2002; Mahle et al., 2007). TCE

pulmonary uptake is rapid during the first 30–60 minutes of exposure, and decreases

significantly as TCE concentrations in tissues approach steady state (Fernandez et al., 1977;

Monster, Boersma & Duba, 1979).

Dermal absorption has been demonstrated in mice (Tsuruta, 1978) and guinea-pigs (Jakobson

et al., 1982). Dermal absorption has also been demonstrated in human volunteers (Stewart &

Dodd, 1964; Sato & Nakajima, 1978), with variability in absorption rates between individuals

(Kezic et al., 2000).

3.2 Distribution

Once absorbed, TCE diffuses readily across biological membranes, and is widely distributed

to tissues and organs via the circulatory system. Studies in animals (e.g. Fernandez et al., 1977;

Dallas et al., 1991; Fisher et al., 1991) and humans (De Baere et al., 1997) have found TCE or

its metabolites in most major organs and tissues. Primary sites of distribution include the lungs,

liver, kidneys and central nervous system (CNS). TCE may accumulate in adipose tissue

because of its lipid solubility. in humans, reported tissue:blood partition coefficients were

highest for fat (52–64); the range for all other tissues and organs is much lower, at 0.5–6.0

(IARC, 2014). Slow release of TCE from adipose stores might act as an internal source of

exposure, ultimately resulting in longer mean residence times and bioavailability of TCE

(Fernandez et al., 1977; Dallas et al., 1991; Fisher et al., 1991).

Age-dependent factors may influence TCE distribution in humans (Pastino, Yap & Carroquino,

2000).

3.3 Metabolism

Adverse health effects of TCE are attributed to some of its metabolites (except for solvent

effects that occur at extremely high exposures to the parent compound).

TCE metabolism is quite complex, yielding multiple intermediates and end products (IARC,

2014; Lash et al., 2014; ATSDR, 2019). Experimental animal and human data indicate that

TCE metabolism occurs through two major pathways: cytochrome P450 (CYP)-dependent

oxidation and glutathione (GSH) conjugation catalysed by glutathione S-transferases (GSTs).

Flux through the CYP-dependent oxidation pathway far exceeds that through the GSH

conjugation pathway in all species studied, including humans. Metabolites generated by the

CYP-dependent oxidation pathway are mostly chemically stable. In contrast, the GSH

conjugation pathway generates several highly reactive metabolites. Chemical stability of the

metabolite is an important determinant of systemic availability and fate. Relatively stable TCE

metabolites may be transported from their site of formation into the bloodstream and delivered

to other potential target organs.

TCE metabolism by the oxidative pathway occurs mainly in the liver. Other tissues that are

sites of CYP-mediated TCE metabolism include the lungs, kidneys and male reproductive

organs. Different isozymes of cytochrome P450 oxidize TCE; the highest contribution is by

CYP2E1 (Lash et al., 2014). In the CYP-dependent oxidation pathway, TCE is metabolized to

an epoxide intermediate (TCE epoxide), which spontaneously rearranges to chloral

Page 17: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

9

(trichloroacetaldehyde, trichloroethanal). Chloral is further metabolized to trichloroethanol

(TCOH), trichloroethanol glucuronide (TCOG) and trichloroacetic acid (TCA) as the principal

metabolites. Under certain conditions, TCE epoxide forms dichloroacetyl chloride, which

rearranges to dichloroacetic acid (DCA) (Goeptar et al., 1995). DCA is then further

metabolized by GST. This occurs at a higher rate than metabolism of TCA and TCE; as a result,

measurable DCA concentrations are not often generated in vivo.

TCE metabolism by the GST-catalysed GSH conjugation pathway occurs more slowly than

metabolism by the CYP-catalysed pathway. The initial GSH conjugation step occurs primarily

(but not exclusively) as first-pass metabolism in the liver, which has a high content of GSTs.

The liver is very efficient at excreting GSH conjugates as the S-dichlorovinyl glutathione (1,1-

DCVG and 1,2-DCVG) into either bile or plasma. Subsequently, through enterohepatic and

renal–hepatic circulation, S-dichlorovinyl-L-cysteine (DCVC) or the mercapturate N-acetyl-S-

dichlorovinyl-L-cysteine are delivered to the kidneys for further metabolism or excretion.

DCVG may undergo N-acetylation and be excreted in the urine or metabolized by a lyase

enzyme to reactive metabolites, including a thioacetaldehyde and a thioketene (Clewell et al.,

2001). Additionally, in situ GSH conjugation of TCE can occur within the kidneys themselves,

primarily the proximal tubules, establishing an intra-organ cycle of GSH conjugate transport

and metabolism (Lash et al., 2014).

Exposure to TCE clearly results in exposure of tissues to a complex mixture of metabolites (US

EPA, 2011b).

The very high clearance of TCE seen at low oral doses, which is associated with first-pass

metabolism in the liver, essentially favours the oxidative metabolism pathway. This is the main

reason that the GSH conjugation pathway does not seem to contribute much to the clearance

of TCE at low doses. In addition, the enterohepatic circulation of TCOG is believed to play a

very important role in maintaining TCA levels, and therefore has a major impact on the

oxidative metabolite dosimetry (Stenner et al., 1997, 1998; Barton et al., 1999). The oxidative

metabolites are clearly responsible for the effects on the liver (both cancer and noncancer; see

sections 4 and 5). This implies that the oral route is most important for liver effects, whereas

other routes of exposure may preferentially affect other organs (e.g. kidney).

There are several interspecies differences in TCE metabolism. For example, human hepatic

microsomes have less activity towards TCE than rat or mouse hepatic microsomes (Nakajima

et al., 1993), and humans are less efficient at metabolizing TCE than rodents. Furthermore, a

comparison of renal β-lyase activities in the kidney indicates that rats are more efficient than

humans at metabolizing DCVC to reactive metabolites (Clewell et al., 2000).

TCE metabolism also differs within species. In humans, variations between individuals have

been reported in enzyme expression and activity – for example, in the activity of CYP2E1 and

GST. These reflect differences between the sexes, pathological status, genetic polymorphisms,

or induction and inhibition of the enzymes (Lash, Parker & Scott, 2000; Lash et al., 2014). For

example, chronic exposure to ethanol, a CYP2E1 inducer, is expected to increase TCE

metabolism (Lash et al., 2014). In addition, genetic polymorphisms of OAT1 and OAT3 have

been reported to result in different capacity to accumulate DCVG or DCVC (Lash, Putt &

Parker, 2006), which is likely to affect nephrotoxicity.

Page 18: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

10

3.4 Elimination

The database pertaining to the elimination of TCE is large, and TCE clearance is well

characterized in both animals and humans. In humans, it has been estimated that during and up

to 5 days after a 4-hour inhalation exposure period, pulmonary excretion accounts for 19–35%

of TCE intake, urinary excretion of metabolites accounts for 24–39% of TCE intake, and the

balance is retained in the body (Monster, Boersma & Duba, 1976; Opdam, 1989; Chiu et al.,

2007). Although the elimination kinetics of TCE and its metabolites vary by route of exposure,

elimination pathways appear to be similar for ingestion and inhalation. The half-life of TCE in

alveolar air has been estimated as about 6–44 hours. Half-lives of trichloroethanol and TCA in

urine are 15–50 and 36–73 hours, respectively (IARC, 2014). No data were found regarding

elimination of TCE and its metabolites following dermal exposure.

TCE is eliminated either unchanged in expired air or as metabolites, primarily in urine. The

excreted metabolites are TCA, TCOH or TCOG (following oxidative metabolism), or DCVG

or the cysteine conjugate N-acetyl-S-dichlorovinyl-L-cysteine (following GSH conjugation).

Studies in human volunteers have shown that urinary TCOH is first produced more quickly

and in larger amounts than urinary TCA. However, over time, TCA production eventually

exceeds that of TCOH. Small amounts of metabolized TCE are excreted in the bile or as TCOH

in exhaled air. The total radioactivity recovered in mouse and rat faeces after oral exposure to

radiolabelled TCE accounted for about 1–5% of total radiolabel administered (Dekant, Metzler

& Henschler, 1984; Kim & Ghanayem, 2006), although higher values (up to 24%) were also

reported (Green & Prout, 1985) in another strain of mice. TCE may also be excreted in breast

milk (Pellizzari, Hartwell & Harris, 1982; Fisher et al., 1987; Fisher, Whittaker & Taylor,

1989).

Elimination is more rapid in mice than in rats (Lash, Parker & Scott, 2000), but formation of

TCA is approximately 10 times faster in mice than in rats. These observations help explain

interspecies differences in toxicity associated with TCE, given that the toxicity of TCE is linked

to the formation of its metabolites (Parchman & Magee, 1982; Stott, Quast & Watanabe, 1982;

Dekant, Metzler & Henschler, 1984; Buben & O’Flaherty, 1985; Mitoma et al., 1985; Prout,

Provan & Green, 1985; Rouisse & Chakrabarti, 1986). In humans, differences between

individuals have been seen in the metabolism and elimination of TCE (Nomiyama &

Nomiyama, 1971; Fernandez et al., 1975; Monster, Boersma & Duba, 1976).

3.5 Physiologically based pharmacokinetic modelling

Toxicity studies have been conducted for the inhalation route in humans (occupationally

exposed individuals) and in experimental animals. In contrast, the database on TCE ingestion

via drinking-water is limited. Therefore, many targets of toxicity from chronic exposure to TCE

largely focus on the inhalation route of exposure.

Considering the main features of TCE kinetics, as summarized above, a linear extrapolation

from high-dose studies in rodents to low-dose human exposures seems not be appropriate, for

the following reasons:

• TCE is rapidly and well absorbed by both the oral and inhalation routes of exposure

(ATSDR, 2019).

• The metabolic pathways and kinetics of excretion for oral exposure are similar to those for

inhalation exposure (ATSDR, 2019).

• Data for oral exposure indicate a pattern of effects similar to that of inhalation exposure.

Page 19: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

11

• Differences in first-pass effects (affecting systemic bioavailability) between oral and

inhalation exposures can be adequately accounted for by a physiologically based

pharmacokinetic (PBPK) model.

• Quantitative differences in TCE metabolism between humans and rodents exist.

• Metabolite production is not linear because the oxidative pathway is saturated at the high

doses at which the GST pathway starts to be active.

The use of PBPK modelling allows a route-to-route extrapolation, as well as estimation of the

internal exposure.

Several PBPK models of TCE have been developed, progressively increasing in complexity to

address specific problems in extrapolation of kinetics from rats and mice to humans (Fisher,

2000; Poet et al., 2000; Thrall & Poet, 2000; Simmons et al., 2002; Keys et al., 2003; Hack et

al., 2006; Chiu, Okino & Evans, 2009; Evans et al., 2009; US EPA, 2011b). Recent models

include the kinetics of the relevant TCE oxidative metabolites (chloral hydrate [CH], TCA,

TCOH and trichloroethanol-glucuronide conjugate) (Fisher, 2000; Hack et al., 2006), as well

as the metabolites formed via GSH conjugation in the liver or kidney leading to the appearance

of DCVC (Clewell et al., 2000). The US EPA has derived its own model, based on previous

models but incorporating newer data (Chiu, Okino & Evans, 2009; Evans et al., 2009), and has

applied the updated model to dosimetry extrapolations to support its toxicological review of

TCE (US EPA, 2011b). The model features have been extensively described (ATSDR, 2019).

Starting from the lowest-observed-adverse-effect level (LOAEL) and no-observed-adverse-

effect level (NOAEL) or benchmark dose (BMD) values, PBPK modelling was used to apply

a route-to-route extrapolation and calculate an internal dose based on present understanding of

the role that different TCE metabolites play and the mode of action for TCE toxicity (US EPA,

2011b). The PBPK model was also used to estimate interspecies and intraspecies

pharmacokinetic variability. This resulted in 99th percentile estimates of human equivalent

dose (HED99) for the critical effects.

The PBPK model simulated 100 weeks of human exposure. This was considered representative

of continuous lifetime exposure because longer simulations did not add substantially to the

average (e.g. doubling the simulated exposure time resulted in a change in the resulting HED

of less than a few percent).

4 Effects on humans

4.1 Acute exposure

CNS effects were the primary effects noted from acute inhalation exposure to TCE in humans.

Symptoms included sleepiness, fatigue, headache, confusion and feelings of euphoria (ATSDR,

2019). Simultaneous exposure to TCE and ethanol results in a marked inhibition of the

metabolism of TCE, which leads to accumulation of TCE in the blood and increases the extent

of CNS depression (Muller, Spassovski & Henschler, 1975). Effects on the liver, kidneys,

gastrointestinal system and skin have also been noted (ATSDR, 2019). In its past use as an

inhalant anaesthetic drug in humans, concentrated solutions of TCE have proved quite irritating

to the gastrointestinal tract, and have caused nausea and vomiting (DeFalque, 1961).

Page 20: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

12

4.2 Short-term exposure

Information from medium-term (to long-term) TCE exposures via inhalation and the dermal

route has been reviewed (ATSDR, 2019). These studies indicated that the CNS is the most

sensitive organ for toxicity, followed by the liver and kidneys. Case reports of intermediate and

chronic occupational exposures included effects such as dizziness, headache, sleepiness,

nausea, confusion, blurred vision, facial numbness, and weakness. Liver effects noted included

liver enlargement and increases of liver enzymes in serum. Kidney effects included increased

N-acetyl-β-D-glucosaminidase. Cardiovascular, immunological, reproductive and

carcinogenic effects were also observed (ATSDR, 2019).

4.3 Long-term exposure

4.3.1 Systemic effects

The systemic effects elicited by TCE are not specific to the exposure route; similar effects can

be elicited via oral and inhalation routes.

There is some evidence for TCE-induced hepatic effects (e.g. changes in blood and urine

indices of liver function, enlarged liver) in occupationally exposed humans. However, study

limitations include lack of quantifiable exposure data and confounding due to concomitant

exposure to other chemicals.

Renal toxicity was reported in occupationally exposed humans (although workers were

sometimes also exposed to other chemicals in the workplace). No clear evidence of kidney

effects has been reported in studies examining the association between long-term exposure to

TCE in drinking-water and adverse health effects.

4.3.2 Neurological effects

Reported neurological effects, as described in section 4.2, have been associated with relatively

high exposure to TCE.

4.3.3 Reproductive and developmental effects

Most epidemiological studies have found no convincing association between adverse

reproductive effects in humans and exposure to TCE in contaminated drinking-water (IPCS,

1985; ATSDR, 2019). Epidemiological data are typically limited by concomitant exposure to

other potentially hazardous substances, and case–control studies are limited by small numbers

of cases.

Although an epidemiological study of 2000 male and female workers exposed to TCE via

inhalation found no increase in infant malformations following exposure (IPCS, 1985), an

association was found between the occurrence of congenital heart disease in children and a

drinking-water supply contaminated with TCE and similar chemicals (IPCS, 1985). These

studies were confounded by several factors, including potential exposure to many other

contaminants or compounds that produce similar metabolites, a lack of characterization of the

exposure levels and the exposed populations, and failure to characterize the nature of the

“congenital heart disease” (which may not necessarily be equivalent to cardiac anomalies).

Therefore, use of these studies to indicate a causal association between TCE and congenital

cardiac anomalies remains very limited.

Page 21: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

13

Other epidemiological studies of women exposed to degreasing solvents, including TCE, have

reported elevated risks of cardiac anomalies in their offspring (Goldberg et al., 1990; Ferencz,

Loffredo & Correa-Villaseñor, 1997; Wilson et al., 1998). Large, statistically significant

excesses were observed for specific cardiac defects: left-sided obstructive defects (odds ratio

[OR] = 6.0; 95% confidence interval [CI] = 1.7–21.3) and hypoplastic left heart (OR = 3.4; 95%

CI = 1.6–6.9), with an attributable risk1 of 4.6% (Wilson et al., 1998). Neural tube defects have

also been noted with either occupational or drinking-water exposure to solvents, including TCE

(Holmberg & Nurminen, 1980; Holmberg et al., 1982; Bove, Fulcomer & Klotz, 1995).

