You are on page 1of 127

Reconsideration of chlorpyrifos

Toxicology update
JUNE 2019
© Australian Pesticides and Veterinary Medicines Authority 2019

Ownership of intellectual property rights in this publication


Unless otherwise noted, copyright (and any other intellectual property rights, if any) in this publication is owned by the Australian Pesticides and
Veterinary Medicines Authority (APVMA).

Creative Commons licence


With the exception of the Coat of Arms and other elements specifically identified, this publication is licensed under a Creative Commons
Attribution 4.0 Australia Licence. This is a standard form agreement that allows you to copy, distribute, transmit and adapt this publication
provided that you attribute the work.

A summary of the licence terms is available from www.creativecommons.org/licenses/by/3.0/au/deed.en. The full licence terms are available
from www.creativecommons.org/licenses/by/3.0/au/legalcode.

The APVMA’s preference is that you attribute this publication (and any approved material sourced from it) using the following wording:

Source: Licensed from the Australian Pesticides and Veterinary Medicines Authority (APVMA) under a Creative Commons Attribution 4.0
Australia Licence.

In referencing this document the Australian Pesticides and Veterinary Medicines Authority should be cited as the author, publisher and copyright
owner.

Cover image: iStockphoto (www.istockphoto.com)


iStockphoto images are not covered by this Creative Commons licence.

Use of the Coat of Arms


The terms under which the Coat of Arms can be used are set out on the Department of the Prime Minister and Cabinet website.

Disclaimer

The material in or linking from this report may contain the views or recommendations of third parties. Third party material does not necessarily

reflect the

views of the APVMA, or indicate a commitment to a particular course of action. There may be links in this document that will transfer you to
external websites. The APVMA does not have responsibility for these websites, nor does linking to or from this document constitute any form of
endorsement. The APVMA is not responsible for any errors, omissions or matters of interpretation in any third-party information contained within
this document.

Comments and enquiries regarding copyright:


Assistant Director, Communications
Australian Pesticides and Veterinary Medicines Authority
PO Box 6182
KINGSTON ACT 2604 Australia

Telephone: +61 2 6210 4701

Email: communications@apvma.gov.au.

This publication is available from the APVMA website.


CONTENTS

CONTENTS
EXECUTIVE SUMMARY 5

1 INTRODUCTION 7
1.1 Scope and objectives 7
1.2 Objective 1: Evaluation of the 2015–18 published mammalian studies on the putative low dose effects of
chlorpyrifos 8
Literature search and triage strategy 8
In vivo thresholds for blood and brain cholinesterase inhibition 9
Hazard assessment of low dose chlorpyrifos exposure 13
1.3 Objective 2: To evaluate whether the 2015–18 published mammalian studies affects the validity of the
current APVMA health based guidance values for chlorpyrifos 13
1.4 Objective 3: To re-evaluate Coulston et al Safety evaluation of DOWCO 179 in human volunteers. 1972 Dow
AgroSciences Institute of Experimental Pathology and Toxicology, Albany Medical College, Albany, New
York, USA 14
Posology, experimental design and evaluated endpoints 14
Results 15
Study quality 15
No observed effect level 16
1.5 Objective 4: To re-evaluate Kisicki, et al 1999 A rising dose toxicology study to determine the no-
observable-effect-levels (NOEL) for erythrocyte acetylcholinesterase (AChE) inhibition and cholinergic
signs and symptoms of chlorpyrifos at three dose levels. Dow Agrosciences, report No. DR#K-044793-284
17
Posology, experimental design and evaluated endpoints 17
Results 18
Study quality 18
No observed effect level 19
1.6 Objective 5: To propose the new APVMA health based guidance values for chlorpyrifos based on all of the
relevant, currently available information 20
Acceptable daily intake 20
Acute reference dose 22

2 CONCLUSIONS 23

3 REFERENCES 24

APPENDIX 1: LITERATURE SEARCH RESULTS 27

APPENDIX 2: STUDY ABSTRACTS AND STUDY EVALUATIONS 71

List of table
RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Table 1: Summary of online literature search strategy 8


Table 2: Toxicological thresholds for in vivo cholinesterase inhibition in mice 10
Table 3: Toxicological thresholds for peak cholinesterase inhibition in rats based on Marty et al (2012) and Reiss et al
(2012) 11
Table 4: Toxicological thresholds for in vivo rat blood cholinesterase used in this evaluation 12
Table 5: Experimental design of Coulson et al (1972) 14
Table 6: Toxicological thresholds in other studies 20
EXECUTIVE SUMMARY 5

EXECUTIVE SUMMARY
The scope of this 2019 toxicology update was:

 to evaluate the recent emergent published literature regarding the hypothesised adverse effects of low
dose (doses below the threshold for inhibition of blood cholinesterases) chlorpyrifos treatment in vivo

 to re-evaluate the regulatory studies suporting the current APVMA health based guidance values for
chlorpyrifos

 to propose new APVMA health based guidance values for chlorpyrifos.

The specific objectives of this update were:

 to evaluate the recently published (2015–18) in vivo studies in laboratory mammals on the putative
hazards of low dose exposure to chlorpyrifos

 to evaluate whether the results of the recently published (2015–18) in vivo laboratory mammal studies
affect the validity of the current APVMA health based guidance values for chlorpyrifos

 to determine reliability and suitability of the human studies that support the current APVMA health based
guidance values for chlorpyrifos

 to determine if changes to the APVMA health based guidance values for chlorpyrifos were required.

Evaluation of the recently published (2015–18) in vivo mammalian studies on the putative hazards of low dose
exposure to chlorpyrifos:

 a systematic literature search covering the 2015–18 period identified a total of 500 published references.
Using a literature triage strategy, 76 published studies were selected for detailed evaluation. No regulatory
quality studies were identified. Overall, the putative effects of low dose chlorpyrifos (ie doses below the
thresholds for inhibition of blood cholinesterases or ≤ 0.1 mg/kg bw/day in rats) remain poorly evaluated in
human health-relevant in vivo animal models.

Evaluation of whether the recently published (2015–18) in vivo laboratory mammal studies affect the validity of the
current APVMA health based guidance values for chlorpyrifos:

 no new, regulatory quality data was identified

 there is currently little regulatory quality in vivo experimental animal evidence that exposure to chlorpyrifos
at doses below those that inhibit blood cholinesterases results in human health relevant adverse effects.
Critically, the hypothesis that low doses of chlorpyrifos (ie doses below the thresholds for inhibition of
blood cholinesterases or ≤ 0.1 mg/kg bw/day in rats) cause adverse developmental effects remains
incompletely studied

 however, there is now a substantial body of experimental evidence that exposure to chlorpyrifos at levels
that result in detectable inhibition of blood cholinesterases (inhibition of butyryl- and/or
acetylcholinesterase) may be a potentially serious neurodevelopmental and neurobehavioural
developmental health hazard for humans.
6 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Determine the reliability and suitability of the human studies that form the basis for the current APVMA health
based guidance values for chlorpyrifos:

 the single oral dose study in humans that forms the basis for the current APVMA acute reference dose for
chlorpyrifos (0.1 mg/kg bw/day) is regarded as being appropriate, with limitations (low statistical power),
for determination of APVMA’s acute reference dose for chlorpyrifos

 the repeat daily oral dosing study in humans that forms the basis for the current APVMA acceptable daily
intake for chlorpyrifos (0.003 mg/kg bw/day), while of appropriate quality for the era in which it was
conducted, is no longer regarded as being compliant with modern regulatory expectations. Accordingly, it
is no longer regarded as an appropriate basis for determination of the acceptable daily intake for
chlorpyrifos.

Proposed new APVMA health based guidance values for chlorpyrifos:

 the APVMA proposes to set a new acute reference dose based on the single oral dose no observed effect
level for inhibition of erythrocyte cholinesterase in humans of 1 mg/kg bw. Because the acute reference
dose is derived from humans, an interspecies uncertainty factor is not required. APVMA has retained the
full ten-fold intraspecies uncertainty factor. APVMA has also applied an additional uncertainty factor of
100.5 to account for any additional uncertainties associated with the limitations of the underlying study.
Accordingly the proposed new chlorpyrifos acute reference dose is 1/(10 x 10 0.5) ≈ 0.03 mg/kg bw

 the APVMA proposes to set a new acceptable daily intake for chlorpyrifos based on the repeat oral dose
no observed effect level for inhibition of blood cholinesterases (blood acetyl- and butyrylcholinesterases)
in rats of 0.1 mg/kg bw/day. The proposed total intra- and interspecies uncertainty factors are 10 x 10 =
100. Accordingly the proposed new chlorpyrifos acceptable daily intake is 0.1/(10 x 10) = 0.001 mg/kg
bw/day.
INTRODUCTION 7

1 INTRODUCTION

1.1 Scope and objectives


The 2019 toxicology update was intended to supplement the published April 2017 reconsideration of chlorpyrifos:
supplementary toxicology assessment report. The scope of this update was:

 to evaluate the recent (2015–18) emergent published literature regarding the hypothesised adverse effects of
low dose (doses below the threshold for inhibition of butyrylcholinesterase (plasma cholinesterase in humans,
[EC 3.1.1.8]) and erythrocyte acetylcholinesterase [EC 3.1.1.7]) chlorpyrifos exposure in vivo. This evaluation
is necessary because of the emerging hypothesis that the thresholds for inhibition of blood cholinesterases
are inadequately protective surrogate biomarker points of departure for the establishment of human health
based guidance values. The term ‘low dose chlorpyrifos’ within the context of this report is specifically defined
as chlorpyrifos doses below the threshold doses for inhibition of blood cholinesterases

 to re-evaluate the regulatory studies forming the basis of the current APVMA health based guidance values for
chlorpyrifos

 if indicated, to propose the new APVMA health based guidance values for chlorpyrifos.

The specific objectives of this report were as follows:

 to evaluate the 2015–18 published in vivo mammalian studies on the putative hazards of low dose exposure to
chlorpyrifos

 to evaluate whether the 2015–18 published in vivo mammalian studies affects the validity of the current
APVMA health based guidance values for chlorpyrifos

 to re-evaluate Coulson and Goldberg 1972, Safety evaluation of DOWCO 179 in human volunteers, Institute
of Experimental Pathology and Toxicology, Albany Medical College, Albany, New York, USA (Dow
AgroSciences). This study is the basis for the current chlorpyrifos APVMA acceptable daily intake of 0.003
mg/kg bw/day

 to re-evaluate Kisicki, J.C., Seip, C.W. & Combs, M.L. 1999, A rising dose toxicology study to determine the
no-observable-effect-levels (NOEL) for erythrocyte acetylcholinesterase (AChE) inhibition and cholinergic
signs and symptoms of chlorpyrifos at three dose levels, Report No. LLC Study ID: DR#K-044793-284 (Dow
AgroSciences). This study is the basis for the current chlorpyrifos APVMA acute reference dose of 0.1 mg/kg
bw

 to propose the new APVMA health based guidance values for chlorpyrifos based on all of the relevant,
currently available information.
8 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

1.2 Objective 1: Evaluation of the 2015–18 published mammalian


studies on the putative low dose effects of chlorpyrifos

Literature search and triage strategy

The literature search strategy that was used to identify the published studies evaluated in this section is shown in
Table 1.

Table 1: Summary of online literature search strategy

Tool Search characteristics

Combined search criteria (search field) Results

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


39 references
(date of addition in PubMed) AND ‘neuro*’ (all)

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


274 references
(date of addition in PubMed) AND ‘development’ (all)

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


118 references
(date of addition in PubMed) AND ‘behaviour’ (all)

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


120 references
(date of addition in PubMed) AND ‘rat’ (all)

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


58 references
(date of addition in PubMed) AND ‘mouse’ (all)

PubMed* ‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


6 references
(date of addition in PubMed) AND ‘guinea pig’ (all)

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


2 references
(date of addition in PubMed) AND ‘dog’ (all)

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


12 references
(date of addition in PubMed) AND ‘human’ (all) AND ‘behaviour’

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


34 references
(date of addition in PubMed) AND ‘human’ (all) AND ‘development’

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


11 references
(date of addition in PubMed) AND ‘human’ (all) AND ‘neuro*’

‘Chlorpyrifos’ (all) AND ‘2015/01/01 to 2018/12/31’


30 references
(date of addition in PubMed) AND ‘human’ (all) AND ‘model*’

Total including duplicate/replicate references: 704 references


INTRODUCTION 9

Tool Search characteristics

Combined search criteria (search field) Results

Total excluding duplicate/replicate references: 500 references

*PubMed; ncbi.nlm.nih.gov/pubmed

The complete list of identified published references is shown in Appendix 1.

The literature triage strategy applied to the references (identified in Appendix 1) was as follows:

 non-mammalian studies were excluded since the objective was to evaluate the putative low dose effects of
chlorpyrifos in laboratory mammals in vivo

 in vitro studies were excluded since the objective was to evaluate the putative low dose effects of chlorpyrifos
in vivo in laboratory mammals

 studies on overt, high dose, acute chlorpyrifos cholinergic poisoning were excluded since the objective was to
evaluate the putative low dose effects of chlorpyrifos. These papers provided no useful additional data for
setting low dose human health based guidance values

 studies that had inadequate dosimetry data or that inadequately described the doses used were excluded

 studies that utilised multiple chemical co-exposures in the absence of a chlorpyrifos only treatment cohort
were excluded

 studies that were primarily ecotoxicological and/or ecochemical and/or bioremediative in nature were excluded

 analytical method and other method development type studies were excluded

 reviews that provided no additional new data were excluded.

The list of 76 studies that were identified for detailed evaluation following literature triage are shown in Appendix 2.

In vivo thresholds for blood and brain cholinesterase inhibition

Since the objective of this report was to determine if the 2015–18 published studies provide evidence for adverse
effects following low dose chlorpyrifos exposure it was necessary to establish the thresholds for such effects in the
evaluated species. The following sections examine the toxicological thresholds for chlorpyrifos effects on blood
and brain cholinesterases in mice, rats and guinea pigs.

A common theme was identified throughout almost all of the rodent (rats, mice and guinea pig) studies evaluated:
almost all of them deliberately selected chlorpyrifos doses (typically in the 1 to 6 mg/kg bw/day range for repeated
dosing) that are known to produce adverse neurobehavioural effects in the absence of manifestly overt toxicity, ie
study dose selections were designed not to result in deaths, overt clinical signs, overtly adverse pregnancy
outcomes or overtly adverse effects on pup growth. Their objective was very specifically not to evaluate the low
dose response relationships and/or dose thresholds of chlorpyrifos. Accordingly many of these of studies have
limited value in terms of establishing human health based guidance values. However, these studies do recapitulate
and reinforce earlier findings that exposure to chlorpyrifos at levels higher than those that inhibit blood
cholinesterases (ie ‘high dose chlorpyrifos exposure’) results in adverse effects; particularly adverse effects on
10 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

neurodevelopment and neurobehavioural development. Thus these types of studies do have an important role in
guiding public health policy regarding chlorpyrifos.

A second common theme was identified throughout almost all of the evaluated rodent studies: very few of them
evaluated the interaction between dosing and time to peak blood and brain cholinesterase inhibition. Many of the
studies that have made claims of an absence of detectable blood cholinesterase inhibition relied upon non-
concurrent studies and non-concurrent control groups where measurements were not performed at the post-
dosing peak inhibition time point (Marty et al 2012). Thus these studies may have erroneously underestimated the
levels of blood and brain cholinesterase inhibition that were present in the experiments.

Mice

The thresholds for blood and brain cholinesterase in mice exposed to chlorpyrifos have been extensively
evaluated by the FAO/WHO JMPR and APVMA (APVMA 2017, Wagner 1999; summarised in Table 2).

Table 2: Toxicological thresholds for in vivo cholinesterase inhibition in mice

Endpoint Exposure type Threshold type Value


Plasma cholinesterase
Repeated LOEL 0.7 mg/kg bw/d
inhibition
Erythrocyte cholinesterase
Repeated NOEL 0.7 mg/kg bw/d
inhibition
Brain cholinesterase
Repeated NOEL 0.7 mg/kg bw/d
inhibition

Based on Table 2, the blood cholinesterase inhibition threshold used for evaluation of repeated exposure studies in
mice that has been used in this report is 0.7 mg/kg bw/d.

Rats

In a well-conducted series of regulatory quality studies, DOW (2010[a]) and Marty et al (2012) defined the oral
exposure acute and repeated dose NOELs for chlorpyrifos-induced inhibition of blood cholinesterases in adult and
pre-weaning age SD rats. The study of DOW (2010[b]) and Reis et al (2012) used the studies of Marty et al (2012)
and DOW (2010[a]) to calculate the bench mark doses (BMD) and bench mark dose lows (BMDL; based on the
lower 95 per cent confidence interval of the dose producing the bench mark response) for cholinesterase inhibition
in rats. The dose thresholds derived from these studies are shown in Table 3. Notably, at low levels of chlorpyrifos
exposure age had no consistent effect on the toxicological thresholds for cholinesterase inhibition over the pre-
weaning to adult stages of development. The findings also demonstrated that the results of the benchmark dose
(DOW 2010[b], Reis et al 2012) were broadly consistent (within a factor of ≤ 2-fold) with the NOELs derived from
DOW (2010[a]) and Marty et al (2012). Both approaches also demonstrate that the thresholds for inhibition of
blood cholinesterases were consistently lower (ie are potentially health-protective) than the threshold for inhibition
of brain acetylcholinesterase. For the purposes of this report the lowest threshold for cholinesterases derived from
DOW 2010[a], DOW 2010[b], Marty et al (2012) and Reis et al (2012) were used (Table 4).
INTRODUCTION 11

Table 3: Toxicological thresholds for peak cholinesterase inhibition in rats based on Marty et al (2012) and
Reiss et al (2012)

Benchmark dose

Duration of exposure Peak RBC AChE inhibition Peak brain AChE inhibition

BMD20 BMDL* BMD10 BMDL*

Acute dose (mg/kg bw) † 1.0 0.90 1.7 1.3

Repeat dose (mg/kg


0.21 0.19 0.67 0.53
bw/d)†

Age range No observed effect level


Duration Vehicle
n per over
of dosing Peak RBC AChE Peak plasma ChE Peak brain
dose treatment
exposure method inhibition ֍ inhibition AChE† inhibition
period days

Corn oil 8 ♂ 0.5 mg/kg bw ♂ 0.5 mg/kg bw ♂ 2.0 mg/kg bw


Acute† PND 11
gavage sex/dose ♀ 0.5 mg/kg bw ♀ 0.5 mg/kg bw ♀ 2.0 mg/kg bw

8 ♂ 2.0 mg/kg bw ♂ 0.5 mg/kg bw ♂ 2.0 mg/kg bw


Acute† Milk feeding PND 11
sex/dose ♀ 0.5 mg/kg bw ♀ 0.5 mg/kg bw ♀ 2.0 mg/kg bw

Corn oil 8
Acute† ≈ 70 ♀ 0.5 mg/kg bw ♀ 0.5 mg/kg bw ♀ 2.0 mg/kg bw
gavage ♀/dose

Food
8
Acute† 12 h ≈ 70 ♀ 0.5 mg/kg bw ♀ 0.5 mg/kg bw ♀ 2.0 mg/kg bw
♀/dose
feeding

♂ 0.1 mg/kg ♂ 0.5 mg/kg


Repeated† Corn oil 8 bw/d ♂ 0.1 mg/kg bw/d bw/d
PND 11 to 21
(11 days) gavage sex/dose ♀ 0.1 mg/kg ♀ 0.1 mg/kg bw/d ♀ 0.5 mg/kg
bw/d bw/d

Repeated† Corn oil 8 ♀ 0.1 mg/kg ♀ 0.5 mg/kg


≈ 70–80 ♀ 0.1 mg/kg bw/d
(11 days) gavage ♀/dose bw/d bw/d

Abbreviations: RBC = erythrocyte; AChE = acetylcholinesterase; ChE = cholinesterase; BMD = bench mark dose; BMDL =
bench mark dose low; PND = post natal day(s) of age; mg = milligram; bw = body weight; d = day.

*Lower 95% confidence interval of the BMD.

† Measurements were at the time of peak cholinesterase inhibition as determined in a preliminary time-to-peak inhibition
study. For PND 11 pups, the times-to-peak inhibition were 6 h post-dosing for chlorpyrifos in corn oil and 8 h post-dosing for
chlorpyrifos in milk. For adult female rats, the times-to peak inhibition were 8 h post-dosing for chlorpyrifos in corn oil, and 8
h post-exposure for chlorpyrifos in diet.
12 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

֍ Note: butyrylcholinesterase (EC 3.1.1.8) is the predominant cholinesterase in human serum (> 99%), while in rats there is
an approximately equal distribution of acetylcholinesterase (rat equivalent to EC 3.1.1.7) and butyrylcholinesterase (rat
equivalent to EC 3.1.1.8) in plasma.

Table 4: Toxicological thresholds for in vivo rat blood cholinesterase used in this evaluation

Endpoint Threshold type Exposure type Value

Plasma cholinesterase Acute 0.5 mg/kg bw


NOEL
inhibition (all age ranges) Repeated 0.1 mg/kg bw/d

Erythrocyte cholinesterase Acute 0.5 mg/kg bw


NOEL
inhibition (all age ranges) Repeated 0.1 mg/kg bw/d
BMDL Acute 1 mg/kg bw/d*
Brain cholinesterase
NOEL
inhibition (all age ranges) Repeated 0.5 mg/kg bw/d
BMDL
*Rounded down

It should be noted that plasma cholinesterase determinations in rats are a composite measure of both acetyl (rat
equivalent of EC 3.1.1.7) and butyrylcholinesterase (rat equivalent of EC 3.1.1.8) activities whereas human serum
cholinesterase consists of > 99 per cent butyrylcholinesterase (Nolan 1997).

Further, it should be noted that in rats the relationships between blood cholinesterase activities, brain
cholinesterase activities, and chlorpyrifos metabolism have a complex interaction during development (DOW
2010[a], Marty et al 2012). The maturation-associated changes in cholinesterase activities in blood and brain,
when combined with the process of the maturation-associated changes in chlorpyrifos metabolism, results in a
situation where the relative sensitivity of young vs adult rats to chlorpyrifos-induced cholinesterase inhibition is
dose dependent. Young rats were not more sensitive than adults to chlorpyrifos—or chlorpyrifos oxon-induced
cholinesterase inhibition over the lower portion of the chlorpyrifos dose response curve (doses < 2 mg/kg bw/d;
DOW 2010[a], Marty et al 2012). However, at higher chlorpyrifos exposure levels (> 2 mg/kg bw/d), young rats
became more sensitive than adults. APVMA regards the lower portion of the dose response curve (ie doses < 2
mg/kg bw/d) as being more directly relevant to the objective of establishing human health based guidance values
for chlorpyrifos.

In terms of extrapolating rat data to humans, it is also important to note that central nervous system and blood
brain barrier development in rats is altricial (ie rats are born somewhat less mature cf. humans; Adinolfi and
Haddad 1977, Bonati et al 1981, Clancy et al 2007, Vidair 2004). Based on these characteristics, the full term
human brain at birth is approximately equal to the rat brain at post-natal days 14 to 21 days of age (Bayer et al
1993). Thus findings in post-natal day 11 rats are not necessarily directly correlated with neurodevelopment in
perinatal and neonatal humans.

Guinea pigs

The thresholds for inhibition of blood and brain cholinesterases have not been well established in this species. The
relevant thresholds are assumed to be approximately the same as other rodents (ie about the same as rats and
mice).
INTRODUCTION 13

Hazard assessment of low dose chlorpyrifos exposure

The abstracts and evaluation of the 76 evaluated studies is shown in Appendix 2.

Mice

Of the 21 evaluated publications no reliable low dose chlorpyrifos studies were identified.

Rats

Of the 43 evaluated publications only three studies examined the effects of low dose chlorpyrifos treatment
(Gómez-Giménez et al 2017 and 2018, Silva et al 2017). None of these studies were regarded as being of
regulatory quality due to inadequacies of the statistical methods used, lack of blinding of the observer and/or other
techniques to reduce observer bias, lack of assay validation by incorporation of appropriate control groups, lack of
historical control data and insufficiently detailed short published study reports.

Guinea pigs

Of the three evaluated publications no reliable studies that utilised doses below the threshold for blood
cholinesterases were identified. These studies provide no reliable evidence of low dose chlorpyrifos associated
adverse effects and do not disprove the validity of using blood cholinesterases as health-protective biomarkers for
human health risk assessment.

Humans

The small number of human studies evaluated in this report provide no additional information than what has
already been evaluated in the APVMA 2017 supplementary toxicology report.

1.3 Objective 2: To evaluate whether the 2015–18 published


mammalian studies affects the validity of the current APVMA health
based guidance values for chlorpyrifos
Overall there was insufficient evidence in the evalulated 2015–18 published mammalian studies on chlorpyrifos
that proves the hypothesis that thresholds for blood cholinesterase inhibition are inadequately protective points of
departure for the derivation of human health based guidance values for the Australian population. However, the
body of recent data did clearly demonstrate a systematic lack of regulatory quality data regarding putative low
dose chlorpyrifos effects, particularly in the areas of neurological and neurobehavioural development. The dose
response thresholds for such effects have not been adequately studied in vivo in human-relevant models. APVMA
recognises and acknowledges the regulatory uncertainties created by these important data gaps.
14 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

1.4 Objective 3: To re-evaluate Coulston et al Safety evaluation of


DOWCO 179 in human volunteers. 1972 Dow AgroSciences Institute of
Experimental Pathology and Toxicology, Albany Medical College, Albany,
New York, USA
This study evaluated the effects of administering technical grade chlorpyrifos (syn. DOWCO 179) to adult human
males (basic demographic details not supplied) for up to 27 days (depending on dose). The central premise for
limiting the duration of exposure to a maximum of 27 days was the lack of evidence of cumulative, chronic
toxicological effects at the doses used. The separate statistical report (statistical analysis of data from report:
Coulston, F., L. Golberg, and T. Griffen 1972, Safety Evaluation of DOWCO 179 in Human Volunteers, Institute of
Experimental Pathology and Toxicology, Albany Medical College, Albany, New York (Dow AgroSciences)) was also
reviewed.

Posology, experimental design and evaluated endpoints

Human subjects were assigned to four experimental groups. The study report did not state whether assignment to
treatment group was randomised, if any attempts at cohort matching were made or if any other attempts at
controlling potentially confounding variables and bias were made. It is assumed that no such procedures were
incorporated in the study design. Importantly, the study does not appear to have been conducted using double
blinding. Chlorpyrifos was administered in the form of an oral tablet (composition not stated) at the doses of 0
(placebo), 0.014, 0.03, 0.1 mg/kg bw/day with food (breakfast). The duration of treatment at the different doses
and group sizes are summarised in Table 5.

Table 5: Experimental design of Coulson et al (1972)

Dose mg/kg bw/day Duration of treatment days Recovery period days* n


0
49 0 4
(Placebo)

0.014 27 0 4

0.03 20 6 4

0.1 9 18 4
*Twice weekly blood sample collection was continued until blood cholinesterase levels returned to pre-dosing levels.

Before administration of commencement of treatment, the clinical status of each volunteer was determined by
means of a medical examination which included chest X-ray, electrocardiogram, urinalysis, and studies of
haematology and serum chemistry. No baseline neurological or neurobehavioral evaluations were performed.
There was insufficient information presented in the report regarding the criteria for admission to the study,
including whether subjects not assessed to be clinically normal were excluded.

Following commencement of dosing, blood samples were collected twice each week for blood cholinesterase
determination. Additional blood samples were collected weekly for haematology and clinical chemistry evaluations.
Urine samples for evaluation were collected weekly for routine urinalysis and chemical analysis for chlorpyrifos
metabolites. There was insufficient information in the report to determine the frequency with which clinical signs
were recorded. No neurological or neurobehavioural assessments were performed.
INTRODUCTION 15

The statistical methods included curve-fitting procedures in which linear regressions were calculated for the
response of each individual during the administration phase of the study. The slopes of the resulting regressions in
each dosage group were averaged and the slope means were then submitted to a general analysis of variance.

Results

Termination of treatment

Cessation of treatment at 0.1 mg/kg bw/day appears to have occurred after dosing day nine due to detection of an
approximate 64 per cent decrease in mean plasma cholinesterase activity cf. pre-dose baseline levels. Although
not made clear in the report, inhibition of plasma cholinesterase may have been associated with the possible
detection of muscarinic clinical signs and symptoms of anticholinesterase poisoning in one out of the four
treatment subjects. However, insufficient information was provided to allow for a definitive conclusion. No rationale
was given in the report for the timing of termination of treatment at lower doses.

Clinical signs

One in four subjects in the high dose group displayed signs and symptoms consistent with muscarinic
anticholinesterase poisoning (runny nose, blurred vision, and a feeling of faintness) on day nine of treatment.

Haematology, clinical chemistry, erythrocyte cholinesterase determinations and urinalysis

No chlorpyrifos associated adverse effects were detected.

Plasma cholinesterase determinations

The plasma cholinesterase results through to treatment day 27 of the study are shown in Figure 1.