In a study in which semen parameters of workers exposed to TCE were evaluated (Chia et al.,

1996), sperm density showed a significant difference between low- and high-exposure subjects.

In a recent study involving a small number of subjects, TCE and its metabolites were identified

in seminal fluids of workers exposed to TCE (Forkert et al., 2003), suggesting that TCE may

play a role in the observed effects on sperm parameters.

Overall, epidemiological studies are plagued by lack of clarity on background coexposure. For

example, in the Wilson et al. (1998) study, the investigators asked subjects about their exposure

to “solvents/de-greasing compounds” but not specifically TCE. Subjects at airforce bases are

exposed to jet fuels as well as other solvents on a daily basis (Stewart, Lee & Marano, 1991),

but it is unlikely that they know the exact compounds contained in the degreasing compounds

or solvents. This means that, based on currently available human studies, TCE cannot be

specifically implicated; however, these studies can be used as supporting evidence,

complementary to developmental and reproductive effects reported in animal studies.

4.3.4 Immunological effects

Studies in humans reported some associations between occupational exposure to TCE and

immunotoxicological end-points. In workers, onset of scleroderma (a systemic autoimmune

disease) has been reported, although a meta-analysis indicated that the available data did not

allow clear conclusions, because of the very low incidence of systemic sclerosis (IARC, 2014).

Some changes in levels of inflammatory cytokines were reported in degreasers using TCE, as

well as case reports of hypersensitivity skin disorder (IARC, 2014).

4.3.5 Genotoxicity and carcinogenicity

Studies examining TCE-induced genotoxicity in humans have been largely inconclusive. Four

studies using peripheral lymphocyte cultures from exposed workers showed no, or only minor,

effects on frequency of sister chromatid exchange (Gu et al., 1981a, b; Nagaya, Ishikawa &

Hata, 1989; Brandom et al., 1990; Seiji et al., 1990). As reviewed by IARC (2014), no further

studies of genotoxicity of TCE in humans have been published.

The carcinogenicity of TCE has been investigated in several types of epidemiological studies,

including cohort and case–control studies in occupationally exposed workers and in the general

population exposed via different routes (inhalation, oral and dermal), in addition to ecological

studies of environmental exposures.

1 Attributable risk is the risk or rate difference that may be attributable to the exposure (Rothman, 1986).

Page 22: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

14

The focus has mainly been on tumours of the kidney and liver, and non-Hodgkin lymphoma.

A clear association between any specific type of cancer and exposure to TCE has not been

consistently observed in these studies. Cancer occurrence in populations exposed to drinking-

water contaminated with various concentrations of TCE has been examined in several studies,

but the interpretation of these studies is complicated by methodological problems.

The evidence for TCE-induced cancers in humans has been reviewed in depth by IARC (2014)

and Rusyn et al. (2014). Three cohort studies were available. Two of these studies, in Sweden

and Finland (Axelson et al., 1994; Anttila et al., 1995), involved people who had been

monitored for exposure to TCE by measurement of TCA in urine.

The third study, in the USA (Spirtas et al., 1991), covered 14,444 workers (10 730 men and

3725 women) exposed to TCE during maintenance of military aircraft and missiles for at least

1 year between 1952 and 1956. Radican et al. (2008) extended the follow-up of this cohort until

2000. These workers were also exposed to other solvents and chemicals, including other

potential carcinogens. Personal and area samples were available for some chemicals, including

TCE (Stewart, Lee & Marano, 1991). Exposure frequency and exposure patterns (intermittent

and continuous) for TCE were assessed based on information on job tasks. TCE was used in

degreasers until 1968, when it was replaced by 1,1,1-trichloroethane (Stewart, Lee & Marano,

1991).

In none of these three cohort studies was it possible to control for potential confounding factors,

such as smoking (IARC, 2014). As of 31 December 2000, 68.1% of cohort members had died.

The Cox model hazard ratio for all cancers was 1.03 (95% CI = 0.91–1.17; 854 deaths). No

significantly increased hazard ratio appeared for any specific cancer in either men or women

(IARC, 2014).

Overall, an elevated risk for liver and biliary tract cancer was observed, in addition to a

modestly elevated risk for non-Hodgkin lymphoma seen in cohort studies. A marginally

increased risk for non-Hodgkin lymphoma was suggested to exist in areas where groundwater

is contaminated with TCE (IARC, 1995, 2014).

The occurrence of renal cancer was not elevated in the cohort studies. However, a study of

German workers exposed to TCE yielded five cases of renal cancer compared with none in a

control comparison group (Henschler et al., 1995). This latter study, conducted on 169 workers

in a cardboard factory in Germany who were exposed to TCE for at least 1 year between 1956

and 1975, claimed a causal link between kidney cancer and TCE exposure (Henschler et al.,

1995). By the close of the study in 1992, 50 members of the study group had died, 16 from

malignant neoplasms. In two of these 16 cases, kidney cancer was the cause of death

(standardized mortality ratio = 3.28, versus local population). Five workers were diagnosed

with kidney cancer: four with renal cell cancer and one with a urothelial cancer of the renal

pelvis (standardized incidence ratio = 7.77; 95% CI = 2.50–18.59). After the close of the

observation period, two additional kidney tumours (one renal and one urothelial) were

diagnosed in the study group. For the seven cases of kidney cancer, the average exposure

duration was 15.2 years (range 3–19.4 years). By the end of the study, 52 members of the

control group, which consisted of 190 unexposed workers from the same plant, had died, 16

from malignant neoplasms, but none from kidney cancer. No case of kidney cancer was

diagnosed in the control group. Although this study received some criticism (McLaughlin &

Blot 1997), it is hard to ignore its findings.

Page 23: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

15

A positive association between renal cancer and prolonged occupational exposure to high

levels of TCE was reaffirmed in a case–control study in Germany involving 134 renal cell

cancer patients and 410 controls, comprising workers from industries with and without TCE

exposure (Brüning et al., 2003). When the results were adjusted for age, sex and smoking, a

significant excess risk was determined for the longest-held job in industries with TCE exposure

(OR = 1.80; 95% CI = 1.01–13.32). Any exposure to degreasing agents was found to be a risk

factor for renal cell cancer (OR = 5.57; 95% CI = 2.33–13.32). Self-reported narcotic symptoms,

an indication of peak exposures, were associated with an excess risk for renal cell cancer (OR

= 3.71; 95% CI = 1.80–7.54). However, the levels of occupational exposure in that study were

very high and unlikely to be reached from environmental exposure. The prolonged exposure to

high levels probably affects the metabolism of TCE, with the net production of active

metabolites underlying the development of renal cell cancer in occupationally exposed

industrial workers.

A more recent case–control study in Montreal, Canada (Christensen et al., 2013), included

histologically confirmed cases of cancer in men (n = 3730; participation rate, 82%; for control,

n = 533) occurring between 1979 and 1985 from 18 of the largest hospitals in the Montreal

metropolitan area. On the basis of job history reported by study subjects, exposure was

estimated for 294 substances; only about 3% of the control individuals were exposed to TCE,

limiting the power of the study. A total of 177 cases of cancer of the kidney were included. For

exposure to TCE, the OR was 0.9 (95% CI = 0.4–2.4) when considering any level of exposure,

and 0.6 (95% CI = 0.1–2.8) for substantial exposure, after adjustment for age, income,

education, ethnicity, questionnaire response and smoking.

The GST gene family encodes multifunctional enzymes that catalyse several reactions between

GST and electrophilic as well as hydrophobic compounds (Raunio et al., 1995). Certain

defective GST genes are known to be associated with an increased risk of different kinds of

cancer. A case–control study (Brüning et al., 1997b) investigated the role of GST

polymorphisms in the incidence of renal cell cancer in two occupational groups exposed to

high levels of TCE. The data indicate a higher risk for development of renal cell cancer if TCE-

exposed people carry either the GSTT1 or GSTM1 gene, compared with individuals lacking the

enzyme. These results, which are supported by the study of Henschler et al. (1995), support the

view of the mode of action of TCE-induced kidney cancer as involving metabolites derived

from the GSH-dependent pathway, at least in humans. Involvement of GST-dependent

metabolites is further supported by a hospital-based case–control study on TCE exposure and

renal cell cancer between 1999 and 2003 in seven central and eastern European cities (Moore

et al., 2010). The final study population included 1097 cases and 1476 controls, who were

interviewed to collect information about exposure and other possible confounders

(e.g. smoking habits). A slight increased risk of renal cell cancer was observed among subjects

ever exposed to TCE (OR = 1.63; 95% CI = 1.04–2.54). No increase in risk of renal cell cancer

was observed among subjects with two deleted GSTT1 alleles: the ORs were 0.93 (95% CI =

0.35–2.44) in ever-exposed subjects, 0.81 (95% CI = 0.24–2.72) in subjects with below-

average exposure, and 1.16 (95% CI = 0.27–5.04) in subjects with above-average exposure

intensity. The presence of at least one copy of the GSTT1 gene did not significantly affect the

OR.

Mutations in the Von Hippel–Lindau (VHL) tumour suppressor gene have been associated with

increased risk of renal cell carcinoma (Brüning et al., 1997a; Brauch et al., 1999). Brüning et

al. (1997a) examined VHL mutation by single-stranded conformation polymorphism in

23 renal cell carcinoma patients with documented high occupational TCE exposure. All TCE-

Page 24: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

16

exposed renal cell carcinoma patients had VHL mutations; this was higher than the background

frequency (33–55%) among unexposed renal cell carcinoma patients. Brauch et al. (1999), in

a follow-up study in 44 TCE-exposed renal cell carcinoma patients, found that 75% of TCE-

exposed patients had VHL mutations and 39% had a C to T mutation at nucleotide 454. All the

C to T transitions in the control renal cell carcinoma patients were relatively rare (6% of the

total incidence).

The US EPA conducted a meta-analysis of epidemiological studies, focusing on non-Hodgkin

lymphoma and cancers of the kidney and liver, as part of its evaluation of the carcinogenicity

of TCE (Scott & Jinot, 2011). Twenty-four studies met the inclusion criteria: two studies with

a high relative risk (RR) for renal cancer (i.e. Henschler et al., 1995, and Vamvakas et al., 1998)

were not included in the meta-analysis because they did not meet the inclusion criteria as a

result of incomplete cohort identification or potential selection bias. Overall meta-RRs for

those exposed to TCE were 1.27 (95% CI = 1.13–1.43) for cancer of the kidney, 1.29 (95% CI

= 0.07–1.56) for cancer of the liver and intrahepatic bile ducts, and 1.23 (95% CI = 1.07–1.42)

for non-Hodgkin lymphoma. An adjustment technique to control for possible publication bias

reduced the meta-RR for non-Hodgkin lymphoma to 1.15 (95% CI = 0.97–1.36). A meta-

analysis largely overlapping with that by Scott & Jinot (2011) was conducted by Karami et al.

(2012). The meta-RR for cancer of the kidney from cohort studies was 1.41 (95% CI = 0.98–

2.05), and 1.26 (95% CI = 1.02–1.56) when the study by Henschler et al. (1995) was excluded.

The meta-RR for case–control studies was 1.55 (95% CI = 1.18–2.05), and 1.35 (95% CI =

1.17–1.57) when the study by Vamvakas et al. (1998) was excluded. The combined RR for

cohort and case–control studies was 1.41 (95% CI = 1.16–1.70). IARC (2014) noted that meta-

RRs were stronger when more recent publications were included; it was suggested that this

might reflect improved exposure assessment and less exposure misclassification. In a meta-

analysis of 18 studies (14 cohort and four case–control) of non-Hodgkin lymphoma, Mandel et

al. (2006) reported meta-RRs of 2.33 (95% CI = 1.39 to 3.91) for non-Hodgkin lymphoma from

studies with higher-quality exposure data, 0.84 (95% CI = 0.73–0.98) from studies with lower-

quality exposure data, and 1.39 (95% CI = 0.62–3.10) from case–control studies.

In conclusion, studies in humans show consistent evidence of an association between

occupational TCE exposure and kidney cancer. Associations reported for liver cancer and non-

Hodgkin lymphoma, although positive, are less consistent.

5 Effects on experimental animals and in vitro test systems

Many studies of a wide range of toxic end-points using repeated oral exposures to TCE have

been reviewed (WHO, 2005). Because of the poor solubility of TCE in water, few studies used

water as a vehicle (Tucker et al., 1982), although some drinking-water or water gavage studies

have used emulsifying agents. Many of the studies are therefore confounded by the use of corn

oil as a vehicle, which has been found to alter the pharmacokinetics of TCE, and to affect lipid

metabolism and other pharmacodynamic processes.

The best documented systemic effects are neurotoxicity, hepatotoxicity, nephrotoxicity and

pulmonary toxicity in adult animals. Reproductive and developmental effects have also been

extensively studied.

5.1 Acute exposure

Neurological, lung, kidney and heart effects have been reported in animals acutely exposed to

TCE (US EPA, 2011b; IARC, 2014; ATSDR, 2019). Tests involving acute exposure of rats

Page 25: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

17

and mice have shown TCE to have low toxicity from inhalation exposure and moderate toxicity

from oral exposure (ATSDR, 2019). The 14-day acute oral median lethal dose (LD50) values

for TCE were determined to be 2400 mg/kg bw in mice (Tucker et al., 1982) and

4920 mg/kg bw in rats (IPCS, 1985; ATSDR, 2019). The 4-hour inhalation median lethal

concentration (LC50) was calculated to be 67 600 mg/m3 in rats (Siegel et al., 1971) and

54 700 mg/m3 in mice (Fan, 1988). A review of studies of dermal exposure of TCE in rabbits

indicated that skin irritation occurs after 24 hours at 0.5 mL, and degenerative skin changes

occur within 15 minutes at 1 mL in guinea-pigs (Fan, 1988). Instillation of 0.1 mL to rabbit

eyes caused conjunctivitis and keratitis, with complete recovery within 2 weeks.

5.2 Short-term exposure

In a 13-week oral study, Fischer 344/N rats and B6C3F1 mice (10 per sex per dose) were

administered TCE in corn oil by gavage at doses of up to 1000 mg/kg bw/day in female rats,

up to 2000 mg/kg bw/day in male rats and up to 6000 mg/kg bw/day in mice of both sexes for

5 days per week (NTP, 1990). Body weights were decreased in male rats at 2000 mg/kg bw/day.

Pulmonary vasculitis involving small veins was reported in female rats at 1000 mg/kg bw/day.

Mild to moderate cytomegaly and karyomegaly of the renal tubular epithelial cells occurred in

rats at 1000 mg/kg bw/day (females) or 2000 mg/kg bw/day (males). The NOAEL in rats was

reported as 1000 mg/kg bw/day (males) and 500 mg/kg bw/day (females). Among the mice,

there were decreases in survival in both sexes and body weight gain in males at

750 mg/kg bw/day and above. In both sexes, doses of 3000 mg/kg bw/day and above were

associated with centrilobular necrosis and multifocal calcification in the liver, as well as mild

to moderate cytomegaly and karyomegaly of the renal tubular epithelial cells. A NOAEL was

set at 375 mg/kg bw/day for mice.