Study quality

The study, while possibly compliant with the standards in the era in which it was conducted, was not conducted in
accordance with basic modern standards and study quality expectations. In particular:

 it was not conducted in accordance with modern good clinical practice standards

 it was not conducted in accordance with minimum current human clinical trial standards

 the study report was short and lacking critical details based on modern expectations and standards

 there were unexplained missing data

 there were likely several unexplained protocol deviations

 there was no quality assurance audit of the study

 the study subjects were selected from a sub-population that is not representative of the broader human
population (including putative sensitive sub-groups)

 only one sex was evaluated

 there was no blinding to treatment


16 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

 there was no randomisation or cohort matching

 the validity of the statistical methods is open to question

 the reasons for the timing of cessation of dosing in some cohorts were not explicitly and clearly reported

 there were no neurological/neurobehavioural assessments

 the statistical power of the report (n = 4) was inadequate.

Overall the study is regarded as Klimisch score 3, ie not reliable. The study does supply some limited, low quality
supportive data given that it was conducted in humans.

No observed effect level

Consistent with previous Australian evaluations of this study the NOEL for this study is 0.03 mg/kg bw/day based
on a lack of significant (p > 0.05) inhibition of plasma cholinesterase at this dose. However the reliability of the
NOEL was regarded as being of very low for the reasons stated above. The NOEL for this study is not sufficiently
reliable as a point of departure for the determination of health based guidance values for chlorpyrifos.
INTRODUCTION 17

8
Control
Plasma Cholinesterase Actibity Units/mL*

7 0.014 mg/kg bw/day


0.03 mg/kg bw/day
6 0.10 mg/kg bw/day
Mean “ SD n = 4

0
Baseline 1Baseline 2 1 3 6 9 10 13 16 18 20 23 27

Day of Sampling (day)

Figure 1: Plasma cholinesterase. There was no significant (p < 0.05) differences betweent the control and 0.014
and 0.03 mg/kg bw/day cohorts. The plasma cholinesterase activity in the 0.03 mg/kg bw/day cohort was
approximately 70 per cent of the pre-exposure baseline level and about 87 per cent of the control cohort when
evaluated on a day to day basis (not significantly different cf. control; p > 0.05). The plasma cholinesterase activity
in the 0.10 mg/kg bw/day cohort was significantly lower (p < 0.05) cf. the control group by study day nine (34 per
cent of the baseline level in this cohort). Recovery of plasma cholinesterase activity in the 0.03 mg/kg bw/day
cohort was completed within three weeks post-dosing. For the 0.1 mg/kg bw/day cohort, recovery was completed
within four weeks.

1.5 Objective 4: To re-evaluate Kisicki, et al 1999 A rising dose toxicology


study to determine the no-observable-effect-levels (NOEL) for erythrocyte
acetylcholinesterase (AChE) inhibition and cholinergic signs and
symptoms of chlorpyrifos at three dose levels. Dow Agrosciences, report
No. DR#K-044793-284

Posology, experimental design and evaluated endpoints

The study design was a prospective, double-blinded, randomised, placebo controlled trial. Before being placed on
study the volunteers were screened for general health according to set criteria and instructed to refrain from
alcohol, strenuous exercise, and prescription medications before and during the study. The subjects were
educated on the signs and symptoms of cholinergic toxicity and were instructed to inform the study physician of
any adverse effects before the commencement of dosing.

Groups of six men and six women (per study phase) aged between 18 and 55 were administered a lactose powder
capsule of chlorpyrifos at oral (overnight fasted) doses of 0, 0.5, 1 or 2 mg/kg bw. The study was conducted in two
phases. During Phase I subjects were dosed at 0, 0.5, or 1 mg/kg bw. The subjects were then examined for
possible chlorpyrifos associated effects. During Phase II subjects were dosed at either 0 or 2 mg/kg bw.
18 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

The subject’s vital signs (blood pressure, pulse, respiration, and temperature) were assessed before dosing and
one, two, four, eight, 12, 24, 48, and 168 hours after treatment and the subjects were questioned regarding their
well-being, symptoms and clinical signs at each sampling time point. During both study phases blood samples
were collected 10 and 0 h before treatment and two, four, eight, 12, 24, 36, 48, 72, 96, 120, 144, and 168 hours
after treatment and analysed for erythrocyte cholinesterase activity and chlorpyrifos and its metabolites. Plasma
cholinesterase levels were not measured. In addition, all urine voided from 48 hours before dosing to 168 hours
after dosing was collected at 12 or six hour intervals and analysed for chlorpyrifos and metabolites.
Electrocardiographic evaluations were performed. Haematology, clinical chemistry, urinalysis and an abbreviated
physical examination were performed at study completion. General health parameters and clinical chemistry was
checked at seven days post-dosing. Detailed neurological and neurobehavioural examinations were not
performed.

Results

Protocol deviations and data gaps

There were no protocol deviations that were likely to have affected the outcome of the study. One male in the
Phase 1 control cohort and one woman receiving 2 mg/kg bw did not provide a complete series of blood and urine
samples.

Study completion

One woman withdrew from the study. This was associated with decreased erythrocyte cholinesterase activity.
Although not stated in the report, it is possible that some clinical signs of anticholinesterase poisoning may have
been present in this individual.

Symptoms, clinical signs, haematology, urinalysis and electrocardiogram

No chlorpyrifos-associated adverse effects were reported.

Erythrocyte cholinesterase levels

The only chlorpyrifos-associated effects that may have occurred were in the female subject dosed at 2 mg/kg bw
who withdrew from the study. In this individual erythrocyte cholinesterase activities cf. pre-dose baseline values
were 98.4 per cent of the pre-treatment value at four hours post-dosing, 77 per cent at eight hours post-dosing, 72
per cent at 12 hours post-dosing, 74 per cent at 24 hours post-dosing, 81 per cent at 36 hours post-dosing and 80
per cent at 48 hours post-dosing.

Study quality

The study was broadly consistent with minimum Phase I (initial investigational trial) and Phase II (exploratory trial)
human clinical trial requirements at the time it was conducted. However, in terms of current modern Phase I
human clinical trial requirements, the study is non-compliant minimum expected requirements as per International
Congress on Harmonisation (ICH) Efficacy Guidelines for the following reasons: 1

1
ich.org/products/guidelines/efficacy/article/efficacy-guidelines.html
INTRODUCTION 19

 the statistical evaluation was not consistent with modern ICH guideline requirements

 low statistical power. The study had an n = 6 where n = minimum of 20 is currently more typically required in
order to meet the minimum requirements of a Phase I trial as per ICH guideline E9. Notably, a minimum
decrease in human erythrocyte cholinesterase of 17.3 per cent is needed to identify a statistical depression
when there are two pre-exposure measurements (Gallo and Lawryk, 1991). On this basis, the study design
was statistically underpowered. Furthermore, modern human clinical trial experience demonstrates that an n
of 6 is insufficient for the evaluation of key safety and tolerability endpoints in humans

 plasma cholinesterase levels were not measured

 no pharmacokinetic evaluation was performed (non-compliant with ICH guidance document E8)

 the study was not compliant with good clinical practice (ICH guidance document E6 (R2)). This includes
insufficient reporting details regarding informed consent and lack of quality assurance.

Despite these limitations, it is important to note that the objective of the study was not to conduct a modern ICH
compliant clinical pharmaceutical trial and the results of this study are meant to be interpreted within the context of
the overall large body of regulatory quality data on chlorpyrifos (Wagner 1999). Accordingly, while APVMA
recognises the scientific quality limitations of the study, the agency still regards Kisicki et al1999 as useful for
setting of the acute reference dose for chlorpyrifos provided: (a) it is considered in conjunction with, and supported
by, the available animal data; and (b) the uncertainties associated with the study outcome (particularly the study’s
limited statistical power compared with modern minimum standards and the lack of neurological and
neurobehavioural examinations) are taken into account. The study remains the best available acute exposure data
in humans that is currently available to the agency.

No observed effect level

Consistent with previous evaluations of this study by WHO/FAO JMPR (Wagner 1999) the NOEL for this study was
1 mg/kg bw based on the absence of inhibition of erythrocyte cholinesterase. This finding is consistent with the
lack of inhibition of blood cholinesterases in a published study where human volunteers were (low statistical power
n = 5; no statistical analysis; regarded as unreliable) administered a single oral 1 mg/kg bw dose of chlorpyrifos
(Griffin et al1999). Despite the uncertainties associated with Kisicki et al1999, its NOEL of 1 mg/kg bw is also
broadly consistent (within a factor of two-fold) of the NOEL for acute inhibition of blood cholinesterases derived
from reliable studies in rats (0.5 mg/kg bw; DOW 2010[a], Marty et al 2012,) which in turn, is more than two fold
lower than the NOEL and BMDL for inhibition of brain cholinesterases in rats (≥ 1.3 mg/kg bw; Dow 2010[a], DOW
2010[b], Marty et al 2012, Reiss et al 2012).
20 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

1.6 Objective 5: To propose the new APVMA health based guidance


values for chlorpyrifos based on all of the relevant, currently available
information

Acceptable daily intake

Point of departure

The APVMA no longer regards the current acceptable daily intake for chlorpyrifos of 0.003 mg/kg bw/day based on
Coulston et al 1972 as reliable for regulatory purposes.

APVMA proposes establishing a new acceptable daily intake based on the recent series of studies in young and
adult rats performed by DOW 2010[a] and Marty et al 2012. The NOEL for inhibition of blood cholinesterases
(erythrocyte cholinesterase as well as plasma cholinesterases) for rats from post-natal day 11 of age to adulthood
was 0.1 mg/kg bw/day (consistently five-fold lower than the threshold for inhibition of brain cholinesterases in this
species). This point of departure is supported by the following toxicological thresholds in other studies that have
been evaluated by the agency (Table 6).

Table 6: Toxicological thresholds in other studies

Reference Study type NOAEL Comments


13 week repeat 0.1 mg/kg bw/day Plasma and erythrocyte cholinesterase
daily oral (dietary) based on inhibition of activities were decreased at doses ≥ 1 mg/kg
Szabo et al 1988 dose toxicity study brain and erythrocyte bw/ day, and the activity of brain
in F344 rats cholinesterases at acetylcholinesterase was decreased at 5 and
higher doses 15 mg/kg bw/day
2 year repeat daily 0.1 mg/kg bw/day NOEL for inhibition of brain cholinesterase
oral (dietary) based on inhibition of was 1 mg/kg bw/day based on consistent,
Young and carcinogenicity erythrocyte and plasma statistically significant (p < 0.05) inhibition
Grandjean 1988 study (OECD Test cholinesterases at at 10 mg/kg bw/day
Guideline No. 451) higher doses
in F344 rats
2 generation 0.1 mg/kg bw/d based NOAEL for inhibition of brain cholinesterase
reporoductive on inhibition of blood and maternal toxicity was 1 mg/kg bw/day.
toxicity study in SD cholinesterases at The NOAEL for developmental effects was
Breslin et al 1991
rats higher doses 1 mg/kg bw/day, and the NOAEL for effects
on fertility and reproductive effects was 5
mg/kg bw/day

Uncertainty factors

In order to account for any remaining uncertainties and for health-protective reasons, APVMA proposes applying
the full ten-fold intra- and inter-species uncertainty factors as is consistent with established regulatory toxicology
practice. Thus, the total uncertainty factor applied is 10 x 10 = 100.

Calculation of the proposed new acceptable daily intake

The proposed new acceptable daily intake is 0.1/(10 x 10) = 0.001 mg/kg bw/day (1 µg/kg bw/day).
INTRODUCTION 21

Reliability of the proposed new acceptable daily intake

The reliability of the proposed new acceptable daily intake is regarded as being substantially lower than usual
APVMA standards due to:

 the very limited body of modern, systematic, regulatory quality data on the putative effects of low dose (ie
doses ≤ 0.1 mg/kg bw/day) chlorpyrifos on complex neurobehavioural and neurobehavioural development in
animal models

 the results of the November 2016 US EPA Chlorpyrifos Revised Human Health Risk Assessment which
indicated that the predicted human neurodevelopmental time weighted average (TWA) chlorpyrifos blood
concentration LOAEL of 0.004 µg/L as the point of departure for derivation of human health based guidance
values.2 This equates to points of departure for dietary exposure of between 0.12 to 0.2 µg/kg bw/day, ie five
to ten-fold lower than the proposed new APVMA acceptable daily intake for chlorpyrifos. US EPA has indicated
that a total uncertainty factor of 100-fold (ten-fold for intraspecies uncertainty and ten-fold for the Food Quality
Protection Act factor) would be applied to the modelled chlorpyrifos blood concentration TWA LOAEL. APVMA
notes that the physiologically based pharmacokinetic modelling (PBPK) approaches which are the basis of the
2016 US EPA evaluation, while likely reliable, are also dependent a single time point measurement of
placental cord blood chlorpyrifos levels obtained as part of the Columbia Centre for Children’s Environmental
Health studies on chlorpyrifos (evaluated in the APVMA 2017 supplementary toxicology report on chlorpyrifos).
APVMA has three major remaining concerns regarding this approach adopted in the 2016 US EPA revised
human health risk assessment:

 no regulatory agency (including APVMA) has been able to independently evaluate the underlying
Columbia Centre for Children’s Environmental Health data (despite numerous attempts by both US EPA
and APVMA). A core principle of APVMA’s human health risk assessment approach is that the agency
must be able to evaluate the quality and reliability of the key data that supports the derivation of human
health based guidance values

 the 2016 US EPA PBPK predicted human neurodevelopmental chlorpyrifos TWA blood concentration
LOAEL of 0.004 µg/L has not been replicated in vivo in regulatory quality, human-relevant studies

 a plausible mode of action for low dose chlorpyrifos effects (particularly at the 2016 US EPA TWA blood
concentration LOAEL of 0.004 µg/L) on neuro- and neurodevelopment has not been fully experimentally
established ie cause and effect based on the Bradford-Hill Criteria has not been fully established

 the very limited body of modern, systematic, regulatory quality data on the putative effects of low dose (ie
doses ≤ 0.1 mg/kg bw/day) chlorpyrifos on complex neurobehavioural and neurobehavioural development
in animal models.

The most thorough approach to resolving APVMA’s concerns regarding the lower than usual reliability of the
proposed new acceptable daily intake for chlorpyrifos is further regulatory quality testing regarding the putative low
dose chlorpyrifos effects on neuro- and neuro-behavioural development. In particular, further dosimetry studies
and low dose mode/mechanism of action data are required. APVMA also notes the critical need for independent
evaluation and quality assurance assessment of relevant data generated by the Columbia Centre for Children’s
Environmental Health studies on chlorpyrifos.

2
regulations.gov/document?D=EPA-HQ-OPP-2015-0653-0454
22 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Acute reference dose

Point of departure

APVMA has elected to retain the human acute, single dose NOEL for inhibition of plasma cholinesterase of
1 mg/kg bw derived from Kisicki et al 1999. This point of departure is supported by the NOEL of 0.5 mg/kg bw for
inhibition of blood cholinesterases in rats (Marty et al 2012). Notably, US EPA has not set a specific acute food
exposure human health based guidance value for chlorpyrifos in its 2016 revised human health risk assessment
since this agency has concluded that their steady state dietary assessment is adequately protective of any acute
food exposures.

Uncertainty factors

APVMA has elected to apply the full ten-fold intra-species uncertainty factor for calculating the acute reference
dose. Since the point of departure was determined in humans, an inter-species uncertainty factor is not required.
Because of the statistical power limitations (small n compared with modern human clinical trial standards) and
other concerns associated with the Kisicki et al 1999 study (described above) APVMA is proposing to apply an
additional uncertainty factor of 100.5-fold to account for any remaining uncertainties. The total uncertainty factor
applied is thus 10 x 100.5.

Calculation of the proposed new acute reference dose

The proposed new acute reference dose is 1/(10 x 10 0.5) ≈ 0.03 mg/kg bw (30 µg/kg bw).

Reliability of the proposed new acute reference dose

With the addition of the additional uncertainty factor of 10 0.5 to account for any additional uncertainties associated
with the Kisicki et al 1999 study the proposed new acute reference dose is regarded as being adequately health
protective. Based on the currently available body of data there is little regulatory quality evidence that a single,
acute, low dose exposure to chlorpyrifos at this level would result in adverse effects in humans (including
susceptible sub-populations). However, APVMA notes that there is only a very limited body of modern, systematic,
regulatory quality data on the putative acute effects of low dose (ie doses ≤ 0.1 mg/kg bw/day) chlorpyrifos on
complex neurobehavioural and neurobehavioural development.
CONCLUSIONS 23

2 CONCLUSIONS
The conclusions of the 2019 toxicology update are as follows:

 there is currently little regulatory quality experimental in vivo evidence that exposure to chlorpyrifos at doses
below those that inhibit blood cholinesterases results in human health relevant adverse effects. There is
published, non-regulatory quality, modelling and epidemiological data that is suggestive of this

 however, it is critical to note that the hypothesis that low doses of chlorpyrifos (ie doses below the thresholds
for inhibition of blood cholinesterases or ≤ 0.1 mg/kg bw/day in rats) cause adverse developmental effects
remains inadequately studied

 there is now a very substantial body of experimental evidence that exposure to chlorpyrifos at levels that result
in detectable inhibition of blood cholinesterases is a serious neurodevelopmental and neurobehavioural
developmental health hazard for humans

 APVMA proposes to establish a new acceptable daily intake of 0.001 mg/kg bw/day

 APVMA proposes to establish a new acute reference dose of 0.03 mg/kg bw.
24 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

3 REFERENCES
1. Adinolfi M, Haddad SA, Levels of plasma proteins in human and rat fetal CSF and the development of the
blood-CSF barrier, Neuropacceptable daily intakeatrie, 1977 Nov; 8(4):345–53, PubMed PMID: 74050.

2. APVMA, Reconsideration of chlorpyrifos: supplementary toxicology assessment report, 2018 July 5; available
at apvma.gov.au/node/26831.

3. Bayer SA, Altman J, Russo RJ, Zhang X, Timetables of neurogenesis in the human brain based on
experimentally determined patterns in the rat, Neurotoxicology, 1993 Spring; 14(1):83–144, Review. PubMed
PMID: 8361683.

4. Biedermann SV, Biedermann DG, Wenzlaff F, Kurjak T, Nouri S, Auer MK, Wiedemann K, Briken P, Haaker J,
Lonsdorf TB, Fuss J, An elevated plus-maze invmixed reality for studying human anxiety-related behaviour,
BMC Biol, 2017 Dec 21; 15(1):125. doi: 10.1186/s12915-017-0463-6. PubMed PMID: 29268740;
PubMedvCentral PMCID: PMC5740602.

5. Bonati M, Latini R, Marra G, Assael BM, Parini R, Theophylline distribution in the premature neonate, Dev
Pharmacol Ther, 1981; 3(2):65–73. PubMed PMID: 7318639.

6. Braquenier JB, Quertemont E, Tirelli E, Plumier JC, Anxiety in adult female mice following perinatal exposure
to chlorpyrifos, Neurotoxicol Teratol, 2010 Mar–Apr; 32(2):234–9. doi: 10.1016/j.ntt.2009.08.008, Epub 2009
Aug 28. PubMed PMID: 19716890.

7. Breslin WJ, Liberacki AB, Dittenber DA, Brzak KA, Quast JF, Chlorpyrifos: Two-generation dietary reproduction
study in Sprague-Dawley rats, 1991 Report No. K-044793-088 Toxicology Research Laboratory, Dow
Chemical Co., Midland, Michigan, USA (Dow AgroSciences).

8. Carr RL, Armstrong NH, Buchanan AT, Eells JB, Mohammed AN, Ross MK, Nail CA, Decreased anxiety in
juvenile rats following exposure to low levels of chlorpyrifos during development, Neurotoxicology, 2017 Mar;
59:183–90. doi:10.1016/j.neuro.2015.11.016, Epub 2015 Nov 28, PubMed PMID: 26642910; PubMed Central
PMCID: PMC5580823.

9. Chen WQ, Yuan L, Xue R, Li YF, Su RB, Zhang YZ, Li J, Repeated exposure to chlorpyrifos alters the
performance of adolescent male rats in animal models of depression and anxiety, Neurotoxicology, 2011 Aug;
32(4):355–61, doi: 10.1016/j.neuro.2011.03.008, Epub 2011 Mar 29. PubMed PMID: 21453723.

10. Clancy B, Kersh B, Hyde J, Darlington RB, Anand KJ, Finlay BL, Web-based method for translating
neurodevelopment from laboratory species to humans, Neuroinformatics, 2007 Spring; 5(1):79–94, PubMed
PMID: 17426354.

11. Coulston F, Goldberg L, Griffin T, Safety evaluation of DOWCO 179 in human volunteers, 1972 Institute of
Experimental Pathology and Toxicology, Albany Medical College, Albany, New York, USA (Dow
AgroSciences).

12. Dow [a], Comparison of cholinesterase (che) inhibition in young adult and preweanling CD rats after acute and
repeated chlorpyrifos or chlorpyrifos-oxon exposures, 2010. The Dow Chemical Company Study ID 091107,
pp 1–1062. 2010.

13. Dow [b], Benchmark dose modeling for cholinesterase inhibition of chlorpyrifos and clilorpyrifos-oxon, 2010,
Dow AgroSciences LLC Report ID: 0900956.000 B0T0 06010, pp 1–138.
REFERENCES 25

14. Ennaceur A, Tests of unconditioned anxiety—pitfalls and disappointments, Physiol Behav, 2014 Aug; 135:55–
71, doi: 10.1016/j.physbeh.2014.05.032, Epub 2014 Jun 5, Review, PubMed PMID: 24910138.

15. Gallo MA, Lawryk NJ, Organic phosphorous pesticides, in Handook of Pesticide Toxicology, Vol 2. Classes of
Pesticides (W.J. Hayes and E.R. Laws, Jr., Eds.), 1991 pp. 917–1123, Academic Press, New York.

16. Griffin P, Mason H, Heywood K, Cocker J, Oral and dermal absorption of chlorpyrifos: a human volunteer
study, Occup Environ Med, 1999 Jan; 56(1):10–3, PubMed PMID: 10341740; PubMed Central PMCID:
PMC1757654.

17. Kisicki JC, Seip CW, Combs ML, A rising dose toxicology study to determine the no-observable-effect-levels
(NOEL) for erythrocyte acetylcholinesterase (AChE) inhibition and cholinergic signs and symptoms of
chlorpyrifos at three dose levels, 1999 Dow Agrosciences, Report No. DR#K-044793-284.

18. López-Crespo GA, Carvajal F, Flores P, Sánchez-Santed F, Sánchez-Amate MC, Time course of biochemical
and behavioural effects of a single high dose of chlorpyrifos, Neurotoxicology, 2007 May; 28(3):541–7, Epub
2007 Feb 4, PubMed PMID: 17350100.

19. López-Crespo GA, Flores P, Sánchez-Santed F, Sánchez-Amate MC, Acute high dose of chlorpyrifos alters
performance of rats in the elevated plus-maze and the elevated T-maze, Neurotoxicology, 2009 Nov;
30(6):1025–9. doi: 10.1016/j.neuro.2009.07.009, Epub 2009 Jul 24, PubMed PMID: 19632271.

20. Marty MS, Andrus AK, Bell MP, Passage JK, Perala AW, Brzak KA, Bartels MJ, Beck MJ, Juberg DR,
Cholinesterase inhibition and toxicokinetics in immature and adult rats after acute or repeated exposures to
chlorpyrifos or chlorpyrifos-oxon, Regul Toxicol Pharmacol, 2012 Jul; 63(2):209–24, doi:
10.1016/j.yrtph.2012.03.015, Epub 2012 Apr 7, PubMed PMID: 22504667.

21. Nolan RJ, Rick DL, Freshour NL, Saunders JH, Chlorpyrifos: pharmacokinetics in human volunteers, Toxicol
Appl Pharmacol, 1984 Mar 30; 73(1):8–15, PubMed PMID: 6200956.

22. Nolan RJ, Proposed reference dose for acute and chronic exposure to chlorpyrifos based on the criteria
described by the Acute Cholinesterase Risk Assessment Task Force and the available animal and human
data, 1997, Report No. K-044793-110, Dow AgroSciences.

23. Poet TS, Timchalk C, Bartels MJ, Smith JN, McDougal R, Juberg DR, Price PS, Use of a probabilistic
PBPK/PD model to calculate Data Derived Extrapolation Factors for chlorpyrifos, Regul Toxicol Pharmacol,
2017 Jun; 86:59–73, doi: 10.1016/j.yrtph.2017.02.014, Epub 2017 Feb 24. PubMed PMID: 28238854.

24. Reiss R, Neal B, Lamb JC 4th, Juberg DR, Acetylcholinesterase inhibition dose-response modeling for
chlorpyrifos and chlorpyrifos-oxon, Regul Toxicol Pharmacol, 2012 Jun; 63(1):124–31, doi:
10.1016/j.yrtph.2012.03.008, Epub 2012 Mar 17. PubMed PMID: 22446730.

25. Sánchez-Amate MC, Flores P, Sánchez-Santed F, Effects of chlorpyrifos in the plus-maze model of anxiety,
Behav Pharmacol, 2001 Jul; 12(4):285–92, PubMed PMID: 11548114.

26. Savy CY, Fitchett AE, McQuade R, Gartside SE, Morris CM, Blain PG, Judge SJ, Low-level repeated exposure
to diazinon and chlorpyrifos decrease anxiety-like behaviour in adult male rats as assessed by marble
burying behaviour, Neurotoxicology, 2015 Sep; 50:149–56, doi: 10.1016/j.neuro.2015.08.010, Epub 2015 Aug
19, PubMed PMID: 26297601.

27. Szabo JR, Young JT, Granjean M, Chlorpyrifos: 13-week dietary toxicity study in Fisher 344 rats, 1988 Report
No. TexasT: K-044793-071 Lake Jackson Research Center, Health and Environmental Sciences, Freeport,
Texas, USA (Dow AgroSciences).
26 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

28. Vidair CA, Age dependence of organophosphate and carbamate neurotoxicity in the postnatal rat:
extrapolation to the human, Toxicol Appl Pharmacol, 2004 Apr 15; 196(2):287–302, Review, PubMed PMID:
15081274.

29. Wagner D, Chlorpyrifos, Joint meeting of the FAO Panel of Experts on Pesticide Residues in Food and the
Environment and the WHO Core Assessment Group, Rome, 20–29 September 1999, available at
inchem.org/documents/jmpr/jmpmono/v99pr03.htm.

30. Young JT, Grandjean M, Chlorpyrifos: 2 year dietary chronic toxicity-oncogenicity study in Fischer 344 rats,
1988 Report No. K-044793-079 Lake Jackson Research Center, Dow Chemical Co., Freeport, Texas, USA
(Dow AgroSciences).
APPENDIX 1 27

APPENDIX 1: LITERATURE SEARCH RESULTS


1. Abbas N, Ali Shad S, Ismail M, Resistance to Conventional and New Insecticides in House Flies (Diptera:
Muscidae) From Poultry Facilities in Punjab, Pakistan, J Econ Entomol, 2015 Apr; 108(2):826–33, doi:
10.1093/jee/tou057, Epub 2015 Feb 2, PubMed PMID: 26470195.

2. Abbas N, Khan H, Shad SA, Cross-resistance, stability, and fitness cost of resistance to imidacloprid in Musca
domestica L., (Diptera: Muscidae), Parasitol Res, 2015 Jan; 114(1):247–55, doi: 10.1007/s00436-014-4186-0,
Epub 2014 Oct 24, PubMed PMID: 25342464.

3. Abolaji AO, Awogbindin IO, Adedara IA, Farombi EO, Insecticide chlorpyrifos and fungicide carbendazim,
common food contaminants mixture, induce hepatic, renal, and splenic oxidative damage in female rats, Hum
Exp Toxicol, 2017 May; 36(5):483–93, doi: 10.1177/0960327116652459, Epub 2016 Jun 6, PubMed PMID:
27268782.

4. Abolaji AO, Ojo M, Afolabi TT, Arowoogun MD, Nwawolor D, Farombi EO, Protective properties of 6-gingerol-
rich fraction from Zingiber officinale (Ginger) on chlorpyrifos-induced oxidative damage and inflammation in the
brain, ovary and uterus of rats, Chem Biol Interact, 2017 May 25; 270:15–23, doi:10.1016/j.cbi.2017.03.017,
Epub 2017 Mar 31, PubMed PMID: 28373059.

5. Aceves-Diez AE, Estrada-Castañeda KJ, Castañeda-Sandoval LM, Use of Bacillus thuringiensis supernatant
from a fermentation process to improve bioremediation of chlorpyrifos in contaminated soils, J Environ
Manage, 2015 Jul 1; 157:213–9, doi: 10.1016/j.jenvman.2015.04.026, Epub 2015 Apr 21, PubMed PMID:
25910975.

6. Adedara IA, Owoeye O, Ajayi BO, Awogbindin IO, Rocha JBT, Farombi EO, Diphenyl diselenide abrogates
chlorpyrifos-induced hypothalamic-pituitary-testicular axis impairment in rats, Biochem Biophys Res Commun,
2018 Sep 3; 503(1):171–6, doi: 10.1016/j.bbrc.2018.05.205, Epub 2018 Jun 5, PubMed PMID: 29859936.