In drinking-water studies (Sanders et al., 1982; Tucker et al., 1982), CD-1 and ICR outbred

albino mice (140 per sex per dose) were administered TCE in a 1% solution of Emulphor in

drinking-water at doses of 0, 0.1, 1.0, 2.5 or 5.0 mg/L (equivalent to 0, 18.4, 216.7, 393 or

660 mg/kg bw/day) for 4 or 6 months. Females at 5.0 mg/L and males at and above 2.5 mg/L

consumed less water than the controls. A decrease in body weight gain in both sexes and an

increase (P < 0.05) in kidney weight in males occurred at 5.0 mg/L. In addition, at 5.0 mg/L,

there were elevated urinary protein and ketone levels in both sexes, decreases in leukocyte and

red blood cell counts in males, altered coagulation times in both sexes and shortened

prothrombin times in females. At 2.5 mg/L, enlargement of the liver, and an increase in urinary

protein and ketone levels in males were observed. Inhibition of humoral immunity, cell-

mediated immunity and bone marrow stem cell colonization was seen among females at

2.5 mg/L and above. The LOAEL was considered to be 2.5 mg/L (equivalent to 393 mg/kg

bw/day) based on decreased water consumption, enlargement of the liver, increases in urinary

protein and ketone levels in males (an indication of renal effects), and changes in

immunological parameters in females. A NOAEL of 1.0 mg/L (equivalent to

216.7 mg/kg bw/day) was determined from these studies. Several previous oral studies in

animals had not found evidence of renal toxicity in mice or rats exposed to TCE (Stott, Quast

& Watanabe, 1982).

Several studies have evaluated the toxicity of TCE to rodents following short-term inhalation

exposure. In a 14-week inhalation study, rats were exposed to TCE at 0, 270, 950 or

1800 mg/m3 for 4 hours per day, 5 days per week, for 14 weeks. Another group was exposed

to TCE cat 300 mg/m3 for 8 hours per day, 5 days per week, for 14 weeks. There were

significant increases (P < 0.01) in the absolute and relative liver weights in treated animals

compared with controls, although liver and kidney function tests of treated animals remained

Page 26: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

18

within normal limits (Kimmerle & Eben, 1973). In a study in which mice, rats and gerbils

(unspecified strains) were exposed to TCE continuously by inhalation at 810 mg/m3 for 30 days,

there was a significant increase (P < 0.05) in the liver weights of all three species (Kjellstrand

et al., 1981). Renal effects of inhaled TCE have also been reported (Kjellstrand et al., 1981,

1983a, b). Male and female gerbils exposed to TCE at 810 mg/m3 continuously for 30 days had

increased (P < 0.05) kidney weight. NMRI mice exposed to TCE at 200, 410, 810 or

1600 mg/m3 continuously for 30 days had significantly increased (P < 0.05) kidney weight, at

410 mg/m3 in males and above 810 mg/m3 in females. No kidney effects were evident in the

remaining strains of mice (Kjellstrand et al., 1983a).

5.3 Long-term exposure

5.3.1 Systemic effects

Administration of high doses of TCE by gavage for long durations in rats and mice has been

associated with nephropathy, with characteristic degenerative changes in the renal tubular

epithelium (NCI, 1976). Toxic nephrosis, characterized by cytomegaly of the renal tubular

epithelium, has been reported in cancer bioassays in mice and rats (NTP, 1983, 1988, 1990).

The toxicity of TCE was investigated in F344 rats and B6C3F1 mice (50 per sex per dose)

given 0, 500 or 1000 mg/kg bw/day (rats) and 0 or 1000 mg/kg bw/day (mice) in corn oil,

5 days per week, for 103 weeks. Survival was reduced in male rats and mice but not in females

(NTP, 1983). Toxic nephrosis, characterized by cytomegaly of the renal tubular epithelium,

occurred in rats at 500 mg/kg bw/day and above, and in mice at 1000 mg/kg bw/day. LOAELs

of 500 mg/kg bw/day in rats and 1000 mg/kg bw/day in mice were defined for long-term

effects. A NOAEL was not determined (NTP, 1990).

5.3.2 Neurological effects

Reported neurological effects have been associated with relatively high exposure to parent TCE;

therefore, they are more frequent after acute and short-term exposure. Intermediate- duration

exposures to TCE have produced neurological effects similar to those found in acute-exposure

situations, including hearing loss (Crofton, Lassiter & Rebert, 1994), increased latency in

visual discrimination (Blain, Lachapelle & Molotchnikoff, 1992), and increased disinhibition

or excitability (ATSDR, 2019).

Most of these effects were found to be reversible when the exposure period ended. When the

neurological effects were evaluated at different doses and time of exposure, concentration was

more relevant than time of exposure in determining effects (Bushnell 1997; ATSDR, 2019).

5.3.3 Reproductive and developmental effects

In a reproductive toxicity study, Long–Evans rats were exposed by inhalation to TCE at

9700 mg/m3 for 6 hours per day, 5 days per week, for 12 weeks before mating; for 6 hours per

day, 7 days per week, only during pregnancy through gestation day (GD) 21; or for 6 hours per

day, 5 days per week, for 2 weeks before mating and for 6 hours per day, 7 days per week,

during pregnancy through GD 21.

Incomplete ossification of the sternum, indicative of delay in maturation, occurred in animals

exposed during pregnancy, and a significant decrease in postnatal weight gain occurred in

offspring of the premating exposed group. No maternal toxicity, teratogenicity or other effects

on reproductive parameters were observed (Dorfmueller et al., 1979).

Page 27: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

19

In a two-generation reproductive toxicity study, male and female Fischer 344 rats were fed

diets containing microencapsulated TCE at doses of approximately 0, 75, 150 or

300 mg/kg bw/day from 7 days before mating through to the birth of the F2 generation.

Although left testicular and epididymal weights decreased in the F0 and F1 generations, no

associated histopathological changes were observed. The weight changes were attributed to

general toxicity, rather than reproductive toxicity (NTP, 1986). In a similar two-generation

reproductive toxicity study in CD-1 mice given TCE at up to 750 mg/kg bw/day, sperm

motility was reduced by 45% in F0 males and 18% in F1 males, but there were no treatment-

related effects on mating, fertility or reproductive performance in the F0 or F1 animals (NTP,

1985).

Numerous teratogenicity studies have been conducted using TCE administered by oral or

inhalation routes. Exposure of Swiss Webster mice to TCE by inhalation at 1600 mg/m3 for

7 hours per day on GD 6–15 did not result in treatment-related maternal toxicity or

teratogenicity (Leong, Schwetz & Gehring, 1975). When Swiss Webster mice and Sprague–

Dawley rats were exposed to TCE by inhalation at a concentration of 1600 mg/m3, 7 hours per

day on GD 6–15, a significant decrease (P < 0.05) in maternal weight gain and some evidence

of haemorrhages in the cerebral ventricles were observed, but no teratogenic or reproductive

effects were seen (Schwetz, Leong & Gehring, 1975). In contrast, a significant decrease in fetal

weight and some increase in fetal resorptions were reported in rats (strain not specified)

exposed to TCE at 540 mg/m3 for 4 hours per day during GD 8–21 (Healy, Poole & Hopper,

1982).

In a study of the effect of exposure to TCE on developmental/reproductive function, female

Sprague–Dawley rats were exposed to TCE in drinking-water at 0, 1.5 or 1100 mg/L (equal to

0, 0.18 or 132 mg/kg bw/day) in one of three dose regimens: for 3 months before pregnancy;

for 2 months before and 21 days during pregnancy; or for 21 days during pregnancy only

(Dawson et al., 1993). No maternal toxicity was observed at any dose level or regimen. An

increase in incidence of fetal heart defects was observed in treated animals at both dose levels

(8.2% at 0.18 mg/kg bw/day and 9.2% at 132 mg/kg bw/day, versus 3% in controls) in dams

exposed before and during pregnancy, and only at the high dose (132 mg/kg bw/day; 10.4%,

versus 3% in controls) in animals exposed only during pregnancy. The LOAEL was set at

0.18 mg/kg bw/day, based on the increased incidence of heart defects in fetuses born to dams

that were exposed before and during gestation. However, the study was limited in that it

expressed the incidence of malformations only as a proportion of the total number of fetuses in

the dose group and did not attempt to establish the incidence of heart defects on a per-litter

basis. Despite this shortcoming, the study lends support to similar findings of increased

congenital defects in epidemiological studies (Goldberg et al., 1990; Bove, Fulcomer & Klotz,

1995), although a clear dose–response relationship is lacking.

A subsequent study (Fisher et al., 2001) conducted with Sprague–Dawley rats treated with TCE,

TCA and DCA at dose levels as high as 400 mg/kg bw/day failed to reproduce the heart

malformations reported by Dawson et al. (1993). However, the two studies differed in design,

which may partly account for the incongruence of the results. First, the Fisher et al. (2001)

study used soybean oil as a vehicle, whereas the Dawson et al. (1993) study used water as a

vehicle. Second, Fisher et al. (2001) administered a very large dose of TCE (400 mg/kg bw/day)

in soybean oil in boluses on GD 5–16 only, whereas Dawson et al. (1993) administered TCE

in drinking-water at lower doses (maximum 1100 mg/L, or 129 mg/kg bw/day) ad libitum,

either during the entire gestation period (GD 1–21) or before and throughout pregnancy; both

the form of test agent and the timing of the dosage may partly account for the variations

Page 28: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

20

between the two studies. Third, the Fisher et al. (2001) study had a very high background

incidence of heart malformations (52% on a per-litter basis) among the control fetuses that

were dosed with soybean oil only – this rate is much higher than the incidence of heart

malformations in the parallel water controls (37%). The Dawson et al. (1993) study reported a

much lower incidence of heart malformations (25% on a per-fetus basis) in the control fetuses

that were dosed with water only. The high background incidence of heart malformations

associated with the controls in the Fisher et al. (2001) study might have masked the effects in

the TCE treatment groups.

In another developmental toxicity study by Johnson et al. (2003), pregnant Sprague–Dawley

rats (9–13 per exposure level) were exposed to TCE throughout pregnancy in drinking-water

at concentrations of 0, 0.0025, 0.25, 1.5 or 1000 ppm. On GD 22, there was a significant

increase in the percentage of offspring with abnormal hearts in the treated groups. The

percentage of litters with abnormal hearts ranged from 0 to 66.7%, while 16.4% of control

litters had abnormal hearts. Although this study appears to suggest the presence of a dose–

response relationship, with the effects beginning to manifest at a dose of 250 μg/L (0.25 ppm;

corresponding to 0.048 mg/kg bw/day) and a NOAEL of 2.5 μg/L (0.00045 mg/kg bw/day),

the dose–response relationship is not as clear on closer examination of the data. However, US

EPA (2011b) calculated a rat BMDL01 (lower 95% confidence limit on the benchmark dose for

a 1% response) of 0.0207 mg/kg bw/day from the fetal heart malformation incidence data. The

BMDL01 was preferred to the NOAEL because the selected nested model accounts for

intralitter effects (i.e. the tendency for littermates to respond more similarly to one another than

to the other litters in a dose group), using pups as the unit of analysis since using litter as the

unit may not be optimal for detecting effects.

5.3.4 Immunological effects

Studies on animals indicate TCE as producing some immunotoxic effects, as reviewed by

IARC (2014). These include accelerated autoimmune responses in mice that are prone to

autoimmune disease, autoimmune hepatitis, inflammatory skin lesions and reduced thymus

weight.

The potential developmental immunotoxicity induced by TCE was studied in B6C3F1 mice

(Peden-Adams et al., 2006), by treating parents (C3H/HeJ male and C57BL/6N female mice;

five per sex per group) with TCE in the drinking-water at 0, 1.4 or 14 ppm, beginning at pairing

(1:1) and continuing for 7 days of mating, and throughout gestation and lactation. Pups were

evaluated for body length, timing of eye opening and ear unfolding. At weaning of the pups at

3 weeks of age, 5–7 pups per treatment group were weighed and sacrificed to assess kidney,

liver, thymus and spleen weights. TCE-related effects on the immune system were assessed by

measuring splenic lymphocyte proliferation, NK cell activity, plaque-forming cell (PFC)

response, splenic B220+ cells, and thymic and splenic T-cell immunophenotypes. The

remaining pups (4–5 pups per treatment group) were assessed at 8 weeks of age in a manner

similar to those assessed at 3 weeks of age, with additional assessments of autoantibodies to

dsDNA (double-stranded DNA) and delayed-type hypersensitivity response. At the lower dose

tested (estimated maternal dose of 0.37 mg/kg bw/day), a decreased PFC response was

observed in 3- and 8-week-old pups, and increased delayed-type sensitivity was noted in 8-

week-old pups. The LOAEL derived from the study was therefore 0.37 mg/kg bw/day.

In the study of Keil et al. (2009), developmental immunotoxicity was evaluated by treating

groups of 9-week-old female B6C3F1 mice (9–10 per group), which are not prone to

spontaneous autoimmune disorders, with TCE in drinking-water at 0, 1.4 or 14 ppm in 1%

Page 29: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

21

Emulphor vehicle for 30 weeks (the low dose was estimated at 0.35 mg/kg bw/day). During

the exposure period, serum levels of total IgG and some autoantibodies (anti-ssDNA [single-

stranded DNA], anti-dsDNA and anti-glomerular antigen) were monitored. At sacrifice, spleen,

thymus, liver and kidney were weighed. Spleen and thymus were processed for assessment of

cell counts and activity, whereas kidneys were processed for histopathologic evaluation. TCE

did not alter NK cell activity, or T- and B-cell proliferation. At the low dose, decreased thymus

weight (30% lower than controls), and increased serum levels of IgG and selected

autoantibodies were observed, but overall there was no evidence that TCE accelerated the onset

of autoimmune disease (ATSDR, 2019).

5.3.5 Genotoxicity and carcinogenicity

A range of assays, covering a wide spectrum of genetic end-points, has been performed to

assess possible genotoxic effects of TCE or its metabolites. DNA- or chromosome-damaging

effects have been evaluated in bacteria, fungi, yeast, plants, insects, rodents and humans, using

many different end-points.

TCE-induced genotoxicity and its possible mechanism have been reviewed (WHO, 2005;

IARC, 2014; ATSDR, 2019).

Evidence for TCE genotoxicity is often conflicting, in part because of the presence of

impurities or mutagenic stabilizers in the test material. In fact, the information from many of

the early studies (until the mid-1990s) may not be adequate for complete evaluation of the

genotoxic potential of TCE, as few of the studies identified the grade and purity of the test TCE.

In addition, some TCE samples used contained a mutagenic stabilizer, and other assays used

pure samples without stabilizers; the TCE in the latter might have decomposed to chemicals

with mutagenic activity, further confounding the interpretation of the significance of the

findings.

TCE is weakly active both in vitro and in vivo, inducing recombination responses, including

sister chromatid exchange, and aneuploidies, including micronuclei; however, it appears to be

unable to induce gene mutations or structural chromosomal aberrations. TCE was also observed

to induce increased DNA synthesis and mitosis in mouse liver in vivo (WHO, 2005). The

potential for TCE to cause DNA strand breaks in rodent liver cells in vivo and in culture at

high concentrations (Bull, 2000) has been questioned (Styles, Wyatt & Coutts, 1991; Chang,

Daniel & DeAngelo, 1992), and was shown not to occur in the kidney (Clay, 2008).

Genotoxicity studies have been conducted for the major metabolites of TCE. CH, DCA and

TCA require very high doses to be genotoxic; insufficient information was available to draw

any conclusions for TCOH, and the conjugates DCVC and DCVG (Moore & Harrington-Brock,

2000). Nevertheless, the US EPA, in revising the genotoxicity of selected TCE metabolites,

concluded that there is relatively strong evidence for genotoxicity of CH and some evidence

for genotoxicity of other TCE metabolites, including DCA, DCVC and DCVG (US EPA,

2011b). Firm conclusions on whether TCE has a mutagenic mode of action cannot be drawn

from the available information.

Overall, results of testing in mammalian and nonmammalian test systems indicate a potential

for TCE to induce chromosomal damage. The weight of evidence suggests that TCE does not

act directly as a mutagenic agent, but that some metabolites have a genotoxic potential.

Page 30: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

22

Carcinogenicity in animal models has been extensively reviewed (US EPA, 2011b; IARC,

2014; ATSDR, 2019). After TCE exposure by the oral route, various types of cancers have

been found in rodents. However, in many studies, rats were administered the maximum

tolerated dose, resulting in a poor survival rate, which affects interpretation of the data. In

addition to other limitations in study design, stabilizers such as epichlorohydrin and other

epoxides were used to prevent TCE degradation when exposed to light. Coexposure to these

stabilizers can be a significant confounding factor; indeed, the observed forestomach tumours

were believed to be induced by the direct alkylating epoxides that were used as stabilizers

(Henschler et al., 1984).

Carcinogenicity studies of TCE by the oral route in mice have demonstrated treatment-related

liver tumours in mice in both sexes (NCI, 1976; NTP, 1983, 1988, 1990). Oral exposure to

TCE has also been shown to increase malignant lymphomas in female mice (US EPA, 2011b).

An increase in the incidence of testicular interstitial cell tumours was reported in male rats;

however, inadequacies in this study mean that the data could not be conclusively interpreted

(NTP, 1988).

In a carcinogenicity assay exposing rodents to TCE by gavage (NTP, 1983), there was a

significant increase in the incidence of hepatocellular carcinomas (P < 0.05) at

1000 mg/kg bw/day in male mice (13/49 relative to 8/48 in controls) and hepatocellular

adenomas (P < 0.05) in female mice (8/49 compared with 2/48 in controls). There were no

treatment-related liver tumours in rats. The male rats at 1000 mg/kg bw/day that survived until

the end of the study showed a higher (P = 0.028) incidence of renal tubular cell

adenocarcinomas (3/16 compared with 0/33 among controls). These kidney tumours were

considered biologically significant, given the rarity of kidney tumours in that rat strain.

A close audit of another carcinogenicity study (NTP, 1988) that exposed four different rat

strains (ACI, August, Marshall and Osborne–Mendel) to TCE by gavage indicated that the

study documentation was inadequate to support proper interpretation of the reported tumour

incidence data. No other treatment-related tumours were reported in these rat strains.

In the NTP (1990) carcinogenicity study, which exposed B6C3F1 mice and F344/N rats to TCE

by gavage, there was a significant (P < 0.05) increase in the incidence of combined

hepatocellular carcinomas and adenomas (P < 0.05) in female mice (22/49 at

1000 mg/kg bw/day, compared with 6/48 in untreated controls). No treatment-related kidney

tumours were observed in mice. Although the study authors considered the results equivocal

because of reduced survival in the treated groups, the kidney tumour incidences in rats were

statistically significant (P < 0.05) when adjusted for reduced survival (2/46 at

500 mg/kg bw/day and 3/33 at 1000 mg/kg bw/day, compared with none in controls); these

tumours were considered toxicologically significant because of the rarity of kidney tumours in

this rat strain.

Carcinogenicity studies of TCE by the inhalation route have shown treatment-related tumours

in the lungs of female and male mice (Fukuda, Takemoto & Tsuruta, 1983; Maltoni et al., 1986),

testes of rats (Maltoni et al., 1986), the lymphoid system (lymphomas) in female mice

(Henschler et al., 1980), the kidney in male rats and the liver in mice of both sexes (Maltoni,

Lefemine & Cotti, 1986).

Overall, animal carcinogenicity studies conducted using pure TCE showed that chronic

exposure to this compound by the oral route resulted in malignant liver tumours in mice of both

sexes and kidney tumours in male rats. Inhalation exposure led to lymphomas in female mice,

Page 31: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

23

malignant liver and lung tumours in mice of both sexes, and malignant kidney tumours in male

rats.

5.4 Mode of action

At TCE concentrations found in most occupational and environmental settings, TCE is

absorbed through the skin by diffusion. At very high concentrations, the absorption can be

enhanced by TCE-induced disruption of the phospholipidic structure of cell membranes

(solvent effect).

After absorption from the gastrointestinal tract, TCE is distributed to the liver first, where toxic

and nontoxic metabolites can be formed (first-pass effect). The toxicity of TCE does not seem

to be heavily dependent on its route of entry.

The similarity between carcinogenic effects induced by the parent compound and tested

metabolites suggests that TCE metabolites are mostly responsible for the liver and kidney

tumours observed in TCE bioassays. This seems to be particularly true for renal cell carcinoma;

additional supporting evidence for the involvement of metabolites derived from the GSH-

dependent pathway, at least in humans, is provided by the lower levels of DNA adduct

formation in individuals with the GST null genotype, who have limited formation of genotoxic

DCVC metabolites (Brüning et al., 1997b).

The mode of action for TCE-induced human renal carcinomas potentially involves mutation of

the VHL tumour suppressor gene, followed by induction of neoplasia (Brüning et al., 1997a).

Indeed, multiple mutations of the VHL tumour suppressor gene, primarily C to T changes,

including nucleotide 454, were found in renal carcinoma patients with prolonged exposure to

high levels of TCE (Brüning et al., 1997b; Brauch et al., 1999). These findings are consistent

with the finding that exposure to TCE at high levels, as in occupational settings (compared

with environmental levels), is highly likely to produce kidney cancer in humans.

The complexity of TCE metabolism and clearance complicates the identification of a

metabolite that is responsible for TCE-induced effects. More than one mode of action may

explain TCE-induced carcinogenicity, and several hypotheses have been put forward. A

number of events are likely to be significant to tumour development in rodents under bioassay

conditions. However, the events that may be more relevant to human exposure to TCE at

environmental levels are not known.

Peroxisome proliferation has been correlated with TCE-induced mouse liver carcinogenesis,

since tumours arise in parallel with peroxisome proliferation associated with TCE metabolites

(Elcombe, 1985; Elcombe, Rose & Pratt, 1985; Goldsworthy & Popp, 1987; Melnick et al.,

1987; DeAngelo et al., 1989; Cattley et al., 1998). TCA can activate the peroxisome proliferator

activated receptor-alpha (PPARα) and the subsequent cascade of responses, including effects

on gene transcription. However, peroxisome proliferation has not been observed in humans, so

chemicals acting with this mechanism in rodents would be unlikely to present a liver

carcinogenic hazard to humans.

Some observations suggest that PPARα activation is not the only mechanism for the

hepatocarcinogenicity of TCE (Rusyn et al., 2014):

• TCA also induces peroxisome proliferation in rats, although to a lower extent than in mice,

but TCA has been shown not to induce liver tumours in F344 rats (DeAngelo et al., 1997).

Page 32: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

24

• The patterns of H-ras mutation frequency associated with DCA and other peroxisome

proliferators are different.

• Other TCA epigenetic effects (see below), including increased c-myc expression and

hypomethylation of DNA, are not specific to the PPARα activation mechanism and can

contribute to TCE-induced liver cancer (Rusyn et al., 2014).

The potential for peroxisome proliferation to play a role in TCE-induced kidney toxicity has

been assessed and is considered unlikely (Lash, Parker & Scott, 2000; Rusyn et al., 2014).

Although TCE has been reported to cause peroxisome proliferation in rat and mouse kidney,

with mice showing a greater response, it has not been shown to induce kidney cancer in mice.

In addition, studies indicate that renal peroxisomes are generally less responsive to peroxisome

proliferators than hepatic peroxisomes (Lash, Parker & Scott, 2000).

Since TCE does not cause α2µ-globulin accumulation (Goldsworthy et al., 1988), this

mechanism is not plausible. In addition, TCE has been identified as causing kidney damage in

both male and female rats (Barton & Clewell, 2000), whereas only male rats are known to

accumulate α2µ-globulin. As such, α2µ-globulin accumulation does not appear to be a mode of

action of TCE-induced kidney toxicity, as was previously thought.

The cysteine and GSH intermediates formed during the metabolism of TCE, DCVC and DCVG

have been shown to have some genotoxic potential, and also to induce the expression of proto-

oncogenes, including c-jun, c-fos and c-myc, in mouse liver tumours (Tao et al., 2000a, b). The

proto-oncogene c-myc is believed to be involved in control of cell proliferation and apoptosis,

which also points towards epigenetic mechanisms for the induction of liver tumours in mice.

Other evidence supports a cytotoxic mode of action. Most rats chronically exposed to TCE in

the National Cancer Institute (NCI) and National Toxicology Program (NTP) bioassays

developed toxic nephrosis, and more than 90% of rats (and mice) developed cytomegaly, which

was most evident in male rats. Associated with these findings, the incidence of kidney tumours

increased only in male rats. The TCE conjugates formed by the action of the β-lyase enzyme

produce proximal tubular necrosis and other lesions in rat kidney (Goeptar et al., 1995) that

lead to the production of reactive species. These reactive species may be responsible for

nephrotoxicity, as well as repair and proliferative responses along a continuum that may

ultimately result in tumorigenesis (Lash, Parker & Scott, 2000; Vaidya et al., 2003).

The formation of DCVG and DCVC in humans indicates that the mode of action associated

with their genotoxicity may be relevant at relatively high dose of exposure in humans. However,

it is uncertain what role it might play in human cancers induced by TCE at exposure levels

below the one activating the GST-dependent pathway or those expected to cause frank kidney

toxicity due to the different TCE metabolites.

Lung tumours were induced in female mice following exposure to TCE (Odum, Foster & Green,

1992). Accumulation of the TCE metabolite CH is thought to be the cause of TCE lung

carcinogenicity, as CH exposure results in lung lesions identical to TCE-induced tumours

(Green, Mainwaring & Foster, 1997; Green, 2000). Accumulation of CH in the Clara cells of

the lung is thought to lead to lung tumours by causing cell damage and compensatory cell

replication, which leads to tumour formation (Green, Mainwaring & Foster, 1997; Green,

2000). However, the mechanism by which CH results in tumour formation in animals may not

be pertinent to humans, as there is little CYP2E1 activity in human lungs (Green, Mainwaring

& Foster, 1997; Green, 2000). A specific lesion, characterized by vacuolization of Clara cells,

was seen only in mice; mice exposed to CH at 600 mg/m3 had similar lesions. Only mild effects

Page 33: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

25

were seen with inhaled TCOH, and none with intraperitoneally administered TCA. These

results suggest that acute lung toxicity of TCE may be due to accumulation of chloral in Clara

cells in mice. Chloral is also genotoxic, and the toxicity observed with intermittent exposures

to TCE is likely to exacerbate any genotoxic effect through compensatory cell proliferation in

rodents.

In conclusion, the mode of action for tumour induction by TCE may be attributed to:

• nongenotoxic processes related to cytotoxicity, peroxisome proliferation, production of

reactive oxygen species and altered cell signalling; or

• genotoxic processes, such as the production of genotoxic metabolites (e.g DCVC) in the

kidney.

The latter possibility cannot be fully ignored, considering evidence of human DNA adducts

formed from genotoxic DCVC metabolites and the presence of VHL tumour suppressor gene

mutations in TCE-exposed kidney cancer patients (Brüning et al., 1997a).

Information on the mode of action for noncancer effects of TCE is more limited, and support

for hypotheses is largely based on observations of actions of other agents. The major endocrine

system effects associated with TCE exposure include the development of testicular (Leydig

cell) tumours in rats (Maltoni et al., 1988; NTP, 1988). TCE and its metabolites TCA and

TCOH have been found to concentrate in the male reproductive organs of rats following

inhalation exposure (Zenick et al., 1984). They have also been found in seminal fluids of

humans occupationally exposed to TCE (Forkert et al., 2003). Peroxisome proliferating

chemicals have been shown to induce Leydig cell tumours via a modulation of growth factor

expression by estradiol (Cook et al., 1999). The occurrence of Leydig cell tumours in rats

exposed to TCE may therefore act as a signal for disturbance of the endocrine system. This

could point to potential endocrine disturbances in humans as a result of TCE exposure. The

effect of endocrine disruption in human populations exposed to TCE is an area requiring further

research.

Studies of the mode of action for observed developmental effects seen with TCE, TCA and

DCA exposure, and data specific to TCE exposure are also scant. Again, a possible role for

peroxisome proliferation with PPARα activation in the development of eye anomalies

following TCE exposure has been hypothesized, although no data currently support it

(Narotsky & Kavlock, 1995; Narotsky et al., 1995).

The TCE metabolites TCA and DCA both produce cardiac anomalies in rats (WHO, 2005).

DCA also concentrates in rat myocardial mitochondria (Kerbey et al., 1976), freely crosses the

placenta (Smith, Randall & Read, 1992) and has known toxicity to tissues dependent on

glycolysis as an energy source (WHO, 2005).

More research into TCE and its metabolites is needed to more fully elucidate possible modes

of action for the effects observed in standard developmental protocols.

6 Overall database and quality of evidence

6.1 Summary of health effects

TCE has anaesthetic properties when inhaled at high concentrations.

Page 34: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

26

Available human and animal data obtained after repeated exposure to TCE identify the kidney,

liver, immune system, male reproductive system and developing fetus as potential targets of

TCE toxicity and/or carcinogenicity.

The systemic effects elicited by TCE are not exposure or route specific; indeed, similar effects

can be elicited via oral and inhalation exposure routes. PBPK models have been developed for

extrapolation from the inhalation route to the oral route, and also for predicting human exposure

levels that would result in effects similar to those observed in rodents.

There is some evidence for TCE-induced hepatic effects (e.g. changes in blood and urine

indices of liver function, enlarged livers) in occupationally exposed humans. Study limitations

include lack of quantifiable exposure data, and concomitant exposure to other chemicals. Dose-

related increases in liver weight, and hepatocellular hypertrophy and peroxisome proliferation

have been consistently reported in TCE-exposed animals.

Renal toxicity was reported in occupationally exposed humans (although workers were

sometimes also exposed to other chemicals in the workplace). No clear evidence of kidney

effects has been reported in studies examining the association between long-term exposure to

TCE in drinking-water and adverse health effects. Epidemiological data are limited by

concurrent exposure to other organic solvents; case–control studies are limited by small

numbers of cases. Studies in animals demonstrate the toxicity of TCE to the male reproductive

system.

Effects such as decreases in litter size and perinatal survival have been reported in rats at

maternally toxic oral doses. Effects were seen in pups exposed at quite high doses

(≥37 mg/kg bw/day). Cardiac arrhythmias were reported in rats exposed to TCE, but not in

humans, unless at lethal doses.

Increased incidences of tumours of the kidney, liver and lymphoid tissues have been reported

in chronic bioassays of rats and mice exposed to very high levels of TCE via inhalation and

oral exposure. Available human data on occupationally exposed subjects provide strong

support for TCE-induced kidney cancer; there are indications that the individuals carrying a

deletion of the GSTT1 gene are much less susceptible to TCE-induced renal tumours. This

suggests that, in humans at high exposure doses, the oxidative pathway is saturated and the

GST-mediated pathway is actively forming reactive metabolites in the kidney.

There is some evidence in humans for an association between exposure to TCE and non-

Hodgkin lymphoma. Evidence for TCE-induced liver cancer in humans is less convincing and,

according to NRC (2009), inadequate.

Associations between the incidence of leukemia and other cancers and oral exposure to TCE

are suggestive, but not definitive, as a result of confounding factors, co-exposures, and

inconsistency between study results.

NRC (2009) concluded that there is limited/suggestive evidence of an association between TCE

exposure and risk of kidney cancer, and inadequate/insufficient evidence for determining

whether associations exist between exposure to TCE and cancer risk at other sites. US EPA

(2011b) concluded that TCE is “carcinogenic to humans by all routes of exposure” based on

convincing evidence of a causal association between TCE exposure and kidney cancer in

humans. IARC (2014) classified TCE as carcinogenic to humans (Group 1), concluding that

the epidemiologic data provide sufficient evidence for an association between exposure to TCE

Page 35: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

27

and human kidney cancer, whereas the associations reported for liver cancer and non-Hodgkin

lymphoma, although positive, are less consistent and thus characterized as limited for these two

cancers. The evidence for other tumours was classified as inadequate (IARC, 2014).