7. Adedara IA, Owoeye O, Awogbindin IO, Ajayi BO, Rocha JBT, Farombi EO, Diphenyl diselenide abrogates
brain oxidative injury and neurobehavioural deficits associated with pesticide chlorpyrifos exposure in rats,
Chem Biol Interact, 2018 Sep 27; 296:105–16, doi: 10.1016/j.cbi.2018.09.016, PubMed PMID: 30267645.

8. Adedara IA, Rosemberg DB, de Souza D, Farombi EO, Aschner M, Souza DO, Rocha JBT, Neurobehavioural
and biochemical changes in Nauphoeta cinerea following dietary exposure to chlorpyrifos, Pestic Biochem
Physiol, 2016 Jun; 130:22–30, doi: 10.1016/j.pestbp.2015.12.004, Epub 2015 Dec 11, PubMed PMID:
27155480.

9. Afzal MB, Ijaz M, Farooq Z, Shad SA, Abbas N, Genetics and preliminary mechanism of chlorpyrifos
resistance in Phenacoccus solenopsis Tinsley (Homoptera: Pseudococcidae), Pestic Biochem Physiol, 2015
Mar; 119:42–7, doi: 10.1016/j.pestbp.2015.02.008, Epub 2015 Feb 27, PubMed PMID: 25868815.

10. Agatz A, Brown CD, Introducing the 2-DROPS model for two-dimensional simulation of crop roots and
pesticide within the soil-root zone, Sci Total Environ, 2017 May 15; 586:966–75, doi:
10.1016/j.scitotenv.2017.02.076, Epub 2017 Feb 16, PubMed PMID: 28215806.

11. Agatz A, Schumann MM, French BW, Brown CD, Vidal S, Assessment of acute toxicity tests and rhizotron
experiments to characterize lethal and sublethal control of soil-based pests, Pest Manag Sci, 2018 Nov;
74(11):2450–59, doi:10.1002/ps.4922, Epub 2018 May 17, PubMed PMID: 29575759.

12. Ahammed GJ, He BB, Qian XJ, Zhou YH, Shi K, Zhou J, Yu JQ, Xia XJ, 24-Epibrassinolide alleviates organic
pollutants-retarded root elongation by promoting redox homeostasis and secondary metabolism in Cucumis
28 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

sativus, L. Environ Pollut, 2017 Oct; 229:922–31, doi: 10.1016/j.envpol.2017.07.076, Epub 2017 Jul 31,
PubMed PMID: 28774551.

13. E, Khosroushahi AY, Eghbal MA, Eftekhari A, Betanin reduces organophosphate induced cytotoxicity in
primary hepatocyte via an anti-oxidative and mitochondrial dependent pathway, Pestic Biochem Physiol, 2018
Jan; 144:71–8, doi: 10.1016/j.pestbp.2017.11.009, Epub 2017 Dec 2, PubMed PMID: 29463411.

14. Akande MG, Aliu YO, Ambali SF, Ayo JO, Co-treatment of chlorpyrifos and lead induce serum lipid disorders in
rats: Alleviation by taurine, Toxicol Ind Health, 2016 Jul; 32(7):1328–34, doi: 10.1177/0748233714560394,
Epub 2014 Dec 23, PubMed PMID: 25537622.

15. Akande MG, Shittu M, Uchendu C, Yaqub LS, Taurine ameliorated thyroid function in rats co-administered with
chlorpyrifos and lead, Vet Res Commun, 2016 Dec; 40(3–4):123–9, Epub 2016 Aug 25, PubMed PMID:
27562511.

16. Akbar S, Sultan S, Soil bacteria showing a potential of chlorpyrifos degradation and plant growth
enhancement, Braz J Microbiol, 2016 Jul-Sep; 47(3):563–70, doi: 10.1016/j.bjm.2016.04.009, Epub 2016 Apr
20, PubMed PMID: 27266625; PubMed Central PMCID: PMC4927687.

17. Al Naggar Y, Codling G, Vogt A, Naiem E, Mona M, Seif A, Giesy JP, Organophosphorus insecticides in honey,
pollen and bees (Apis mellifera L.) and their potential hazard to bee colonies in Egypt, Ecotoxicol Environ Saf,
2015 Apr; 114:1–8, doi: 10.1016/j.ecoenv.2014.12.039, Epub 2015 Jan 6, PubMed PMID: 25574845.

18. Al Naggar Y, Wiseman S, Sun J, Cutler GC, Aboul-Soud M, Naiem E, Mona M, Seif A, Giesy JP, Effects of
environmentally-relevant mixtures of four common organophosphorus insecticides on the honey bee (Apis
mellifera L.), J Insect Physiol, 2015 Nov; 82:85–91, doi: 10.1016/j.jinsphys.2015.09.004, Epub 2015 Sep 21,
PubMed PMID: 26403075.

19. Alaa-Eldin EA, El-Shafei DA, Abouhashem NS, Individual and combined effect of chlorpyrifos and
cypermethrin on reproductive system of adult male albino rats, Environ Sci Pollut Res Int, 2017 Jan;
24(2):1532–43, doi:10.1007/s11356-016-7912-6, Epub 2016 Oct 26, PubMed PMID: 27785720.

20. Alharbi HA, Alcorn J, Al-Mousa A, Giesy JP, Wiseman SB, Toxicokinetics and toxicodynamics of chlorpyrifos is
altered in embryos of Japanese medaka exposed to oil sands process-affected water: evidence for inhibition of
P-glycoprotein, J Appl Toxicol, 2017 May; 37(5):591–601, doi: 10.1002/jat.3397, Epub 2016 Oct 24, PubMed
PMID: 27774651.

21. Alharbi HA, Letcher RJ, Mineau P, Chen D, Chu S, Organophosphate pesticide method development and
presence of chlorpyrifos in the feet of nearctic-neotropical migratory songbirds from Canada that over-winter in
Central America agricultural areas, Chemosphere, 2016 Feb; 144:827–35, doi:
10.1016/j.chemosphere.2015.09.052, Epub 2015 Sep 28, PubMed PMID: 26421621.

22. Alloisio S, Nobile M, Novellino A, Multiparametric characterisation of neuronal network activity for in vitro
agrochemical neurotoxicity assessment, Neurotoxicology, 2015 May; 48:152–65, doi:
10.1016/j.neuro.2015.03.013, Epub 2015 Apr 4, PubMed PMID: 25845298.

23. Alves TM, Marston ZP, MacRae IV, Koch RL, Effects of Foliar Insecticides on Leaf-Level Spectral Reflectance
of Soybean, J Econ Entomol, 2017 Dec 5; 110(6):2436–42, doi: 10.1093/jee/tox250, PubMed PMID:
29029168.
APPENDIX 1 29

24. Amani N, Soodi M, Daraei B, Dashti A, Chlorpyrifos Toxicity in Mouse Cultured Cerebellar Granule Neurons at
Different Stages of Development: Additive Effect on Glutamate-Induced Excitotoxicity, Cell J, 2016 Fall;
18(3):464–72, Epub 2016 Aug 24, PubMed PMID: 27602329; PubMed Central PMCID: PMC5011335.

25. Anderson T, Liu J, McMurry S, Pope C, Comparative in vitro and in vivo effects of chlorpyrifos oxon in the
outbred CD-1 mouse (Mus musculus) and great plains toad (Anaxyrus cognatus), Environ Toxicol Chem, 2018
Jul; 37(7):1898–1906, doi: 10.1002/etc.4139, Epub 2018 May 7, PubMed PMID: 29573455.

26. Anh HQ, Tomioka K, Tue NM, Tuyen LH, Chi NK, Minh TB, Viet PH, Takahashi S, A preliminary investigation of
942 organic micro-pollutants in the atmosphere in waste processing and urban areas, northern Vietnam:
Levels, potential sources, and risk assessment, Ecotoxicol Environ Saf, 2018 Oct 22; 167:354–64, doi:
10.1016/j.ecoenv.2018.10.026, PubMed PMID: 30359902.

27. Appenzeller BMR, Hardy EM, Grova N, Chata C, Faÿs F, Briand O, Schroeder H, Duca RC, Hair analysis for
the biomonitoring of pesticide exposure: comparison with blood and urine in a rat model, Arch Toxicol, 2017
Aug; 91(8):2813–25, doi: 10.1007/s00204-016-1910-9, Epub 2016 Dec 23, PubMed PMID: 28011991;
PubMed Central PMCID: PMC5515982.

28. Arellano-Aguilar O, Betancourt-Lozano M, Aguilar-Zárate G, Ponce de Leon-Hill C, Agrochemical loacceptable


daily intakeng in drains and rivers and its connection with pollution in coastal lagoons of the Mexican Pacific,
Environ Monit Assess, 2017 Jun; 189(6):270, doi: 10.1007/s10661-017-5981-8, Epub 2017 May 16, PubMed
PMID: 28510105.

29. Arnold SM, Morriss A, Velovitch J, Juberg D, Burns CJ, Bartels M, Aggarwal M, Poet T, Hays S, Price P,
Derivation of human Biomonitoring Guidance Values for chlorpyrifos using a physiologically based
pharmacokinetic and pharmacodynamic model of cholinesterase inhibition, Regul Toxicol Pharmacol, 2015
Mar; 71(2):235–43, doi: 10.1016/j.yrtph.2014.12.013, Epub 2014 Dec 25, PubMed PMID: 25543108.

30. Aroonvilairat S, Tangjarukij C, Sornprachum T, Chaisuriya P, Siwadune T, Ratanabanangkoon K, Effects of


topical exposure to a mixture of chlorpyrifos, cypermethrin and captan on the hematological and
immunological systems in male Wistar rats, Environ Toxicol Pharmacol, 2018 Apr; 59:53–60, doi:
10.1016/j.etap.2018.02.010, Epub 2018 Mar 2, PubMed PMID: 29529450.

31. Arrizubieta M, Simón O, Torres-Vila LM, Figueiredo E, Mendiola J, Mexia A, Caballero P, Williams T,
Insecticidal efficacy and persistence of a co-occluded binary mixture of Helicoverpa armigera
nucleopolyhedrovirus (HearNPV) variants in protected and field-grown tomato crops on the Iberian Peninsula,
Pest Manag Sci, 2016 Apr; 72(4):660–70, doi: 10.1002/ps.4035, Epub 2015 Jun 1, PubMed PMID: 25960129.

32. Askari-Saryazdi G, Hejazi MJ, Ferguson JS, Rashidi MR, Selection for chlorpyrifos resistance in Liriomyza
sativae Blanchard: Cross-resistance patterns, stability and biochemical mechanisms, Pestic Biochem Physiol,
2015 Oct; 124:86–92, doi: 10.1016/j.pestbp.2015.05.002, Epub 2015 May 28, PubMed PMID: 26453235.

33. Asselborn V, Fernández C, Zalocar Y, Parodi ER, Effects of chlorpyrifos on the growth and ultrastructure of
green algae, Ankistrodesmus gracilis, Ecotoxicol Environ Saf, 2015 Oct; 120:334–41, doi:
10.1016/j.ecoenv.2015.06.015, Epub 2015 Jun 19, PubMed PMID: 26099464.

34. Augusiak J, Van den Brink PJ, The influence of insecticide exposure and environmental stimuli on the
movement behaviour and dispersal of a freshwater isopod, Ecotoxicology, 2016 Sep; 25(7):1338–52, doi:
10.1007/s10646-016-1686-y, Epub 2016 Jun 15, PubMed PMID: 27307165; PubMed Central PMCID:
PMC4961728.
30 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

35. Avci B, Bilge SS, Arslan G, Alici O, Darakci O, Baratzada T, Ciftcioglu E, Yardan T, Bozkurt A, Protective
effects of dietary omega-3 fatty acid supplementation on organophosphate poisoning, Toxicol Ind Health, 2018
Feb; 34(2):69–82, doi: 10.1177/0748233717737646, Epub 2017 Nov 15, PubMed PMID: 29141517.

36. Baba N, Raina R, Verma P, Sultana M, Free racceptable daily intakecal-induced nephrotoxicity following
repeated oral exposureto chlorpyrifos alone and in conjunction with fluoride in rats, Turk J Med Sci, 2016 Feb
17; 46(2):512–7, doi: 10.3906/sag-1403-109, PubMed PMID: 27511519.

37. Babazadeh M, Najafi G, Effect of chlorpyrifos on sperm characteristics and testicular tissue changes in adult
male rats, Vet Res Forum, 2017; 8(4):319–26, Epub 2017 Dec 15, PubMed PMID: 29326791; PubMed Central
PMCID: PMC5756252.

38. Báez ME, Espinoza J, Silva R, Fuentes E, Sorption-desorption behaviour of pesticides and their degradation
products in volcanic and nonvolcanic soils: interpretation of interactions through two-way principal component
analysis, Environ Sci Pollut Res Int, 2015 Jun; 22(11):8576–85, doi: 10.1007/s11356-014-4036-8, Epub 2015
Jan 7, PubMed PMID: 25561264.

39. Balsebre A, Báez ME, Martínez J, Fuentes E, Matrix solid-phase dispersion associated to gas chromatography
for the assessment in honey bee of a group of pesticides of concern in the apicultural field, J Chromatogr A,
2018 Sep 14; 1567:47–54, doi: 10.1016/j.chroma.2018.06.062, Epub 2018 Jun 26, PubMed PMID: 29960737.

40. Barberis CL, Carranza CS, Magnoli K, Benito N, Magnoli CE, Development and removal ability of non-
toxigenic Aspergillus section Flavi in presence of atrazine, chlorpyrifos and endosulfan, Rev Argent Microbiol,
2018 Jun 6. pii: S0325-7541(18)30032-4, doi: 10.1016/j.ram.2018.03.002, PubMed PMID: 29885942.

41. Barski D, Spodniewska A, Effect of chlorpyrifos and enrofloxacin on selected enzymes in rats, Pol J Vet Sci,
2018 Mar; 21(1):39–46, doi: 10.24425/119020, PubMed PMID: 29624020.

42. Basaure P, Guardia-Escote L, Cabré M, Peris-Sampedro F, Sánchez-Santed F, Domingo JL, Colomina MT,
Postnatal chlorpyrifos exposure and apolipoprotein E (APOE) genotype differentially affect cholinergic
expression and developmental parameters in transgenic mice, Food Chem Toxicol, 2018 Aug; 118:42–52, doi:
10.1016/j.fct.2018.04.065, Epub 2018 May 3, PubMed PMID: 29729306.

43. Basaure P, Peris-Sampedro F, Cabré M, Reverte I, Colomina MT, Two cholinesterase inhibitors trigger
dissimilar effects on behaviour and body weight in C57BL/6 mice: The case of chlorpyrifos and rivastigmine,
Behav Brain Res, 2017 Feb 1; 318:1–11, doi: 10.1016/j.bbr.2016.10.014, Epub 2016 Oct 11, PubMed PMID:
27732893.