6.2 Quality of evidence

TCE is a data-rich compound: many good-quality studies are available on its kinetics and toxic

effects in both animals and humans. Studies of workers and volunteers have provided most of

the data on health effects of inhaled TCE in humans. Although available data on oral exposure

have been considered of questionable validity as a result of co-exposure to other contaminants,

the availability of PBPK modelling allows inhalation-to-oral route extrapolation.

Aspects that are not fully elucidated include:

• uncertainty about the mode of action for tumour induction, for which many mechanisms

have been proposed, but none of them definitely established; the potential genotoxicity of

some GST-mediated TCE metabolites at high exposure doses make the picture even more

complex; and

• regardless of the mode of action, the relevance for humans, which could not be completely

assessed for some tumours.

Kidney cancer has been reported in humans in studies considered to be of adequate quality, but

only at high levels of exposure, such as in occupational settings. Liver effects have been

consistently reported, although TCE has been shown to have multiple targets in animals and

humans. None of the studies can be considered optimal, but they all concur to build a weight-

of-evidence approach. For this reason, an overall reference value was derived by the US EPA

(2011b), with tolerable daily intake (TDI) values derived from some recent studies falling

within a narrow range of 0.0003–0.0006 mg/kg bw/day (see section 8.1), thus enhancing the

strength of the findings.

7 Practical considerations

7.1 Analytical methods and achievability

TCE can be analysed together with trichloroethene by gas chromatography using

ISO 10301:1997, which has a limit of quantification of 0.1 μg/L (ISO, 1997).

Four methods for measuring TCE in drinking-water have been approved by the US EPA:

• Method 502.2, which employs purge and trap capillary gas chromatography with

photoionization detectors and electrolytic conductivity detectors in series, has a detection

limit in the range of 0.01–3.0 μg/L (US EPA, 1999).

• Method 524.2, which uses purge and trap capillary gas chromatography with mass

spectrometric detectors in series, has a detection limit of 0.5 μg/L (US EPA, 1999).

• Method 503.1, which uses purge and trap capillary gas chromatography with

photoionization conductivity detectors, has a detection limit of 0.01–3.0 μg/L (US EPA,

1999).

• Method 551.1, which uses liquid–liquid extraction and gas chromatography with electron

capture detectors, has a method detection limit of 0.01 μg/L (US EPA, 1999).

Page 36: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

28

7.2 Source control

TCE is primarily, if not exclusively, a groundwater contaminant because it is lost to the

atmosphere from surface waters. The primary cause of contamination is poor handling and

disposal practices, which result in soil contamination and, in vulnerable aquifers, subsequent

water contamination. This may occur some distance from the source, but contaminants may be

drawn into the source by pumping. Control at source should be relatively cheap and

straightforward by improving handling and disposal practices.

7.3 Treatment methods and performance

Treatment of surface water sources is not needed because TCE volatilizes to the atmosphere.

The most effective technique for removal of TCE from groundwater is aeration, including

packed tower aeration. Granular activated carbon (GAC) adsorption has also been shown to be

effective. TCE concentrations below 2 μg/L should be achievable by air stripping (Duan, Ito &

Ohkawa, 2001). Combining air stripping and activated carbon has been shown to improve TCE

removal.

Aeration has been used to treat contaminated well water (27 μg/L) at pilot scale. For an air-to-

water ratio of 10, a rate of 25 m/hour and a 3.75 m contact height, the process achieved a 67%

reduction in TCE (Simon & Mitchell, 1992).

Pilot-scale tests using air stripping achieved TCE removals from water with an influent

concentration of 204 μg/L of 82% and 87% for air-to-water ratios of 75:1 and 125:1,

respectively (McKinnon & Dyksen, 1984). Other pilot-scale studies using diffused aeration

have achieved removals of 70–92% using an air-to-water ratio of 4:1 and a 10 minute contact

time (Kruithof et al., 1985). One study investigated the effect of media depth on the removal

rate. A packed tower with a media depth of 4.5 m, an air-to-water ratio of 30:1 and a liquid

loading rate of 13.8 l L/m2∙s achieved a removal of 98.2%, whereas a packed tower with a

media depth of 1.2 m achieved a removal of 45% under the same conditions (Amy, Narbaitz

& Cooper, 1987).

Gross & TerMaath (1985) studied the performance of packed tower aeration for stripping TCE

from groundwater at full scale. The groundwater contained TCE concentrations ranging from

50 to 8000 µg/L. During the study, two towers were operated in both series and parallel

configuration at air-to-water ratios of 10, 18 and 25. Higher removals were observed in the

series configuration (96–99.9%) than in the parallel configuration (86–99%). The highest

removal was observed at an air-to-water ratio of 25. Hand et al. (1988) studied the effectiveness

of air stripping for removing TCE from contaminated well water and found that, at an air-to-

water ratio of 60:1, approximately 98% of the influent TCE was removed (influent

concentration of 72 µg/L). Cummins (1985) found that, with a packing height of 5.5 m and an

air-to-water ratio of 39:1, approximately 98.5% of the influent TCE was removed (influent

concentrations of 20–38 mg/L) from three wells used for source.

Full-scale spray aeration of well water containing TCE at up to 10 μg/L achieved 90% removal

to below 1 μg/L (Kruithof & Koppers, 1989).

GAC has been used to remove high concentrations of TCE at pilot scale. The carbon effectively

removed 100% of the influent concentration (approximately 2500 μg/L), for 30 bed volumes

at an empty bed contact time (EBCT) of 2.5 minutes and 40 bed volumes at an EBCT of 10 min

Page 37: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

29

(Hand et al., 1994). The presence of humic substances can decrease GAC adsorption (Urano,

1991).

A study of 68 full-scale treatment facilities in the USA treating water for volatile organic

compounds (VOCs) found that facilities employing GAC achieved TCE removal efficiencies

of >99%. Influent concentrations ranged from 3 to 400 µg/L, and the treatment facilities

operated under a variety of different configurations and design parameters; loading rates varied

from 4.6 to 18.5 m/hour, and EBCT varied from 9 to 30 minutes (AWWA, 1991).

In a municipal-scale treatment plant combining air stripping and GAC, TCE was removed to

levels below 1 µg/L (US EPA, 1985b).

Ozone doses of 2, 6 and 20 mg/L achieved TCE removals of 39%, 76% and 95%, respectively

(Fronk, 1987). Pilot plant studies have shown that ozonation can almost completely remove

trace concentrations of TCE from groundwater (Slagle, 1990).

Karimi et al. (1997) investigated the use of a two contactor ozone/hydrogen peroxide advanced

oxidation process for TCE oxidation at full scale. The influent TCE concentration ranged from

32 to 477 µg/L. In all tests, the treated water concentration of TCE was below 5 µg/L in the

second contactor. Greater than 90% removal of TCE was achieved in all tests after the first

contactor, with effluent TCE concentrations <5 µg/L when the influent concentration was

<100 µg/L. The greatest reduction in TCE (≥99%) occurred for ozone and H2O2 doses of 4.0–

4.6 mg/L and 2.2–2.4 mg/L, respectively. The authors noted that the optimum ratio of

H2O2/ozone was 0.5–0.6.

A combination of H2O2 and ultraviolet (UV) irradiation has been used to treat groundwater

contaminated with VOCs, including TCE (0.89–1.30 mg/L). At 38 L/minute, with a reactor

volume of 57 L and with H2O2 dosed at 65 mg/L, the effluent concentration of TCE was

generally below detection limits (maximum removal efficiency was >99.9%) (Topudurti et al.,

1994). Other research has confirmed that TCE is readily removed from water by ozone and that

UV irradiation gave only a slight improvement; 75 mg/L was removed by an ozonation rate of

6 mg/L and UV fluence of 100 mW∙s/cm2 (Pailard, Brunet & Dore, 1987).

Microbial remediation of TCE has also been described (Pant & Pant, 2010).

8 Conclusion

8.1 Derivation of the guideline value

The previous World Health Organization (WHO) evaluation (WHO, 2005) considered both

cancer and noncancer end-points in deriving the guideline value (GV) for TCE in drinking-

water. The GV for TCE was ultimately based on the noncancer end-points, and was protective

for both cancer and noncancer end-points.

In the previous WHO evaluation (WHO, 2005), the developmental toxicity study from Dawson

et al. (1993) was chosen as the point of departure (POD), based on the appropriateness of the

route (drinking-water), the low dose at which the effects were observed (which coincides with

the LOAEL in all animal studies reviewed), the severity of the end-point, the evidence of

similar effects (e.g. cardiac anomalies) from epidemiological studies, and the observation of

similar malformations in studies of TCE metabolites. However, it has been recognized that the

Dawson et al. (1993) study has several significant methodological limitations, including the

spontaneous incidence of the critical end-point (heart malformations).

Page 38: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

30

In the present evaluation, it was considered more appropriate to take into account potential

PODs for candidate chronic TDI values from various studies by using the LOAEL/NOAEL

approach, benchmark dose (BMD) analysis, and PBPK modelling of human and animal data

considered suitable for assessment of a dose–response relationship (US EPA, 2011b).

Therefore, all the possible candidate PODs, rather than a single key study, were included in the

derivation of the TDI, considering various end-points. End-points included TCE-induced

neurological effects in humans and animals; effects on kidney, liver and body weight in animals;

immunological effects in animals; reproductive effects in humans and animals; and

developmental effects in animals.

PBPK modelling was used to calculate internal doses from a number of studies for which

plausible internal dose metrics were available, based on present understanding of the role that

different metabolites play in TCE toxicity. The exception was the Peden-Adams et al. (2006)

study, because model parameters were not available; therefore the LOAEL was used.

The PBPK model estimated interspecies and intraspecies pharmacokinetic variability, and

resulted in 99th percentile estimates of human equivalent dose (HED99) for candidate critical

effects. The PBPK model simulated 100 weeks of human exposure. This was considered

representative of continuous lifetime exposure because longer simulations did not add

substantially to the average (e.g. doubling the exposure time resulted in a change of less than a

few percentage points in the resulting HED).

Among the available studies, three were considered critical to deriving the TDI:

• Keil et al. (2009), from which a LOAEL of 0.35 mg/kg bw/day was identified, based on

decreased thymus weight in female mice exposed to TCE in the drinking-water for

30 weeks. A PBPK model was used to derive an HED99 of 0.048 mg/kg bw/day for lifetime

continuous exposure, which was used as the POD.

• Peden-Adams et al. (2006), from which a LOAEL of 0.37 mg/kg bw/day was identified

and considered as the POD, based on developmental immunotoxicity effects: decreased

plaque-forming cell response (at 3 and 8 weeks of age) and increased delayed-type

hypersensitivity (at 8 weeks of age) in pups exposed from GD 0 until 3 or 8 weeks of age

through drinking-water (placental and lactational transfer, and pup ingestion). A BMD

could not be calculated because of inadequate model fit, and no PBPK modelling was

applied because of lack of appropriate models and parameters to account for fetal and pup

exposure patterns.

• Johnson et al. (2003), in which pregnant Sprague–Dawley rats were administered TCE in

drinking-water during GD 1–22 at concentrations ≥0.0025 ppm. Increased incidences of

fetal cardiac malformations at maternal exposure levels ≥0.25 ppm (estimated maternal

doses ≥0.048 mg/kg bw/day) were identified as the critical effect. A PBPK model was

applied to the rat BMDL01 external dose of 0.0207 mg/kg bw/day to calculate the rat

internal dose; this was converted to an HED99 of 0.0051 mg/kg bw/day.

No new carcinogenicity data were identified since the previous evaluation (WHO, 2005). The

key studies have been selected among the ones identifying noncancer effects, with a POD that

is lower than that used in the previous evaluation.

Page 39: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

31

8.1.1 Noncancer effects

Limitations of individual studies were overcome through selection of multiple critical effects,

rather than selecting the lowest NOAEL or BMDL as a POD.

By using this approach, as described by US EPA (2011b), a PBPK model was used to calculate

an internal dose POD from a number of studies for which plausible internal dose metrics could

be determined, based on present understanding of the role that different metabolites play in

TCE toxicity and the mode of action for toxicity.

The PBPK model was used to estimate interspecies and intraspecies pharmacokinetic

variability, and resulted in an HED99 for candidate critical effects.

From the three critical studies, the TDI derivation was as follows:

• Keil et al. (2009)

HED99 = 0.048 mg/kg bw/day

Uncertainty factor (UF) = 10 to account for use of a LOAEL

UF = 2.5 to account for remaining uncertainty associated with interspecies

toxicodynamic differences, because a PBPK model was used to characterize interspecies

toxicokinetic differences

UF = 3.2 to account for remaining uncertainty associated with human variability in

toxicodynamics, because a PBPK model was used to characterize human toxicokinetic

variability

TDI = 0.048/80 = 0.0006 mg/kg bw/day

• Peden-Adams et al. (2006)

LOAEL = 0.37 mg/kg bw/day

UF = 10 to account for use of a LOAEL

UF = 10 for interspecies extrapolation (a default factor was used, because of lack of

adequate toxicokinetic data to develop a PBPK model)

UF = 10 for human variability (a default factor was used, because of lack of adequate

toxicokinetic data to develop a PBPK model)

TDI = 0.37/1000 = 0.00037 mg/kg bw/day

• Johnson et al. (2003)

HED99 = 0.0051 mg/kg bw/day (derived from a BMDL01)

UF = 2.5 to account for remaining uncertainty associated with interspecies

toxicodynamic differences, because a PBPK model was used to characterize interspecies

toxicokinetic differences

Page 40: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

32

UF = 3.2 to account for remaining uncertainty associated with human variability in

toxicodynamics, because a PBPK model was used to characterize human toxicokinetic

variability

TDI = 0.0051/8 = 0.00064 mg/kg bw/day

The TDI values fall within a narrow range of 0.0003–0.0006 mg/kg bw/day. The PBPK model-

based TDI value is 0.0006 mg/kg bw/day for both heart malformations in rats (Johnson et al.,

2003) and decreased thymus weights in mice (Keil et al., 2009). The lowest TDI comes from a

third study (Peden-Adams et al., 2006), which allowed derivation of a TDI value of

0.00037 mg/kg bw/day, based on the applied dose LOAEL for developmental immunotoxicity.

Further supporting data in the database are toxic nephropathy in rats (NTP, 1988;

0.0003 mg/kg bw/day) and increased kidney weight in rats (Boverhof et al., 2013;

0.0008 mg/kg bw/day by using route-to-route extrapolation from the inhalation study).

An overall TDI of 0.0005 mg/kg bw/day (0.5 µg/kg bw/day) was considered appropriate, being

supported by multiple effects rather than an individual value. This approach is less sensitive to

limitations of individual studies.

8.1.2 Guideline value

Based on the TDI as described above, the GV is:

GV= 0.5 µg/kg bw/day × 60 kg bw × 0.5 = 7.5 µg/L (rounded to 8 µg/L or 0.008 mg/L)

2 L/day

where

• 0.5 µg/kg bw/day is the TDI, as derived above

• 60 kg is the average body weight of an adult

• 0.5 is the fraction of the total daily intake that is allocated to drinking-water

• 2 L is the daily volume of water consumed by an adult.

An allocation factor higher than the default 20% factor is used, since the occurrence of TCE in

food is low (see section 2.2) and human exposures to TCE overall have been declining (see

section 2.5), as a result of increased environmental regulations governing TCE emissions

(IARC, 2014; ATSDR, 2019).

8.2 Considerations in applying the guideline value

In certain circumstances, there may be a need to adapt the GV by adjusting the allocation factor

or considering the Leq/day corresponding to inhalation exposure from the domestic use of

water to account for local conditions, including inhalational exposure from high rates of

showering and bathing, and/or from living in poorly ventilated buildings. This indirect

exposure may be particularly relevant in such settings because of TCE evaporation rates; the

contribution of indirect exposure via indoor air might be similar to, or higher than, that from

oral water ingestion.

Requirements for monitoring TCE in drinking-water regulations and standards should be

limited to groundwater sources where a catchment risk assessment indicates the possibility of

presence of TCE, such as where TCE is, or was, used as a degreasing agent in large or small

Page 41: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

33

industries with limited resources for safe handling and disposal. In some cases, this may include

TCE drawn in from contamination elsewhere in the aquifer. Monitoring can be conducted at

the treatment works. If concentrations are shown to be stable or effective treatment is in place,

the frequency of monitoring can be quite low.

If monitoring data show elevated levels of TCE, it is suggested that a plan be developed and

implemented to address these situations. Monitoring is not needed for surface water sources

because TCE volatilizes to the atmosphere.

Page 42: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

34

References

Akiyama M, Matsui Y, Kido J, Matsushita T, Shirasaki N (2018). Monte-Carlo and multi-exposure

assessment for the derivation of criteria for disinfection byproducts and volatile organic

compounds in drinking water: allocation factors and liter-equivalents per day. Regul Toxicol

Pharmacol. 95:161–74.

Amoore JE, Hautala E (1983). Odor as an aid to chemical safety: odor thresholds compared with

threshold limit values and volatilities for 214 industrial chemicals in air and water dilution. J

Appl Toxicol. 3:272–90.

Amy GL, Narbaitz RM, Cooper WJ (1987). Removing VOCs from groundwater containing humic

substances by means of coupled air stripping and adsorption. J Am Water Works Assoc.

79(8):49–54.

Anttila A, Pukkala E, Sallmén M, Hernberg S, Hemminki K (1995). Cancer incidence among Finnish

workers exposed to halogenated hydrocarbons. J Occup Environ Med. 37:797–806.

ATSDR (Agency for Toxic Substances and Disease Registry) (2019). Toxicological profile for

trichloroethylene. Atlanta, Georgia: United States Department of Health and Human Services.

AWWA (American Water Works Association) (1991). Existing VOC treatment installations: design,

operation and cost. Report of the Organics Contaminant Control Committee. Denver, Colorado:

Water Quality Division, AWWA (Report No. 0033986).

Axelson O, Seldén A, Andersson K, Hogstedt C (1994). Updated and expanded Swedish cohort study

on trichloroethylene and cancer risk. J Occup Med. 36:556–62.

Barton HA, Clewell HJ III (2000). Evaluating noncancer effects of trichloroethylene: dosimetry, mode

of action, and risk assessment. Environ Health Perspect. 108(Suppl. 2):323–34.

Barton HA, Bull R, Schultz I, Andersen ME (1999). Dichloroacetate (DCA) dosimetry: interpreting

DCA-induced liver cancer dose response and the potential for DCA to contribute to

trichloroethylene-induced liver cancer. Toxicol Lett. 106:9–21.

Blain L, Lachapelle P, Molotchnikoff S. 1992. Evoked potentials are modified by long term exposure

to trichloroethylene. Neurotoxicology 13:203–6.

Bogen KT, Hall LC, Perry L, Fish R, McKone TE, Dowd P, et al. (1988). Health risk assessment of

trichloroethylene (TCE) in California drinking water. Livermore, California: Environmental

Sciences Division, Lawrence Livermore National Laboratory, University of California

(DE88005364).

Boutonnet JC, De Rooij C, Garny V, Lecloux A, Papp R, Thompson RS, et al. (1998). Euro Chlor risk

assessment for the marine environment OSPARCOM region: North Sea – trichloroethylene.

Environ Monit Assess. 53:467–87.

Bove FL, Fulcomer MC, Klotz JB (1995). Public drinking water contamination and birth outcomes.

Am J Epidemiol. 141:850–62.

Boverhof DR, Krieger SM, Hotchkiss JA, Stebbins KE, Thomas J, Woolhiser MR (2013). Assessment

of the immunotoxic potential of trichloroethylene and perchloroethylene in rats following

inhalation exposure. J Immunotoxicol. 10(3):311–20.

Brandom WF, McGavran L, Bristline RW, Bloom AD (1990). Sister chromatid exchanges and

chromosome aberration frequencies in plutonium workers. Int J Radiat Biol. 58:195–207.

Brauch H, Weirich G, Hornauer MA, Störkel S, Wöhl T, Brüning T (1999). Trichloroethylene exposure

and specific somatic mutations in patients with renal cell carcinoma. J Natl Cancer Inst. 91:854–

61.

Brenner D (2010). Results of a long-term study of vapor intrusion at four large buildings at the NASA

Ames Research Center. J Air Waste Manag Assoc. 60(6):747–58.

Page 43: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

35

Brugnone F, Perbellini L, Giuliari M, Cerpelloni M, Soave M (1994). Blood and urine concentrations

of chemical pollutants in the general population. Med Lav. 85:370–89.

Brüning T, Weirich G, Hornauer MA, Höfler H, Brauch H (1997a). Renal cell carcinomas in

trichloroethylene (TRI) exposed persons are associated with somatic mutations in the Von

Hippel–Lindau (VHL) tumour suppressor gene. Arch Toxicol. 71:332–5.

Brüning T, Lammert M, Kempkes M, Thier R, Golka K, Bolt HM (1997b). Influence of polymorphisms

of GSTM1 and GSTT1 for risk of renal cell cancer in workers with long-term high occupational

exposure to trichloroethylene. Arch Toxicol. 71:596–9.

Brüning T, Vamvakas S, Makropoulos V, Birner G (1998). Acute intoxication with trichloroethylene:

clinical symptoms, toxicokinetics, metabolism, and development of biochemical parameters for

renal damage. Toxicol Sci. 41:157–65.

Brüning T, Pesch B, Wiesenhütter B, Rabstein S, Lammert M, Baumüller A, et al. (2003). Renal cell

cancer risk and occupational exposure to trichloroethylene: results of a consecutive case–

control study in Arnsberg, Germany. Am J Ind Med. 43(3):274–85.

Buben JA, O’Flaherty EJ (1985). Delineation of the role of metabolism in the hepatotoxicity of

trichloroethylene and perchloroethylene: a dose-effect study. Toxicol Appl Pharmacol. 78:105–

22.

Bull RJ (2000). Mode of action for liver tumour induction by trichloroethylene and its metabolites,

trichloroacetate and dichloroacetate. Environ Health Perspect. 108(Suppl. 2):241–59.

Burk T, Zarus G (2013). Community exposures to chemicals through vapor intrusion: a review of past

Agency for Toxic Substances and Disease Registry public health evaluations. J Environ Health.

75(9):36–41.

Bushnell PJ (1997). Concentration–time relationships for the effects of inhaled trichloroethylene on

signal detection behavior in rats. Fund Appl Toxicol. 36:30–8.

Canadian Department of National Health and Welfare (1993). Trichloroethylene. Supporting

documentation, health related sections for the Canadian Environmental Protection Act (CEPA)

Priority Substances List assessment report. Ottawa, Ontario: Health Canada.

Cattley RC, DeLuca J, Elcombe C, Fenner-Crisp P, Lake BG, Marsman DS, et al. (1998). Do

peroxisome proliferating compounds pose a hepatocarcinogenic hazard to humans? Regul

Toxicol Pharmacol. 27:47–60.

Chang LW, Daniel FB, DeAngelo AB (1992). Analysis of DNA strand breaks induced in rodent liver

in vivo, hepatocytes in primary culture, and a human cell line by chlorinated acetic acids and

chlorinated acetaldehydes. Environ Mol Mutagen. 20:277–88.

Chia SE, Ong CN, Tsakok MF, Ho A (1996). Semen parameters in workers exposed to

trichloroethylene. Reprod Toxicol. 10:295–9.

Chiu WA, Okino MS, Evans MV (2009). Characterizing uncertainty and population variability in the

toxicokinetics of trichloroethylene and metabolites in mice, rats, and humans using an updated

database, physiologically based pharmacokinetic (PBPK) model, and Bayesian approach.

Toxicol Appl Pharmacol. 241(1):36–60.

Chiu WA, Micallef S, Monster AC, Bois FY (2007). Toxicokinetics of inhaled trichloroethylene and

tetrachloroethylene in humans at 1 ppm: empirical results and comparisons with previous

studies. Toxicol Sci. 95:23–36.

Chiu WA, Jinot J, Siegel Scott C, Makris SL, Cooper GS, Dzubow RC, et al. (2013). Human health

effects of trichloroethylene: key findings and scientific issues. Environ Health Perspect.

121:303–11.

Page 44: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

36

Christensen KY, Vizcaya D, Richardson H, Lavoué J, Aranson K, Siemiatycki J (2013). Risk of selected

cancers due to occupational exposure to chlorinated solvents in a case–control study in

Montreal. J Occup Environ Med. 55:198–208.

Clay P (2008). Assessment of the genotoxicity of trichloroethylene and its metabolite, S-(1,2-

dichlorovinyl)-L-cysteine (DCVC), in the comet assay in rat kidney. Mutagenesis. 23:27–33.

Clewell HJ, Gentry PR, Gearhart JM, Allen BC, Andersen ME (1995). Considering pharmacokinetic

and mechanistic information in cancer risk assessments for environmental contaminants:

examples with vinyl chloride and trichloroethylene. Chemosphere. 31:2561–78.

Clewell HJ, Gentry PR, Covington TR, Gearhart JM (2000). Development of a physiologically based

pharmacokinetic model of trichloroethylene and its metabolites for use in risk assessment.

Environ Health Perspect. 108(Suppl. 2):283–305.

Clewell H, Barton HA, Maull EA, Andersen ME (2001). Under what conditions is trichloroethylene

likely to be a carcinogen in humans? Hum Ecol Risk Assess. 7(4):687–716.

Cook JC, Klinefelter GR, Hardisty JF, Sharpe RM, Foster PM (1999). Rodent Leydig cell

tumorigenesis: a review of the physiology, pathology, mechanisms and relevance to humans.

Crit Rev Toxicol. 29:169–261.

Crofton KM, Lassiter TL, Rebert CS (1994). Solvent-induced ototoxicity in rats: an atypical selective

mid-frequency hearing deficit. Hear Res. 80:25–30.

Cummins MD (1985). Field evaluation of packed column air stripping at Valley Park, Missouri.

Cincinnati, Ohio: United States Environmental Protection Agency.

Dallas CE, Gallo RM, Ramanathan R, Muralidhara S, Bruckner JV (1991). Physiological

pharmacokinetic modeling of inhaled trichloroethylene in rats. Toxicol Appl Pharmacol.

110:303–14.

Daniel JW (1963). The metabolism of 36Cl-labelled trichloroethylene and tetrachloroethylene in the rat.

Biochem Pharmacol. 12:795–802.

Dawson BV, Johnson PD, Goldberg SJ, Ulreich JB (1993). Cardiac teratogenesis of halogenated

hydrocarbon-contaminated drinking water. J Am Coll Cardiol. 21:1466–72.

DeAngelo AB, Daniel FB, McMillan L, Wernsing P, Savage RE (1989). Species and strain sensitivity

to the induction of peroxisome proliferation by chloroacetic acids. Toxicol Appl Pharmacol.

101:285–98.

DeAngelo AB, Daniel FB, Most BM, Olson GR (1997). Failure of monochloroacetic acid and

trichloroacetic acid administered in the drinking water to produce liver cancer in male F344/N

rats. J Toxicol Environ Health. 52:425–45.

De Baere S, Meyer E, Dirinck I, Lambert W, Piette M, Van Peteghem C, et al. (1997). Tissue

distribution of trichloroethylene and its metabolites in a forensic case. J Anal Toxicol. 21:223–

7.

DeFalque RJ (1961). Pharmacology and toxicology of trichloroethylene: a critical review of world

literature. Clin Pharmacol Ther. 2:665–88.

Dekant W, Henschler D (1983). New pathways of trichloroethylene metabolism. Dev Toxicol Environ

Sci. 11:399–402.

Dekant W, Metzler M, Henschler D (1984). Novel metabolites of trichloroethylene through

dechlorination reactions in rats, mice, and humans. Biochem Pharmacol. 33:2021–37.

Dorfmueller MA, Henne SP, York RG, Bornschein RL, Manson JM (1979). Evaluation of

teratogenicity and behavioral toxicity with inhalation exposure of maternal rats to

trichloroethylene. Toxicology. 14:153–66.

Page 45: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

37

ECSA (European Chlorinated Solvents Association) (2012). Product safety summary on

trichloroethylene. Brussels: ECSA.

Elcombe CR (1985). Species differences in carcinogenicity and peroxisome proliferation due to

trichloroethylene: a biochemical human hazard assessment. Arch Toxicol Suppl. 8:6–17.

Elcombe CP, Rose MS, Pratt IS (1985). Biochemical, histological, and ultrastructural changes in rat

and mouse liver following administration of trichloroethylene: possible relevance to species

differences in hepatocarcinogenicity. Toxicol Appl Pharmacol. 79:365–76.

Evans MV, Chiu WA, Okino MS, Caldwell JC (2009). Development of an updated PBPK model for

trichloroethylene and metabolites in mice, and its application to discern the role of oxidative

metabolism in TCE-induced hepatomegaly. Toxicol Appl Pharmacol. 236(3):329–40.

Fan AM (1988). Trichloroethylene: water contamination and health risk assessment. Rev Environ

Contam Toxicol. 101:55–92.

Ferencz C, Loffredo CA, Correa-Villaseñor A (1997). Genetic and environmental risk factors of major

cardiovascular malformation. the Baltimore–Washington Infant Study 1981–1989.

Perspectives in Pediatric Cardiology, Vol. 5. New York: Futura Publishing.

Fernandez JG, Humbert BE, Droz PO, Caperos JR (1975). [Exposition au trichloroéthylène. Bilan de

l'absorption, de l'excrétion et du metabolisme sur des sujets humains]. Arch Mal Prof. 35:397–

407.

Fernández JG, Droz PO, Humbert BE, Caperos JR (1977). Trichloroethylene exposure simulation of

uptake, excretion, and metabolism using a mathematical model. Br J Ind Med. 34:43–55.

Fisher JW (2000). Physiologically based pharmacokinetic models for trichloroethylene and its oxidative

metabolites. Environ Health Perspect. 108(Suppl. 2):265–73.

Fisher JW, Whittaker TA, Taylor DH (1989). Physiologically based pharmacokinetic modeling of the

pregnant rat: a multiroute exposure model for trichloroethylene and its metabolite,

trichloroacetic acid. Toxicol Appl Pharmacol. 99:395–414.

Fisher JW, Andersen ME, Clewell HJ, Taylor D (1987). Kinetics of trichloroethylene in pregnant and

lactating rats and rat pups. Toxicology. 47:206–7.

Fisher JW, Gargas ML, Allen BC, Andersen ME (1991). Physiologically based pharmacokinetic

modeling with trichloroethylene and its metabolite, trichloroacetic acid, in the rat and mouse.

Toxicol Appl Pharmacol. 109:183–95.

Fisher JW, Channel SR, Eggers JS, Johnson PD, MacMahon KL, Goodyear CD, et al. (2001).

Trichloroethylene, trichloroacetic acid, and dichloroacetic acid: do they affect fetal rat heart

development? Int J Toxicol. 20:257–67.

Fleming-Jones ME, Smith RE (2003). Volatile organic compounds in foods: a five-year study. J Agric

Food Chem. 51:8120–7.

Forkert PG, Lash L, Tardif R, Tanphaichitr N, Vandevoort C, Moussa M (2003). Identification of

trichloroethylene and its metabolites in human seminal fluid of workers exposed to

trichloroethylene. Drug Metab Dispos. 31(3):306–11.

Fronk CA (1987). Destruction of volatile organic contaminants in drinking water by ozone treatment.

Ozone Sci Eng. 9(3):265–88.

Fukuda K, Takemoto K, Tsuruta H (1983). Inhalation carcinogenicity of trichloroethylene in mice and

rats. Ind Health. 21:243–54.

Goeptar AR, Commandeur JN, van Ommen B, van Bladeren PJ, Vermeulen NP (1995). Metabolism

and kinetics of trichloroethylene in relation to toxicity and carcinogenicity: relevance of the

mercapturic acid pathway. Chem Res Toxicol. 8:3–21.