44. Becker JM, Liess M, Biotic interactions govern genetic adaptation to toxicants, Proc Biol Sci, 2015 May 7;
282(1806):20150071, doi: 10.1098/rspb.2015.0071, PubMed PMID: 25833856; PubMed Central PMCID:
PMC4426622.

45. Bedia C, Dalmau N, Jaumot J, Tauler R, Phenotypic malignant changes and untargeted lipidomic analysis of
long-term exposed prostate cancer cells to endocrine disruptors, Environ Res, 2015 Jul; 140:18–31,
doi:10.1016/j.envres.2015.03.014, Epub 2015 Mar 29, PubMed PMID: 25817993.

46. Bedia C, Tauler R, Jaumot J, Analysis of multiple mass spectrometry images from different Phaseolus vulgaris
samples by multivariate curve resolution, Talanta, 2017 Dec 1; 175:557–65, doi:
10.1016/j.talanta.2017.07.087, Epub 2017 Jul 29. PubMed PMID: 28842033.

47. Beloti VH, Alves GR, Araújo DF, Picoli MM, Moral Rde A, Demétrio CG, Yamamoto PT, Lethal and Sublethal
Effects of Insecticides Used on Citrus, on the Ectoparasitoid Tamarixia racceptable daily intakeata, PLoS One,
APPENDIX 1 31

2015 Jul 1; 10(7):e0132128, doi: 10.1371/journal.pone.0132128, eCollection 2015, PubMed PMID: 26132327;
PubMed Central PMCID: PMC4488444.

48. Benedetto A, Brizio P, Squadrone S, Scanzio T, Righetti M, Gasco L, Prearo M, Abete MC, Oxidative stress
related to chlorpyrifos exposure in rainbow trout: Acute and medium term effects on genetic biomarkers, Pestic
Biochem Physiol, 2016 May; 129:63–9, doi: 10.1016/j.pestbp.2015.10.019, Epub 2015 Oct 23, PubMed PMID:
27017883.

49. Bertucci F, Jacob H, Mignucci A, Gache C, Roux N, Besson M, Berthe C, Metian M, Lecchini D, Decreased
retention of olfactory predator recognition in juvenile surgeon fish exposed to pesticide, Chemosphere, 2018
Oct; 208:469–75, doi: 10.1016/j.chemosphere.2018.06.017, Epub 2018 Jun 4, PubMed PMID: 29886335.

50. Bilal M, Freed S, Ashraf MZ, Zaka SM, Khan MB, Activity of acetylcholinesterase and acid and alkaline
phosphatases in different insecticide-treated Helicoverpa armigera (Hübner), Environ Sci Pollut Res Int, 2018
Aug; 25(23):22903–10, doi: 10.1007/s11356-018-2394-3, Epub 2018 Jun 1, PubMed PMID: 29858991.

51. Bonansea RI, Wunderlin DA, Amé MV, Behavioural swimming effects and acetylcholinesterase activity
changes in Jenynsia multidentata exposed to chlorpyrifos and cypermethrin individually and in mixtures,
Ecotoxicol Environ Saf, 2016 Jul; 129:311–9, doi: 10.1016/j.ecoenv.2016.03.043, Epub 2016 Apr 7, PubMed
PMID: 27060258.

52. Bonifacio AF, Ballesteros ML, Bonansea RI, Filippi I, Amé MV, Hued AC, Environmental relevant
concentrations of a chlorpyrifos commercial formulation affect two neotropical fish species, Cheirodon
interruptus and Cnesterodon decemmaculatus, Chemosphere, 2017 Dec; 188:486–93, doi:
10.1016/j.chemosphere.2017.08.156, Epub 2017 Aug 30, PubMed PMID: 28903091.

53. Bosch D, Rodríguez MA, Avilla J, Monitoring resistance of Cydia pomonella (L.) Spanish field populations to
new chemical insecticides and the mechanisms involved, Pest Manag Sci, 2018 Apr; 74(4):933–43, doi:
10.1002/ps.4791, Epub 2018 Jan 5, PubMed PMID: 29148167.

54. Bradley PM, Journey CA, Romanok KM, Barber LB, Buxton HT, Foreman WT, Furlong ET, Glassmeyer ST,
Hlacceptable daily intakek ML, Iwanowicz LR, Jones DK, Kolpin DW, Kuivila KM, Loftin KA, Mills MA, Meyer
MT, Orlando JL, Reilly TJ, Smalling KL, Villeneuve DL, Expanded Target-Chemical Analysis Reveals Extensive
Mixed-Organic-Contaminant Exposure in U.S. Streams, Environ Sci Technol, 2017 May 2; 51(9):4792–802,
doi: 10.1021/acs.est.7b00012, Epub 2017 Apr 12, PubMed PMID: 28401767; PubMed Central PMCID:
PMC5695041.

55. Brandt C, Burnett DC, Arcinas L, Palace V, Gary Anderson W, Effects of chlorpyrifos on in vitro sex steroid
production and thyroid follicular development in adult and larval Lake Sturgeon, Acipenser fulvescens,
Chemosphere, 2015 Aug; 132:179–87, doi: 10.1016/j.chemosphere.2015.03.031, Epub 2015 Apr 5, PubMed
PMID: 25855011.

56. Buck JC, Hua J, Brogan WR 3rd, Dang TD, Urbina J, Bendis RJ, Stoler AB, Blaustein AR, Relyea RA, Effects
of Pesticide Mixtures on Host-Pathogen Dynamics of the Amphibian Chytrid Fungus, PLoS One, 2015 Jul 16;
10(7):e0132832, doi: 10.1371/journal.pone.0132832, eCollection 2015, PubMed PMID: 26181492; PubMed
Central PMCID: PMC4504700.

57. Bugel SM, Tanguay RL, Multidimensional chemobehaviour analysis of flavonoids and neuroactive compounds
in zebrafish, Toxicol Appl Pharmacol, 2018 Apr 1; 344:23–34, doi: 10.1016/j.taap.2018.02.019, Epub 2018 Feb
27, PubMed PMID: 29499247; PubMed Central PMCID: PMC5912343.
32 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

58. Buntyn RW, Alugubelly N, Hybart RL, Mohammed AN, Nail CA, Parker GC, Ross MK, Carr RL, Inhibition of
Endocannabinoid-Metabolizing Enzymes in Peripheral Tissues Following Developmental Chlorpyrifos
Exposure in Rats, Int J Toxicol, 2017 Sep/Oct; 36(5):395–402, doi: 10.1177/1091581817725272, Epub 2017
Aug 18, PubMed PMID: 28820005; PubMed Central PMCID: PMC5585048.

59. Burke RD, Todd SW, Lumsden E, Mullins RJ, Mamczarz J, Fawcett WP, Gullapalli RP, Randall WR, Pereira
EFR, Albuquerque EX, Developmental neurotoxicity of the organophosphorus insecticide chlorpyrifos: from
clinical findings to preclinical models and potential mechanisms, J Neurochem, 2017 Aug; 142 Suppl 2:162–
77, doi: 10.1111/jnc.14077, Review, PubMed PMID: 28791702; PubMed Central PMCID: PMC5673499.

60. Buzzetti K, Chorbadjian RA, Nauen R, Resistance Management for San Jose Scale (Hemiptera: Diaspididae),
J Econ Entomol, 2015 Dec; 108(6):2743–52, doi: 10.1093/jee/tov236, Epub 2015 Jul 31, PubMed PMID:
26470382.

61. Calatayud-Vernich P, Calatayud F, Simó E, Suarez-Varela MM, Picó Y, Influence of pesticide use in fruit
orchards during blooming on honeybee mortality in 4 experimental apiaries, Sci Total Environ, 2016 Jan 15;
541:33–41, doi: 10.1016/j.scitotenv.2015.08.131, Epub 2015 Sep 19, PubMed PMID: 26398448.

62. Cao F, Souders CL 2nd, Li P, Pang S, Qiu L, Martyniuk CJ, Biological impacts of organophosphates
chlorpyrifos and diazinon on development, mitochondrial bioenergetics, and locomotor activity in zebrafish
(Danio rerio), Neurotoxicol Teratol, 2018 Oct 2; 70:18–27, doi: 10.1016/j.ntt.2018.10.001, PubMed PMID:
30290195.

63. Cao Y, Yang Q, Tu XH, Li SG, Liu S, Molecular characterization of a typical 2-Cys thioredoxin peroxidase from
the Asiatic rice borer Chilo suppressalis and its role in oxidative stress, Arch Insect Biochem Physiol, 2018
Sep; 99(1):e21476, doi: 10.1002/arch.21476, Epub 2018 Jun 5, PubMed PMID: 29873106.

64. Capoferri D, Álvarez-Diduk R, Del Carlo M, Compagnone D, Merkoçi A, Electrochromic Molecular Imprinting
Sensor for Visual and Smartphone-Based Detections, Anal Chem, 2018 May 1; 90(9):5850-5856, doi:
10.1021/acs.analchem.8b00389, Epub 2018 Apr 11, PubMed PMID: 29617110.

65. Carazo-Rojas E, Pérez-Rojas G, Pérez-Villanueva M, Chinchilla-Soto C, Chin-Pampillo JS, Aguilar-Mora P,


Alpízar-Marín M, Masís-Mora M, Rodríguez-Rodríguez CE, Vryzas Z, Pesticide monitoring and
ecotoxicological risk assessment in surface water bodies and sediments of a tropical agro-ecosystem, Environ
Pollut, 2018 Oct; 241:800–9, doi: 10.1016/j.envpol.2018.06.020, Epub 2018 Jun 14, PubMed PMID:
29909306.

66. Cardoza YJ, Drake WL, Jordan DL, Schroeder-Moreno MS, Arellano C, Brandenburg RL, Impact of Location,
Cropping History, Tillage, and Chlorpyrifos on Soil Arthropods in Peanut, Environ Entomol, 2015 Aug;
44(4):951–9, doi: 10.1093/ee/nvv074, Epub 2015 May 18, PubMed PMID: 26314040.

67. Carr RL, Armstrong NH, Buchanan AT, Eells JB, Mohammed AN, Ross MK, Nail CA, Decreased anxiety in
juvenile rats following exposure to low levels of chlorpyrifos during development, Neurotoxicology, 2017 Mar;
59:183–90, doi: 10.1016/j.neuro.2015.11.016, Epub 2015 Nov 28, PubMed PMID: 26642910; PubMed Central
PMCID: PMC5580823.

68. Carr RL, Dail MB, Chambers HW, Chambers JE, Species differences in paraoxonase mediated hydrolysis of
several organophosphorus insecticide metabolites, J Toxicol, 2015; 2015:470189, doi: 10.1155/2015/470189,
Epub 2015 Feb 15, PubMed PMID: 25784934; PubMed Central PMCID: PMC4345044.

69. Casida JE, Why Prodrugs and Propesticides Succeed, Chem Res Toxicol, 2017 May 15; 30(5):1117–26, doi:
10.1021/acs.chemrestox.7b00030, Epub 2017 Apr 7, Review. PubMed PMID: 28334528.
APPENDIX 1 33

70. Chaklader M, Law S, Alteration of hedgehog signaling by chronic exposure to different pesticide formulations
and unveiling the regenerative potential of recombinant sonic hedgehog in mouse model of bone marrow
aplasia, Mol Cell Biochem, 2015 Mar; 401(1–2):115–31, doi: 10.1007/s11010-014-2299-5, Epub 2014 Dec 4.
PubMed PMID: 25472879.

71. Chauhan LK, Varshney M, Pandey V, Sharma P, Verma VK, Kumar P, Goel SK, ROS-dependent genotoxicity,
cell cycle perturbations and apoptosis in mouse bone marrow cells exposed to formulated mixture of
cypermethrin and chlorpyrifos, Mutagenesis, 2016 Nov; 31(6):635–42, Epub 2016 Jul 28, PubMed PMID:
27470700.

72. Che W, Huang J, Guan F, Wu Y, Yang Y, Cross-resistance and Inheritance of Resistance to Emamectin
Benzoate in Spodoptera exigua (Lepidoptera: Noctuidae), J Econ Entomol, 2015 Aug; 108(4):2015–20, doi:
10.1093/jee/tov168, Epub 2015 Jul 9, PubMed PMID: 26470348.

73. Chebab S, Mekircha F, Leghouchi E, Potential protective effect of Pistacia lentiscus oil against chlorpyrifos-
induced hormonal changes and oxidative damage in ovaries and thyroid of female rats, Biomed
Pharmacother, 2017 Dec; 96:1310–6, doi: 10.1016/j.biopha.2017.11.081, Epub 2017 Dec 6, PubMed PMID:
29217167.

74. Chen R, Cui Y, Zhang X, Zhang Y, Chen M, Zhou T, Lan X, Dong W, Pan C, Chlorpyrifos Induced Testicular
Cell Apoptosis through Generation of Reactive Oxygen Species and Phosphorylation of AMPK, J Agric Food
Chem, 2018 Oct 31, doi: 10.1021/acs.jafc.8b03407, PubMed PMID: 30378422.

75. Chen S, Chen M, Wang Z, Qiu W, Wang J, Shen Y, Wang Y, Ge S, Toxicological effects of chlorpyrifos on
growth, enzyme activity and chlorophyll a synthesis of freshwater microalgae, Environ Toxicol Pharmacol,
2016 Jul; 45:179–86, doi: 10.1016/j.etap.2016.05.032, Epub 2016 Jun 1, PubMed PMID: 27314761.

76. Chen S, Lu M, Zhang N, Zou X, Mo M, Zheng S, Nuclear factor erythroid-derived 2-related factor 2 activates
glutathione S-transferase expression in the midgut of Spodoptera litura (Lepidoptera: Noctuidae) in response
to phytochemicals and insecticides, Insect Mol Biol, 2018 Aug; 27(4):522–32, doi: 10.1111/imb.12391, Epub
2018 May 10, PubMed PMID: 29749087.

77. Chen XP, Chao YS, Chen WZ, Dong JY, Mother gestational exposure to organophosphorus pesticide induces
neuron and glia loss in daughter adult brain, J Environ Sci Health B, 2017 Feb; 52(2):77–83, doi:
10.1080/03601234.2016.1239973, Epub 2016 Oct 24, PubMed PMID: 28099088.

78. Chen XP, Wang TT, Wu XZ, Wang DW, Chao YS, An in vivo study in mice: mother's gestational exposure to
organophosphorus pesticide retards the division and migration process of neural progenitors in the fetal
developing brain, Toxicol Res (Camb), 2016 Jun 14; 5(5):1359–70, doi: 10.1039/c5tx00282f, eCollection 2016
Sep 1, PubMed PMID: 30090440; PubMed Central PMCID: PMC6062264.