Page 46: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

38

Goldberg SJ, Lebowitz MD, Graver EJ, Hicks S (1990). An association of human congenital cardiac

malformations and drinking water contaminants. J Am Coll Cardiol. 16:155–64.

Goldsworthy TL, Popp JA (1987). Chlorinated hydrocarbon-induced peroxisomal enzyme activity in

relation to species and organ carcinogenicity. Toxicol Appl. Pharmacol. 86:225–33.

Goldsworthy TL, Lyght O, Burnett VL, Popp JA (1988). Potential role of alpha-2 mu-globulin, protein

droplet accumulation, and cell replication in the renal carcinogenicity of rats exposed to

trichloroethylene, perchloroethylene, and pentachloroethane. Toxicol Appl Pharmacol.

96:367–79.

Green T (2000). Pulmonary toxicity and carcinogenicity of trichloroethylene: species differences and

modes of action. Environ Health Perspect. 108(Suppl. 2):261–4.

Green T, Prout MS (1985). Species differences in response to trichloroethylene. II. Biotransformation

in rats and mice. Toxicol Appl Pharmacol. 79:401–11.

Green T, Mainwaring GW, Foster JR (1997). Trichloroethylene-induced mouse lung tumors: studies of

the mode of action and comparisons between species. Fundam Appl Toxicol. 37:125–30.

Gross R, Termaath S (1985). Packed tower aeration strips trichloroethylene from groundwater.

Environmental Progress. 4:119.

Gu ZW, Sele B, Chmara D, Jalbert P, Vincent M, Vincent M, et al. (1981a). [Effets du trichloroéthylène

et de ses métabolites sur le taux d’échanges de chromatides-soeurs.] Ann Génét. 24:105–6.

Gu ZW, Sele B, Jalbert P, Vincent M, Vincent F, Marka C, et al. (1981b). [Induction d’échanges entre

les chromatides soeurs (SCE) par le trichloréthylène et ses métabolites.] Toxicol Eur Res.

111(2):63–67.

Hack CE, Chiu WA, Jay Zhao Q, Clewell HJ (2006). Bayesian population analysis of a harmonized

physiologically based pharmacokinetic model of trichloroethylene and its metabolites. Regul

Toxicol Pharmacol. 46(1):63–83.

Hand DW, Crittenden JC, Miller JM, Gehin JL (1988). Performance of air stripping and granular

activated carbon for synthetic organic chemical and volatile organic chemical removal from

groundwater. Cincinnati, Ohio: United States Environmental Protection Agency (EPA 600/J-

88/053).

Hand DW, Herlevich JA, Perram DL, Crittenden JC (1994). Synthetic adsorbent versus GAC for TCE

removal. J Am Water Works Assoc. 86(8):64–72.

Healy TFJ, Poole TR, Hopper A (1982). Rat fetal development and maternal exposure to

trichloroethylene at 100 ppm. Br J Anaesth. 54:337–41.

Health Canada (2005) Guidelines for Canadian Drinking Water Quality: Supporting Documentation —

Trichloroethylene. Water Quality and Health Bureau, Healthy Environments and Consumer

Safety Branch, Health Canada, Ottawa, Ontario.

Henschler DH, Romen W, Elsässer HM, Reichert D, Eder E, Radwan Z (1980). Carcinogenicity study

of trichloroethylene by long-term inhalation in three animal species. Arch Toxicol. 43:237–48.

Henschler D, Elsässer H, Romen W, Eder E (1984). Carcinogenicity study of trichloroethylene, with

and without epoxide stabilizers, in mice. J Cancer Res Clin Oncol. 107:149–56.

Henschler D, Vamvakas S, Lammert M, Dekant W, Kraus B, Thomas B, et al. (1995). Increased

incidence of renal cell tumors in a cohort of cardboard workers exposed to trichloroethylene.

Arch Toxicol. 69:291–9.

Hine J, Mookerjee PK (1975). The intrinsic hydrophilic character of organic compounds: correlations

in terms of structural contributions. J Org Chem. 40:292–303.

Holmberg PC, Nurminen M (1980). Congenital defects of the central nervous system and occupational

factors during pregnancy: a case–referent study. Am J Ind Med. 1:167–76.

Page 47: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

39

Holmberg PC, Hernberg S, Kurppa K, Rantala K, Riala R (1982). Oral clefts and organic solvent

exposure during pregnancy. Int Arch Occup Environ Health. 50:371–6.

IARC (International Agency for Research on Cancer) (1995). Trichloroethylene. In: Drycleaning, some

chlorinated solvents, and other industrial chemicals. Lyon, France: International Agency for

Research on Cancer (IARC Monographs on the Evaluation of Carcinogenic Risks to Humans,

Vol. 63).

IARC (International Agency for Research on Cancer) (2014). Trichloroethylene, tetrachloroethylene,

and some chlorinated agents. Lyon, France: IARC (IARC Monographs on the Evaluation of

Carcinogenic Risks to Humans, Vol. 106).

IPCS (International Programme on Chemical Safety) (1985). Trichloroethylene. Geneva: World Health

Organization (Environmental Health Criteria No. 50).

ISO (International Organization for Standardization) (1997). ISO 10301:1997. Water quality:

determination of highly volatile halogenated hydrocarbons – gas-chromatographic methods.

Geneva: ISO.

Jakobson I, Wahlberg JE, Holmberg B, Johansson G (1982). Uptake via the blood and elimination of

10 organic solvents following epicutaneous exposure of anesthetized guinea pigs. Toxicol Appl

Pharmacol. 63:181–7.

Johnson PD, Goldberg SJ, Mays MZ, Dawson BV (2003). Threshold of trichloroethylene contamination

in maternal drinking waters affecting fetal heart development in the rat. Environ Health

Perspect. 111(3):289–92.

Karami S, Lan Q, Rothman N, Stewart PA, Lee KM, Vermeulen R, et al. (2012). Occupational

trichloroethylene exposure and kidney cancer risk: a meta-analysis. Occup Environ

Med. 69(12):858–67.

Karimi AA, Redman JA, Glaze WH, Stolarik GF (1997). Evaluating an AOP for TCE and PCE removal.

J Am Water Works Assoc. 89(8):41–53.

Keil DE, Peden-Adams MM, Wallace S, Ruiz P, Gilkeson GS (2009). Assessment of trichloroethylene

(trichloroethylene) exposure in murine strains genetically-prone and non-prone to develop

autoimmune disease. J Environ Sci Health A Tox Hazard Subst Environ Eng. 44:443–53.

Kerbey AL, Randle PJ, Cooper RH, Whitehouse S, Pask HT, Denton RM (1976). Regulation of

pyruvate dehydrogenase in rat heart. Biochem J. 154:327–48.

Keys DA, Bruckner JV, Muralidhara S, Fisher JW (2003). Tissue dosimetry expansion and cross-

validation of rat and mouse physiologically based pharmacokinetic models for

trichloroethylene. Toxicol Sci. 76(1):35–50.

Kezic S, Monster AC, Krüse J, Verberk MM (2000). Skin absorption of some vaporous solvents in

volunteers. Int Arch Occup Environ Health. 73:415–22.

Kim D, Ghanayem BI (2006). Comparative metabolism and disposition of trichloroethylene in Cyp2e1-

/- and wild-type mice. Drug Metab Dispos. 34:2020–7.

Kimmerle G, Eben A (1973). Metabolism, excretion and toxicology of trichloroethylene after

inhalation. 1. Experimental exposure on rats. Archiv für Toxikologie. 30:115–26.

Kjellstrand P, Kanje M, Månsson L, Bjerkemo M, Mortensen I, Lanke J, et al. (1981).

Trichloroethylene: effects on body and organ weights in mice, rats, and gerbils. Toxicology.

21:105–15.

Kjellstrand P, Kanje M, Månsson L, Bjerkemo M, Mortensen I, Lanke J, et al. (1983a). Effects of

continuous trichloroethylene inhalation on different strains of mice. Acta Pharmacol Toxicol

(Copenh). 53:369–74.

Page 48: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

40

Kjellstrand P, Holmquist B, Alm P, Kanje M, Romare S, Jonsson I, et al. (1983b). Trichloroethylene:

further studies of the effects on body and organ weights and plasma butyrylcholinesterase

activity in mice. Acta Pharmacol Toxicol (Copenh). 53:375–84.

Kleinfeld M, Tabershaw IR (1954). Trichloroethylene toxicity: report of five fatal cases. AMA Arch

Ind Hyg Occup Med. 10:141–3.

Krishnan K (2003). Evaluation of the relative importance of dermal and inhalation routes of exposure

for trichloroethylene. Contract report submitted to Water Quality and Health Bureau, Safe

Environments Programme, Health Canada, Ottawa, Ontario.

Krishnan K, Carrier R (2008). Approaches for evaluating the relevance of multiroute exposures in

establishing guideline values for drinking water contaminants. J Environ Sci Health C Environ

Carcinog Ecotoxicol Rev. 26(3):300–16.

Krishnan K, Carrier R (2013). The use of exposure source allocation factor in the risk assessment of

drinking-water contaminants. J Toxicol Environ Health B Crit Rev. 16:39–51.

Kruithof JC, Koppers HMM (1989). Experiences with groundwater treatment and disposal of the

eliminated substances in the Netherlands. Aqua. 38:207–16.

Kruithof JC, Hess AF, Manwaring JF, Beville PB (1985). Removal of organic contaminants from

drinking water. Aqua. 35(2):89–99.

Laparé S, Tardif R, Brodeur J (1995). Effect of various exposure scenarios on the biological monitoring

of organic solvents in alveolar air. II. 1,1,1-trichloroethane and trichloroethylene. Int Arch

Occup Environ Health. 67:375–94.

Lash L, Parker J, Scott C (2000). Modes of action of trichloroethylene for kidney tumorigenesis.

Environ Health Perspect. 108(Suppl. 2):225–40.

Lash LH, Putt DA, Parker JC (2006). Metabolism and tissue distribution of orally administered

trichloroethylene in male and female rats: identification of glutathione- and cytochrome P-450-

derived metabolites in liver, kidney, blood, and urine. J Toxicol Environ Health A. 69:1285–

309.

Lash LH, Chiu WA, Guyton KZ, Rusyn I (2014). Trichloroethylene biotransformation and its role in

mutagenicity, carcinogenicity and target organ toxicity. Mutat Res Rev Mutat Res. 762:22–36.

Leighton DT, Calo JM (1981). Distribution coefficients of chlorinated hydrocarbons in dilute air–water

systems for groundwater contamination applications. J Chem Eng Data. 26:383–5.

Leong KJ, Schwetz BA, Gehring PJ (1975). Embryo and fetotoxicity of inhaled trichloroethylene,

perchloroethylene, methylchloroform and methylene chloride in mice and rats. Toxicol Appl

Pharmacol. 33:136–7.

Lindstrom AB, Pleil JD (1996). A methodological approach for exposure assessment studies in

residence using volatile organic compound-contaminated water. J Air Waste Manag Assoc.

46(11):1058–66.

Mahle DA, Gearhart JM, Grigsby CC, Mattie DR, Barton HA, Lipscombe JC, et al. (2007). Age-

dependent partition coefficients for a mixture of volatile organic solvents in Sprague–Dawley

rats and humans. J Toxicol Environ Health A. 70:1745–51.

Major DW, Hodgins EW, Butler BJ (1991). Field laboratory evidence of in situ biotransformation of

tetrachloroethene to ethene and ethane at a chemical transfer facility in North Toronto. In:

Hinchee RE, Olfenbuttel RF, editors. On-site bioreclamation: processes for xenobiotic and

hydrocarbon treatment. Boston: Butterworth-Heinemann.

Maltoni C, Lefemine G, Cotti G (1986). Archives of research on industrial carcinogenesis. Vol. V.

Experimental research on trichloroethylene carcinogenesis. Princeton, New Jersey: Princeton

Scientific Publishing Company.

Page 49: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

41

Maltoni C, Lefemine G, Cotti G, Perino G (1988). Long-term carcinogenic bioassays on

trichloroethylene administered by inhalation to Sprague–Dawley rats and Swiss and B6C3F1

mice. Ann N Y Acad Sci. 534:316–42.

Mandel JH, Kelsh MA, Mink PJ, Alexander DD, Kalmes RM, Weingart M (2006). Occupational

trichloroethylene exposure and non-Hodgkin’s lymphoma: a meta-analysis and review. Occup

Environ Med. 63:597–607.

McCarthy MC, Hafner HR, Chinkin LR, Charrier JG (2007). Temporal variability of selected air toxics

in the United States. Atmos Environ. 41(34):7180–94.

McKinnon RJ, Dykesen JE (1984). Removing organics from groundwater through aeration plus GAC.

J Am Water Works Assoc. 76(5):42–7.

McKone TE (1987). Human exposure to volatile organic compounds in household tap water: the indoor

inhalation pathway. Environ Sci Technol. 21:1194–1201.

McLaughlin JK, Blot WJ (1997). A critical review of epidemiology studies of trichloroethylene and

perchloroethylene and risk of renal-cell cancer. Int Arch Occup Environ Health. 70:222–31.

McNeill WC (1979). Trichloroethylene. In: Kirk–Othmer encyclopedia of chemical technology, 3rd

ed., Vol. 5. New York: John Wiley & Sons, 745–53.

Melnick RL, Jameson CW, Goehl TJ, Maronpot RR, Collins BJ, Greenwell A, et al. (1987). Application

of microencapsulation for toxicology studies. II. Toxicity of microencapsulated

trichloroethylene in Fischer 344 rats. Fundam Appl Toxicol. 9:432–42.

Mitoma C, Steeger T, Jackson SE, Wheeler KP, Rogers JH, Milman HA (1985). Metabolic disposition

study of chlorinated hydrocarbons in rats and mice. Drug Chem Toxicol. 8:183–94.

Monster AC (1979). Difference in uptake, elimination, and metabolism in exposure to trichloroethylene,

1,1,1-trichloroethane and tetrachloroethylene. Int Arch Occup Environ Health. 42:311–17.

Monster AC, Boersma C, Duba WC (1976). Pharmacokinetics of trichloroethylene in volunteers,

influence of workload and exposure concentration. Int Arch Occup Environ Health. 38:87–102.

Monster AC, Boersma C, Duba WC (1979). Kinetics of trichloroethylene in repeated exposure of

volunteers. Int Arch Occup Environ Health. 42:283–92.

Moore LE, Boffetta P, Karami S, Brennan P, Stewart PS, Hung R, et al. (2010). Occupational

trichloroethylene exposure and renal carcinoma risk: evidence of genetic susceptibility by

reductive metabolism gene variants. Cancer Res. 70:6527–36.

Moore MM, Harrington-Brock K (2000). Mutagenicity of trichloroethylene and its metabolites:

implications for risk assessment of trichloroethylene. Environ Health Perspect. 108(Suppl.

2):215–25.

Muller G, Spassovski M, Henschler D (1975). Metabolism of trichloroethylene in man. III. Interaction

of trichloroethylene and ethanol. Arch Toxicol. 33:173–89.

Murata K, Inoue O, Akutsu M, Iwata T ( 2010). Neuromotor effects of short-term and long-term

exposures to trichloroethylene in workers. Am J Ind Med. 53(9):915–21.

Nagaya T, Ishikawa N, Hata H (1989). Sister-chromatid exchanges in lymphocytes of workers exposed

to trichloroethylene. Mutat Res. 222(3):279–82.

Nakajima T, Wang RS, Elovaara E, Park SS, Gelboin HV, Vainio H (1993). Cytochrome P450-related

differences between rats and mice in the metabolism of benzene, toluene and trichloroethylene

in liver microsomes. Biochem Pharmacol. 45:1079–85.

Narotsky MG, Kavlock RJ (1995). A multidisciplinary approach to toxicological screening. II.