79. Chen XZ, Hu QX, Liu QQ, Wu G, Cloning of Wing-Development-Related Genes and mRNA Expression Under
Heat Stress in Chlorpyrifos-Resistant and -Susceptible Plutella xylostella, Sci Rep, 2018 Oct 15; 8(1):15279,
doi: 10.1038/s41598-018-33315-z, PubMed PMID: 30323169; PubMed Central PMCID: PMC6189056.

80. Chennappa G, Naik MK, Adkar-Purushothama CR, Amaresh YS, Sreenivasa MY, PGP potential, abiotic stress
tolerance and antifungal activity of Azotobacter strains isolated from paddy soils, Indian J Exp Biol, 2016 May;
54(5):322–31. PubMed PMID: 27319051.

81. Chin-Pampillo JS, Masís-Mora M, Ruiz-Hidalgo K, Carazo-Rojas E, Rodríguez-Rodríguez CE, Removal of


carbofuran is not affected by co-application of chlorpyrifos in a coconut fiber/compost based biomixture after
34 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

aging or pre-exposure, J Environ Sci (China), 2016 Aug; 46:182–9, doi: 10.1016/j.jes.2015.12.026, Epub 2016
Mar 8. PubMed PMID: 27521950.

82. Chowdhury R, Faria S, Huda MM, Chowdhury V, Maheswary NP, Mondal D, Akhter S, Akter S, Khan RK, Nabi
SG, Kroeger A, Argaw D, Alvar J, Dash AP, Banu Q, Control of Phlebotomus argentipes (Diptera:
Psychodidae) sand fly in Bangladesh: A cluster randomized controlled trial, PLoS Negl Trop Dis, 2017 Sep 5;
11(9):e0005890, doi: 10.1371/journal.pntd.0005890, eCollection 2017 Sep, PubMed PMID: 28873425;
PubMed Central PMCID: PMC5600390.

83. Christen V, Fent K, Exposure of honey bees (Apis mellifera) to different classes of insecticides exhibit distinct
molecular effect patterns at concentrations that mimic environmental contamination, Environ Pollut, 2017 Jul;
226:48–59, doi: 10.1016/j.envpol.2017.04.003, Epub 2017 Apr 9, PubMed PMID: 28402838.

84. Christen V, Rusconi M, Crettaz P, Fent K, Developmental neurotoxicity of different pesticides in PC-12 cells in
vitro, Toxicol Appl Pharmacol, 2017 Jun 15; 325:25–36, doi: 10.1016/j.taap.2017.03.027, Epub 2017 Apr 3,
PubMed PMID:28385489.

85. Clarke R, Connolly L, Frizzell C, Elliott CT, Challenging conventional risk assessment with respect to human
exposure to multiple food contaminants in food: A case study using maize, Toxicol Lett, 2015 Oct 1; 238(1):54–
64, doi: 10.1016/j.toxlet.2015.07.006, Epub 2015 Jul 18, PubMed PMID: 26196220.

86. Coban A, Carr RL, Chambers HW, Willeford KO, Chambers JE, Comparison of inhibition kinetics of several
organophosphates, including some nerve agent surrogates, using human erythrocyte and rat and mouse brain
acetylcholinesterase, Toxicol Lett, 2016 Apr 25; 248:39–45, doi: 10.1016/j.toxlet.2016.03.002, Epub 2016 Mar
7, PubMed PMID: 26965078; PubMed Central PMCID: PMC4805473.

87. Cogo LA, Santos Filha VA, Murashima Ade A, Hyppolito MA, Silveira AF, Morphological analysis of the
vestibular system of guinea pigs poisoned by organophosphate, Braz J Otorhinolaryngol, 2016 Jan-Feb;
82(1):11–6, doi: 10.1016/j.bjorl.2015.10.001, Epub 2015 Oct 24, PubMed PMID: 26589545.

88. Cooper BY, Flunker LD, Johnson RD, Nutter TJ, Behavioural, cellular and molecular maladaptations covary
with exposure to pyridostigmine bromide in a rat model of gulf war illness pain, Toxicol Appl Pharmacol, 2018
Aug 1; 352:119–31, doi: 10.1016/j.taap.2018.05.023, Epub 2018 May 24, PubMed PMID: 29803855.

89. Cooper BY, Johnson RD, Nutter TJ, Exposure to Gulf War Illness chemicals induces functional muscarinic
receptor maladaptations in muscle nociceptors, Neurotoxicology, 2016 May; 54:99–110, doi:
10.1016/j.neuro.2016.04.001, Epub 2016 Apr 4, PubMed PMID: 27058124.

90. Cowell WJ, Wright RJ, Sex-Specific Effects of Combined Exposure to Chemical and Non-chemical Stressors
on Neuroendocrine Development: a Review of Recent Findings and Putative Mechanisms, Curr Environ
Health Rep, 2017 Dec; 4(4):415–25, doi: 10.1007/s40572-017-0165-9, Review. PubMed PMID: 29027649.

91. Cunha AF, Felippe ISA, Ferreira-Junior NC, Resstel LBM, Guimarães DAM, Beijamini V, Paton JFR, Sampaio
KN, Neuroreflex control of cardiovascular function is impaired after acute poisoning with chlorpyrifos, an
organophosphorus insecticide: Possible short and long term clinical implications, Toxicology, 2018 Apr 1; 398–
9:13–22, doi: 10.1016/j.tox.2018.02.005, Epub 2018 Feb 19, PubMed PMID: 29471072.

92. D'Agostino J, Zhang H, Kenaan C, Hollenberg PF, Mechanism-Based Inactivation of Human Cytochrome P450
2B6 by Chlorpyrifos, Chem Res Toxicol, 2015 Jul 20; 28(7):1484–95, doi: 10.1021/acs.chemrestox.5b00156,
Epub 2015 Jun 30, PubMed PMID: 26075493.
APPENDIX 1 35

93. Dai H, Deng Y, Zhang J, Han H, Zhao M, Li Y, Zhang C, Tian J, Bing G, Zhao L, PINK1/Parkin-mediated
mitophagy alleviates chlorpyrifos-induced apoptosis in SH-SY5Y cells, Toxicology, 2015 Aug 6; 334:72–80,
doi: 10.1016/j.tox.2015.06.003, Epub 2015 Jun 9, PubMed PMID: 26070385.

94. Dai P, Jack CJ, Mortensen AN, Ellis JD, Acute toxicity of five pesticides to Apis mellifera larvae reared in vitro,
Pest Manag Sci, 2017 Nov; 73(11):2282–6, doi: 10.1002/ps.4608, Epub 2017 Jul 24, PubMed PMID:
28485079.

95. Daisley BA, Trinder M, McDowell TW, Collins SL, Sumarah MW, Reid G, Microbiota-Mediated Modulation of
Organophosphate Insecticide Toxicity by Species-Dependent Interactions with Lactobacilli in a Drosophila
melanogaster Insect Model, Appl Environ Microbiol, 2018 Apr 16; 84(9). pii: e02820-17, doi:
10.1128/AEM.02820-17, Print 2018 May 1, PubMed PMID: 29475860; PubMed Central PMCID: PMC5930343.

96. Darwiche W, Delanaud S, Dupont S, Ghamlouch H, Ramadan W, Joumaa W, Bach V,Gay-Quéheillard J,


Impact of prenatal and postnatal exposure to the pesticide chlorpyrifos on the contraction of rat ileal muscle
strips: involvement of an inducible nitric oxide synthase-dependent pathway, Neurogastroenterol Motil, 2017
Feb; 29(2), doi: 10.1111/nmo,12918, Epub 2016 Aug 21, PubMed PMID: 27545116.

97. Darwiche W, Gay-Quéheillard J, Delanaud S, El Khayat El Sabbouri H, Khachfe H, Joumaa W, Bach V,


Ramadan W, Impact of chronic exposure to the pesticide chlorpyrifos on respiratory parameters and sleep
apnea in juvenile and adult rats, PLoS One, 2018 Jan 22; 13(1):e0191237, doi:
10.1371/journal.pone.0191237, eCollection 2018, PubMed PMID: 29357379; PubMed Central PMCID:
PMC5777649.

98. Dawkar VV, Chikate YR, More TH, Gupta VS, Giri AP, The expression of proteins involved in digestion and
detoxification are regulated in Helicoverpa armigera to cope up with chlorpyrifos insecticide, Insect Sci, 2016
Feb; 23(1):68–77, doi: 10.1111/1744-7917.12177, Epub 2015 Jan 21, PubMed PMID: 25284010.

99. de Beeck LO, Verheyen J, Stoks R, Strong differences between two congeneric species in sensitivity to
pesticides in a warming world, Sci Total Environ, 2018 Mar 15; 618:60–69, doi:
10.1016/j.scitotenv.2017.10.311, Epub 2017 Nov 8, PubMed PMID: 29126027.

1. De Felice A, Greco A, Calamandrei G, Minghetti L, Prenatal exposure to the organophosphate insecticide


chlorpyrifos enhances brain oxidative stress and prostaglandin E2 synthesis in a mouse model of idiopathic
autism, J Neuroinflammation, 2016 Jun 14; 13(1):149, doi: 10.1186/s12974-016-0617-4, PubMed PMID:
27301868; PubMed Central PMCID: PMC4908699.

100. De Felice A, Scattoni ML, Ricceri L, Calamandrei G. Prenatal exposure to a common organophosphate
insecticide delays motor development in a mouse model of idiopathic autism, PLoS One, 2015 Mar 24;
10(3):e0121663, doi: 10.1371/journal.pone.0121663, eCollection 2015, PubMed PMID: 25803479; PubMed
Central PMCID: PMC4372449.

101. De Long NE, Holloway AC, Early-life chemical exposures and risk of metabolic syndrome, Diabetes Metab
Syndr Obes, 2017 Mar 21; 10:101–9, doi: 10.2147/DMSO.S95296, eCollection 2017, Review, PubMed PMID:
28367067; PubMed Central PMCID: PMC5370400.

102. Deacon S, Norman S, Nicolette J, Reub G, Greene G, Osborn R, Andrews P, Integrating ecosystem services
into risk management decisions: case study with Spanish citrus and the insecticide chlorpyrifos, Sci Total
Environ, 2015 Feb 1; 505:732–9, doi: 10.1016/j.scitotenv.2014.10.034, Epub 2014 Nov 5, PubMed PMID:
25461076.
36 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

103. Debecker S, Sanmartín-Villar I, de Guinea-Luengo M, Cordero-Rivera A, Stoks R, Integrating the pace-of-life


syndrome across species, sexes and individuals: covariation of life history and personality under pesticide
exposure, J Anim Ecol, 2016 May; 85(3):726–38, doi: 10.1111/1365-2656.12499, Epub 2016 Mar 1, PubMed
PMID: 26845756.

104. Del Pino J, Moyano P, Anadon MJ, García JM, Díaz MJ, García J, Frejo MT, Acute and long-term exposure to
chlorpyrifos induces cell death of basal forebrain cholinergic neurons through AChE variants alteration,
Toxicology, 2015 Oct 2; 336:1–9, doi: 10.1016/j.tox.2015.07.004, Epub 2015 Jul 22, PubMed PMID:
26210949.

105. Del Pino J, Moyano P, Anadon MJ, García JM, Díaz MJ, Gómez G, García J, Frejo MT, SN56 basal forebrain
cholinergic neuronal loss after acute and long-term chlorpyrifos exposure through oxidative stress generation;
P75(NTR) and α7-nAChRs alterations mediated partially by AChE variants disruption, Toxicology, 2016 Apr
15; 353–4:48–57, doi: 10.1016/j.tox.2016.05.007, Epub 2016 May 6, PubMed PMID: 27163631.

106. Del Prado-Lu JL, Insecticide residues in soil, water, and eggplant fruits and farmers' health effects due to
exposure to pesticides, Environ Health Prev Med, 2015 Jan; 20(1):53–62, doi: 10.1007/s12199-014-0425-3,
Epub 2014 Nov 21, PubMed PMID: 25413584; PubMed Central PMCID: PMC4284253.

107. Delpuech JM, Elicitation of superparasitization behaviour from the parasitoid wasp Leptopilina boulardi by the
organophosphorus insecticide chlorpyrifos, Sci Total Environ, 2017 Feb 15; 580:907–11, doi:
10.1016/j.scitotenv.2016.12.037, Epub 2016 Dec 29, PubMed PMID: 28040222.

108. Dempsey JL, Cui JY, Long Non-Coding RNAs: A Novel Paracceptable daily intakegm for Toxicology, Toxicol
Sci, 2017 Jan; 155(1):3–21, doi: 10.1093/toxsci/kfw203, Epub 2016 Nov 17, Review, PubMed PMID:
27864543; PubMed Central PMCID: PMC5216656.

109. Deng S, Chen Y, Wang D, Shi T, Wu X, Ma X, Li X, Hua R, Tang X, Li QX, Rapid biodegradation of
organophosphorus pesticides by Stenotrophomonas sp. G1, J Hazard Mater, 2015 Oct 30; 297:17–24, doi:
10.1016/j.jhazmat.2015.04.052, Epub 2015 Apr 20, PubMed PMID: 25938642.

110. Deng Y, Zhang Y, Lu Y, Zhao Y, Ren H, Hepatotoxicity and nephrotoxicity induced by the chlorpyrifos and
chlorpyrifos-methyl metabolite, 3,5,6-trichloro-2-pyridinol, in orally exposed mice, Sci Total Environ, 2016 Feb
15; 544:507–14, doi: 10.1016/j.scitotenv.2015.11.162, Epub 2015 Dec 8, PubMed PMID: 26674679.

111. Deng ZZ, Zhang F, Wu ZL, Yu ZY, Wu G, Chlorpyrifos-induced hormesis in insecticide-resistant and
-susceptible Plutella xylostella under normal and high temperatures, Bull Entomol Res, 2016 Jun; 106(3):378–
86, doi:10.1017/S000748531600002X, PubMed PMID: 27241230.

112. Deveci HA, Karapehlivan M. Chlorpyrifos-induced parkinsonian model in mice: Behaviour, histopathology and
biochemistry, Pestic Biochem Physiol, 2018 Jan; 144:36–41, doi: 10.1016/j.pestbp.2017.11.002, Epub 2017
Nov 10, PubMed PMID: 29463406.

113. Dewer Y, Pottier MA, Lalouette L, Maria A, Dacher M, Belzunces LP, Kairo G, Renault D, Maibeche M,
Siaussat D, Behavioural and metabolic effects of sublethal doses of two insecticides, chlorpyrifos and
methomyl, in the Egyptian cotton leafworm, Spodoptera littoralis (Boisduval) (Lepidoptera: Noctuidae), Environ
Sci Pollut Res Int, 2016 Feb; 23(4):3086–96, doi: 10.1007/s11356-015-5710-1, Epub 2015 Nov 14, PubMed
PMID: 26566611.