Developmental toxicity. J Toxicol Environ Health. 45:145–71.

Page 50: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

42

Narotsky MG, Weller EA, Chinchilli VM, Kavlock RJ (1995). Nonadditive developmental toxicity in

mixtures of trichloroethylene, di(2- ethylhexyl) phthalate, and heptachlor in a 5 × 5 × 5 design.

Fundam Appl Toxicol. 27:203–16.

NCI (National Cancer Institute) (1976). Carcinogenesis bioassay of trichloroethylene. Bethesda,

Maryland: National Institutes of Health (NCI-CGTR-2, NIH 76-802).

Nomiyama K, Nomiyama H (1971). Metabolism of trichloroethylene in humans: sex difference in

urinary excretion of trichloroacetic acid and trichloroethanol. Int Arch Arbeitsmed. 28:37–48.

NRC (National Research Council) (2009). Contaminated water supplies at Camp Lejeune: assessing

the potential health effects. Washington, DC: National Academies Press.

NTP (National Toxicology Program) (1983). NTP technical report on the carcinogenesis studies of

trichloroethylene (without epichlorohydrin) (CAS No. 79-01-6) in F344/N rats and B6C3F1

mice (gavage studies). Draft report. Research Triangle Park, North Carolina: National Institutes

of Health (NIH Publication No. 83-1799).

NTP (National Toxicology Program) (1985). Trichloroethylene: reproduction and fertility assessment

in CD-1 mice when administered in the feed. Research Triangle Park, North Carolina: National

Institutes of Health (NIH Publication No. 86-068).

NTP (National Toxicology Program) (1986). Trichloroethylene: reproduction and fertility assessment

in F344 rats when administered in the feed. Final report. Research Triangle Park, North

Carolina: National Institutes of Health (NIH Publication No. 86-085).

NTP (National Toxicology Program) (1988). Toxicology and carcinogenesis studies of

trichloroethylene (CAS No. 79-01-6) in four strains of rats (ACI, August, Marshall, Osborne-

Mendel) (gavage studies). Research Triangle Park, North Carolina: National Institutes of

Health (NTP Technical Report Series No. 273; NIH Publication No. 88- 2525).

NTP (National Toxicology Program) (1990). Carcinogenesis studies of trichloroethylene (without

epichlorohydrin) (CAS No. 79-01-6) in F344/N rats and B6C3F1 mice (gavage studies).

Research Triangle Park, North Carolina: National Institutes of Health (NTP Technical Report

Series No. 243).

Odum J, Foster JR, Green T (1992). A mechanism for the development of Clara cell lesions in the

mouse lung after exposure to trichloroethylene. Chem Biol Interact. 83:135–53.

Ömür-Özbek P, Gallagher DL, Dietrich AM (2011). Determining human exposure and sensory

detection of odorous compounds released during showering. Environ Sci Technol. 45:468–73.

Opdam JJ (1989). Intra and interindividual variability in the kinetics of a poorly and highly metabolizing

solvent. Br J Ind Med. 46:831–45.

OSHA (Occupational Safety and Health Administration) (1993). Air contaminants final rule. Fed

Regist. 58:35338–51.

Pailard H, Brunet R, Dore M (1987). Application of oxidation by a combined ozone/ultraviolet radiation

system to the treatment of natural water. Ozone Sci Eng. 9(4):391–418.

Pant P, Pant S (2010). A review: advances in microbial remediation of trichloroethylene (TCE). J

Environ Sci. 22:116–26.

Parchman LG, Magee PN (1982). Metabolism of [14C]trichloroethylene to 14CO2 and interaction of a

metabolite with liver DNA in rats and mice. J Toxicol Environ Health. 9:797–813.

Pastino GM, Yap WY, Carroquino M (2000). Human variability and susceptibility to trichloroethylene.

Environ Health Perspect. 108(Suppl. 2):201–4.

Pearson CR, McConnell G (1975). Chlorinated C1 and C2 hydrocarbons in the marine environment.

Proc R Soc Lond. 189:305–32.

Page 51: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

43

Peden-Adams MM, Eudaly JG, Heesemann LM, Smythe J, Miller J, Gilkeson GS, et al. (2006).

Developmental immunotoxicity of trichloroethylene (trichloroethylene): studies in B6C3F1

mice. J Environ Sci Health A Tox Hazard Subst Environ Eng. 41:249–71.

Pellizzari ED, Hartwell TD, Harris BS (1982). Purgeable organic compounds in mother’s milk. Bull

Environ Contam Toxicol. 28:322–8.

Perbellini L, Olivato D, Zedde A, Miglioranzi R (1991). Acute trichloroethylene poisoning by

ingestion: clinical and pharmacokinetic aspects. Intensive Care Med. 17:234–5.

Poet TS, Corley RA, Thrall KD, Edwards JA, Tanojo H, Weitz KK, et al. (2000). Assessment of the

percutaneous absorption of trichloroethylene in rats and humans using MS/MS real-time breath

analysis and physiologically based pharmacokinetic modeling. Toxicol Sci. 56(1):61–72.

Prout MS, Provan WM, Green T (1985). Species differences in response to trichloroethylene:

pharmacokinetics in rats and mice. Toxicol Appl Pharmacol. 79:389–400.

Radican L, Blair A, Stewart P, Wartenberg D (2008). Mortality of aircraft maintenance workers exposed

to trichloroethylene and other hydrocarbons and chemicals: extended follow-up. J Occup

Environ Med. 50:1306–19.

Rastkari N, Yunesian M, Ahmadkhaniha R (2011). Exposure assessment to trichloroethylene and

perchloroethylene for workers in the dry cleaning industry. Bull Environ Contam Toxicol.

86(4):363–7.

Raunio H, Husgafvel-Pursiainen K, Anttila S, Hietanen E, Hirvonen A, Pelkonen O (1995). Diagnosis

of polymorphisms in carcinogen-activating and inactivating enzymes and cancer susceptibility:

a review. Gene. 159:113–21.

Rosa C (2003). Exposure to trichloroethylene in an insignia manufacturing facility. Appl Occup

Environ Hyg. 18(9):646–8.

Rothman KJ (1986). Modern epidemiology. Boston, Massachusetts: Little Brown and Company.

Rouisse L, Chakrabarti SK (1986). Dose-dependent metabolism of trichloroethylene and its relevance

to hepatotoxicity in rats. Environ Res. 40:450–8.

Rusyn I, Chiu WA, Lash LH, Kromhout H, Hansen J, Guyton KZ (2014). Trichloroethylene:

mechanistic, epidemiologic and other supporting evidence of carcinogenic hazard. Pharmacol

Ther. 141(1):1–32.

Ruth JH (1986). Odor threshold and irritation levels of several chemical substances: a review. Am Ind

Hyg Assoc J. 47:148–51.

Sanders VM, Tucke AN, White KL Jr, Kauffman BM, Hallett P, Carchman RA, et al. (1982). Humoral

and cell-mediated immune status in mice exposed to trichloroethylene in the drinking water.

Toxicol Appl Pharmacol. 62:358–68.

Sato A, Nakajima T (1978). Differences following skin or inhalation exposure in the absorption and

excretion kinetics of trichloroethylene and toluene. Br J Ind Med. 35:43–9.

Sato A, Nakajima T, Fuliwara Y, Murayama N (1977). A pharmacokinetic model to study the excretion

of trichloroethylene and its metabolites after an inhalation exposure. Br J Ind Med. 34:56–63.

Schroll R, Bierling B, Cao G, Dörfler U, Lahaniati M, Langenbach T, et al. (1994). Uptake pathways

of organic chemicals from soil by agricultural plants. Chemosphere. 28:297–303.

Schwetz BA, Leong BKJ, Gehring PJ (1975). The effect of maternally inhaled trichloroethylene,

perchloroethylene, methyl chloroform, and methylene chloride on embryonal and fetal

development in mice and rats. Toxicol Appl Pharmacol. 32:84–96.

Scott CS, Jinot J (2011). Trichloroethylene and cancer: systematic and quantitative review of

epidemiologic evidence for identifying hazards. Int J Environ Res Public Health. 8:4238–72.

Page 52: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

44

Seiji K, Jin C, Wantanabe T, Nakatsuka H, Ikeda M (1990). Sister chromatid exchanges in peripheral

lymphocytes of workers exposed to benzene, trichloroethylene, or tetrachloroethylene, with

reference to smoking habits. Int Arch Occup Environ Health. 62(2):171–6.

Shuong D, Ito A, Ohkawa A (2001). Removal of trichloroethylene from water by aeration,

pervaporation and membrane distillation. Journal of Chemical Engineering of Japan.

34(8):1069–73.

Siegel J, Jones RA, Coon RA, Lyon JP (1971). Effects on experimental animals of acute repeated and

continuous inhalation exposures to dichloroacetylene mixtures. Toxicol Appl Pharmacol.

18:168–74.

Simmons JE, Boyes WK, Bushnell PJ, Raymer JH, Limsakun T, McDonald A, et al. (2002). A

physiologically based pharmacokinetic model for trichloroethylene in the male long-evans rat.

Toxicol Sci. 69:3–15.

Simon P, Mitchell J (1992). The aeration removal of volatile organochlorine compounds from a well

water site near London. Aqua, 41(6):322–9.

Skender L, Karacic V, Bosner B, Prpić-Majić D (1994). Assessment of urban population exposure to

trichloroethylene and tetrachloroethylene by means of biological monitoring. Arch Environ

Health. 49:445–51.

Slagle D (1990). Ozonation as an attractive option for groundwater systems. Public Works. 121(3):45–

6.

Smith MK, Randall JL, Read EJ (1992). Developmental toxicity of dichloroacetate in the rat.

Teratology. 46:217–23.

Spirtas R, Stewart PA, Lee JS, Marano DE, Forbes CD, Grauman DJ, et al. (1991). Retrospective cohort

mortality study of workers at an aircraft maintenance facility. I. Epidemiological results. Br J

Ind Med. 48:515–30.

Stenner RD, Merdink JL, Stevens DK, Springer DL, Bull RJ (1997). Enterohepatic recirculation of

trichloroethanol glucuronide as a significant source of trichloroacetate in the metabolism of

trichloroethylene. Drug Metab Dispos. 25:529–35.

Stenner RD, Merdink JL, Fisher JW, Bull RJ (1998). Physiologically-based pharmacokinetic model for

trichloroethylene considering enterohepatic circulation of major metabolites. Risk Anal.

18:261–9.

Stewart PA, Lee JS, Marano DE (1991). Retrospective cohort mortality study of workers at an aircraft

maintenance facility: II. Exposures and their assessment. Br J Ind Med. 48:531–7.

Stewart RD, Dodd HC (1964). Absorption of carbon tetrachloride, trichloroethylene,

tetrachloroethylene, methylene chloride, and 1,1,1-trichloroethane through human skin. Am

Ind Hyg Assoc J. 25:439–46.

Stott WT, Quast JF, Watanabe PG (1982). The pharmacokinetics and macromolecular interactions of

trichloroethylene in mice and rats. Toxicol Appl Pharmacol. 62:137–51.

Styles JA, Wyatt I, Coutts C (1991). Trichloroacetic acid: studies on uptake and effects on hepatic DNA

and liver growth in mouse. Carcinogenesis. 12:1715–19.

Tao L, Yang S, Xie M, Kramer PM, Pereira MA (2000a). Hypomethylation and overexpression of c-

jun and c-myc protooncogenes and increased DNA methyltransferase activity in dichloroacetic

and trichloroacetic acid-promoted mouse liver tumors. Cancer Lett. 158:185–93.

Tao L, Yang S, Xie M, Kramer PM, Pereira MA (2000b). Effect of trichloroethylene and its metabolites,

dichloroacetic acid and trichloroacetic acid, on the methylation and expression of c-jun and c-

myc protooncogenes in mouse liver: prevention by methionine. Toxicol Sci. 54:399–407.

Page 53: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

45

Thrall KD, Poet TS (2000). Determination of biokinetic interactions in chemical mixtures using

realtime breath analysis and physiologically based pharmacokinetic modeling. J Toxicol

Environ Health A. 59(8):653–70.

Topudurti K, Keefe M, Wooliever P, Lewis N (1994). Field evaluation of perox-pure chemical

oxidation technology. Water Sci Technol. 30(7):95–104.

Tsuruta H (1978). Percutaneous absorption of trichloroethylene in mice. Industrial Health, 15:145–51.

Tucker AN, Sanders VM, Barnes DW, Bradshaw TJ, White KL Jr, Sain LE, et al. (1982). Toxicology

of trichloroethylene in the mouse. Toxicol Appl Pharmacol. 62:351–7.

Urano K (1991). Adsorption of chlorinated organic compounds on activated carbon from water. Water

Res. 25(12):1459–64.

US EPA (United States Environmental Protection Agency) (1985a). Health assessment document for

trichloroethylene. Washington, DC: US EPA (EPA/600/8-82/006F).

US EPA (United States Environmental Protection Agency) (1985b). National primary drinking water

regulations, volatile synthetic organic compounds. Fed Regist. 50(219):46902.

US EPA (United States Environmental Protection Agency) (1999). UCMR List 1 and List 2 chemical

analytical methods and quality control manual. Washington, DC: US EPA.

US EPA (United States Environmental Protection Agency) (2010). EPI Suite results for CAS 000079-

01-6. Washington, DC: US EPA.

US EPA (United States Environmental Protection Agency) (2011b). Toxicological review for

trichloroethylene. Washington, DC: US EPA.

US EPA (United States Environmental Protection Agency) (2014). TSCA work plan chemical risk

assessment. Trichloroethylene: Degreasing, spot cleaning, and arts & crafts uses. Washington,

DC: US EPA (EPA740-R1-4002).

US EPA (United States Environmental Protection Agency) (2018). Air Data: access to air pollution

data. Washington, DC: US EPA.

Vaidya VS, Shankar K, Lock EA, Bucci TJ, Mehendale HM (2003). Renal injury and repair following

S-1,2-dichlorovinyl-L-cysteine administration to mice. Toxicol Appl Pharmacol. 188(2):110–

21.

Vamvakas S, Brüning T, Thomasson B, Lammert M, Baumüller A, Bolt HM, et al. (1998). Renal cell

cancer correlated with occupational exposure to trichloroethene. J Cancer Res Clin Oncol.

124:374–82.

Ward PM (2012). Brown and black grease suitability for incorporation into feeds and suitability for

biofuels. J Food Prot. 75:731–7.

Weisel CP, Jo WK (1996). Ingestion, inhalation and dermal exposures to chloroform and

trichloroethene from tap water. Environ Health Perspect. 104:48–51.

WHO (World Health Organization) (2005). Trichloroethene in drinking-water: background document

for development of WHO guidelines for drinking-water quality. Geneva: WHO

(WHO/SDE/WSH/05.08/22).

Wilson PD, Loffredo CA, Correa-Villaseñor A, Ferencz C (1998). Attributable fraction for cardiac

malformations. Am J Epidemiol. 148:414–23.

Wu S, Schaum J (2000). Exposure assessment of trichloroethylene. Environ Health Perspect. 108(Suppl

2):359–63.

Yoshida M, Fukabori S, Hara K, Yuasa H, Nakaaki K, Yamamura Y, et al. (1996). Concentrations of

trichloroethylene and its metabolites in blood and urine after acute poisoning by ingestion. Hum

Exp Toxicol. 15:254–8.

Page 54: Trichloroethene in drinking-water - WHO

Trichloroethene in drinking-water

46

Yuen W, Zimmer J (2001). Manitoba First Nation community water supplies 2000. Report prepared for

First Nations & Inuit Health Branch, Health Canada, by Saskatchewan Research Council,

Regina, Saskatchewan (SRC Publication No. 10497-1C01).

Zenick H, Blackburn K, Hope E, Richdale N, Smith MK (1984). Effects of trichloroethylene exposure

on male reproductive function in rats. Toxicology. 31:237–50.