114. Diepens NJ, Beltman WHJ, Koelmans AA, Van den Brink PJ, Baveco JM, Dynamics and recovery of a
sediment-exposed Chironomus riparius population: A modelling approach, Environ Pollut, 2016 Jun; 213:741–
50, doi: 10.1016/j.envpol.2016.03.051, Epub 2016 Mar 29, PubMed PMID: 27031571.
APPENDIX 1 37

115. Diepens NJ, Dimitrov MR, Koelmans AA, Smidt H, Molecular Assessment of Bacterial Community Dynamics
and Functional End Points during Sediment Bioaccumulation Tests, Environ Sci Technol, 2015 Nov 17;
49(22):13586–95, doi: 10.1021/acs.est.5b02992, Epub 2015 Oct 28, PubMed PMID: 26466173.

116. Dimzon IKD, Morata AS, Müller J, Yanela RK, Lebertz S, Weil H, Perez TR, Müller J, Dayrit FM, Knepper TP,
Trace organic chemical pollutants from the lake waters of San Pablo City, Philippines by targeted and non-
targeted analysis, Sci Total Environ, 2018 Oct 15; 639:588–95, doi: 10.1016/j.scitotenv.2018.05.217, Epub
2018 May 26, PubMed PMID: 29800852.

117. Dingemans MM, Schütte MG, Wiersma DM, de Groot A, van Kleef RG, Wijnolts FM, Westerink RH, Chronic
14-day exposure to insecticides or methylmercury modulates neuronal activity in primary rat cortical cultures,
Neurotoxicology, 2016 Dec; 57:194–202, doi: 10.1016/j.neuro.2016.10.002, Epub 2016 Oct 5, PubMed PMID:
27720795.

118. Diop A, Diop YM, Thiaré DD, Cazier F, Sarr SO, Kasprowiak A, Landy D, Delattre F, Monitoring survey of the
use patterns and pesticide residues on vegetables in the Niayes zone, Senegal, Chemosphere, 2016 Feb;
144:1715–21, doi: 10.1016/j.chemosphere.2015.10.058, Epub 2015 Nov 11, PubMed PMID: 26519803.

119. Dobrinas S, Soceanu A, Popescu V, Coatu V, Polycyclic aromatic hydrocarbons and pesticides in milk powder,
J Dairy Res, 2016 May; 83(2):261–5, doi: 10.1017/S0022029916000169, PubMed PMID: 27210498.

120. Dokuyucu R, Bilgili A, Hanedan B, Dogan H, Dokuyucu A, Celik MM, Attenuating effects of caffeic acid
phenethyl ester with intralipid on hepatotoxicity of chlorpyrifos in the case of rats, Med Pr, 2016 Dec 22;
67(6):743–9, doi: 10.13075/mp.5893.00462, Epub 2016 Dec 12, PubMed PMID: 28005083.

121. Dollinger J, Schacht VJ, Gaus C, Grant S, Effect of surfactant application practices on the vertical transport
potential of hydrophobic pesticides in agrosystems, Chemosphere, 2018 Oct; 209:78–87, doi:
10.1016/j.chemosphere.2018.06.078, Epub 2018 Jun 14, PubMed PMID: 29913402.

122. Dominah GA, McMinimy RA, Kallon S, Kwakye GF, Acute exposure to chlorpyrifos caused NADPH oxidase
mediated oxidative stress and neurotoxicity in a striatal cell model of Huntington's disease, Neurotoxicology,
2017 May; 60:54–69, doi: 10.1016/j.neuro.2017.03.004, Epub 2017 Mar 12, PubMed PMID: 28300621.

123. Donald CE, Scott RP, Blaustein KL, Halbleib ML, Sarr M, Jepson PC, Anderson KA, Silicone wristbands detect
individuals' pesticide exposures in West Africa, R Soc Open Sci, 2016 Aug 17; 3(8):160433, eCollection 2016
Aug, PubMed PMID: 27853621; PubMed Central PMCID: PMC5108971.

124. Đorđević TM, Đurović-Pejčev RD, The potency of Saccharomyces cerevisiae and Lactobacillus plantarum to
dissipate organophosphorus pesticides in wheat during fermentation, J Food Sci Technol, 2016 Dec;
53(12):4205–15, doi:10.1007/s13197-016-2408-4, Epub 2016 Dec 8, PubMed PMID: 28115761; PubMed
Central PMCID: PMC5223255.

125. Dores EF, Spadotto CA, Weber OL, Dalla Villa R, Vecchiato AB, Pinto AA, Environmental Behaviour of
Chlorpyrifos and Endosulfan in a Tropical Soil in Central Brazil, J Agric Food Chem, 2016 May 25;
64(20):3942–8, doi: 10.1021/acs.jafc.5b04508, Epub 2015 Dec 15, PubMed PMID: 26635198.

126. Dou X, Zhang L, Liu C, Li Q, Luo J, Yang M, Fluorometric competitive immunoassay for chlorpyrifos using
rhodamine-modified gold nanoparticles as a label, Mikrochim Acta, 2017 Dec 8; 185(1):41, doi:
10.1007/s00604-017-2561-0, PubMed PMID: 29594500.

127. Ducharme NA, Reif DM, Gustafsson JA, Bondesson M, Comparison of toxicity values across zebrafish early
life stages and mammalian studies: Implications for chemical testing, Reprod Toxicol, 2015 Aug 1; 55:3–10,
38 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

doi: 10.1016/j.reprotox.2014.09.005, Epub 2014 Sep 28, PubMed PMID: 25261610; PubMed Central PMCID:
PMC5396953.

128. Dusza HM, Cenijn PH, Kamstra JH, Westerink RHS, Leonards PEG, Hamers T, Effects of environmental
pollutants on calcium release and uptake by rat cortical microsomes, Neurotoxicology, 2018 Jul 26, pii: S0161-
813X(18)30300-0, doi: 10.1016/j.neuro.2018.07.015, PubMed PMID: 30056177.

129. Ejaz M, Afzal MB, Shabbir G, Serrão JE, Shad SA, Muhammad W, Laboratory selection of chlorpyrifos
resistance in an Invasive Pest, Phenacoccus solenopsis (Homoptera: Pseudococcidae): Cross-resistance,
stability and fitness cost, Pestic Biochem Physiol, 2017 Apr; 137:8–14, doi: 10.1016/j.pestbp.2016.09.001.
Epub 2016 Sep 7, PubMed PMID: 28364807.

130. Elhalwagy ME, Darwish NS, Shokry DA, El-Aal AG, Abd-Alrahman SH, Nahas AA, Ziada RM, Garlic and alpha
lipoic supplementation enhance the immune system of albino rats and alleviate implications of pesticides
mixtures, Int J Clin Exp Med, 2015 May 15; 8(5):7689–700, eCollection 2015, PubMed PMID: 26221319;
PubMed Central PMCID: PMC4509264.

131. El-Nahhal Y, Lubbad R, Acute and single repeated dose effects of low concentrations of chlorpyrifos, diuron,
and their combination on chicken, Environ Sci Pollut Res Int, 2018 Apr; 25(11):10837–47, doi:
10.1007/s11356-018-1313-y, Epub 2018 Feb 3, PubMed PMID: 29397503.

132. 136. El-Sayed NM, Ahmed AAM, Selim MAA, Cytotoxic effect of chlorpyrifos is associated with activation of
Nrf-2/HO-1 system and inflammatory response in tongue of male Wistar rats, Environ Sci Pollut Res Int, 2018
Apr; 25(12):12072–82, doi: 10.1007/s11356-018-1391-x, Epub 2018 Feb 16, PubMed PMID: 29453720.

133. Espinoza M, Rivero Osimani V, Sánchez V, Rosenbaum E, Guiñazú N, B-esterase determination and
organophosphate insecticide inhibitory effects in JEG-3 trophoblasts, Toxicol In Vitro, 2016 Apr; 32:190–7, doi:
10.1016/j.tiv.2016.01.001, Epub 2016 Jan 12, PubMed PMID: 26790371.

134. Etterson M, Garber K, Odenkirchen E, Mechanistic modeling of insecticide risks to breeding birds in North
American agroecosystems, PLoS One, 2017 May 3; 12(5):e0176998, doi: 10.1371/journal.pone.0176998,
eCollection 2017, PubMed PMID: 28467479; PubMed Central PMCID: PMC5415183.

135. Ezzi L, Belhadj Salah I, Haouas Z, Sakly A, Grissa I, Chakroun S, Kerkeni E, Hassine M, Mehdi M, Ben Cheikh
H, Histopathological and genotoxic effects of chlorpyrifos in rats, Environ Sci Pollut Res Int, 2016 Mar;
23(5):4859–67, doi: 10.1007/s11356-015-5722-x, Epub 2015 Nov 6, PubMed PMID: 26545888.

136. Fang B, Li JW, Zhang M, Ren FZ, Pang GF, Chronic chlorpyrifos exposure elicits diet-specific effects on
metabolism and the gut microbiome in rats, Food Chem Toxicol, 2018 Jan; 111:144–52, doi:
10.1016/j.fct.2017.11.001, Epub 2017 Nov 3, PubMed PMID: 29109040.

137. Farajzadeh MA, Mohebbi A, Development of magnetic dispersive solid phase extraction using toner powder as
an efficient and economic sorbent in combination with dispersive liquid-liquid microextraction for extraction of
some widely used pesticides in fruit juices, J Chromatogr A, 2018 Jan 12; 1532:10–19, doi:
10.1016/j.chroma.2017.11.048, Epub 2017 Nov 22, PubMed PMID: 29174132.

138. Faria M, Garcia-Reyero N, Padrós F, Babin PJ, Sebastián D, Cachot J, Prats E, Arick Ii M, Rial E, Knoll-Gellida
A, Mathieu G, Le Bihanic F, Escalon BL, Zorzano A, Soares AM, Raldúa D, Zebrafish Models for Human Acute
Organophosphorus Poisoning, Sci Rep, 2015 Oct 22; 5:15591, doi: 10.1038/srep15591, Erratum in: Sci Rep.
2016; 6:17244, PubMed PMID: 26489395; PubMed Central PMCID: PMC4614985.
APPENDIX 1 39

139. Felemban SG, Garner AC, Smida FA, Boocock DJ, Hargreaves AJ, Dickenson JM, Phenyl Saligenin
Phosphate Induced Caspase-3 and c-Jun N-Terminal Kinase Activation in Cardiomyocyte-Like Cells, Chem
Res Toxicol, 2015 Nov 16; 28(11):2179–91, doi: 10.1021/acs.chemrestox.5b00338, Epub 2015 Oct 23,
PubMed PMID: 26465378.

140. Feng F, Ge J, Li Y, He S, Zhong J, Liu X, Yu X, Enhanced degradation of chlorpyrifos in rice (Oryza sativa L.)
by five strains of endophytic bacteria and their plant growth promotional ability, Chemosphere, 2017 Oct;
184:505–13, doi: 10.1016/j.chemosphere.2017.05.178, Epub 2017 Jun 3, PubMed PMID: 28622646.

141. Feng L, Yang G, Zhu L, Xu X, Gao F, Mu J, Xu Y, Enhancement removal of endocrine-disrupting pesticides


and nitrogen removal in a biofilm reactor coupling of biodegradable Phragmites communis and elastic filler for
polluted source water treatment, Bioresour Technol, 2015; 187:331–37, doi: 10.1016/j.biortech.2015.03.095,
Epub 2015 Mar 28, PubMed PMID: 25863211.

142. Fernandes MES, Alves FM, Pereira RC, Aquino LA, Fernandes FL, Zanuncio JC, Lethal and sublethal effects
of seven insecticides on three beneficial insects in laboratory assays and field trials, Chemosphere, 2016 Aug;
156:45–55, doi: 10.1016/j.chemosphere.2016.04.115, Epub 2016 May 6, PubMed PMID: 27160634.

143. Ferrario C, Parolini M, De Felice B, Villa S, Finizio A, Linking sub-individual and supra-individual effects in
Daphnia magna exposed to sub-lethal concentration of chlorpyrifos, Environ Pollut, 2018 Apr; 235:411–18, doi:
10.1016/j.envpol.2017.12.113, Epub 2018 Jan 5, PubMed PMID: 29310084.

144. Fiedler N, Rohitrattana J, Siriwong W, Suttiwan P, Ohman Strickland P, Ryan PB, Rohlman DS, Panuwet P,
Barr DB, Robson MG, Neurobehavioural effects of exposure to organophosphates and pyrethroid pesticides
among Thai children, Neurotoxicology, 2015 May; 48:90–9, doi: 10.1016/j.neuro.2015.02.003, Epub 2015 Feb
24, PubMed PMID: 25721160; PubMed Central PMCID: PMC4442703.

145. Firmin S, Bahi-Jaber N, Abdennebi-Najar L, Food contaminants and programming of type 2 diabetes: recent
findings from animal studies, J Dev Orig Health Dis, 2016 Oct; 7(5):505–12, Epub 2016 Jun 13, PubMed
PMID: 27292028.

146. Fluegge KR, Nishioka M, Wilkins JR 3rd, Effects of simultaneous prenatal exposures to organophosphate and
synthetic pyrethroid insecticides on infant neurodevelopment at three months of age, J Environ Toxicol Public
Health, 2016 May; 1:60–73, doi: 10.5281/zenodo.218417, Epub 2016 May 19, PubMed PMID: 28580452;
PubMed Central PMCID: PMC5455339.

147. Flunker LK, Nutter TJ, Johnson RD, Cooper BY, DEET potentiates the development and persistence of
anticholinesterase dependent chronic pain signs in a rat model of Gulf War Illness pain, Toxicol Appl
Pharmacol, 2017 Feb 1; 316:48–62, doi: 10.1016/j.taap.2016.12.014, Epub 2016 Dec 23, PubMed PMID:
28025109.

148. Fu DJ, Li P, Song J, Zhang SY, Xie HZ, Mechanisms of synergistic neurotoxicity induced by two high risk
pesticide residues - Chlorpyrifos and Carbofuran via oxidative stress, Toxicol In Vitro, 2018 Oct 30; 54:338–44,
doi: 10.1016/j.tiv.2018.10.016, PubMed PMID: 30385350.

149. Fu Y, Liu F, Zhao C, Zhao Y, Liu Y, Zhu G, Distribution of chlorpyrifos in rice paddy environment and its
potential dietary risk, J Environ Sci (China), 2015 Sep 1; 35:101–07, doi: 10.1016/j.jes.2015.02.015, Epub
2015 Jun 29. PubMed PMID: 26354698.

150. Galea KS, MacCalman L, Jones K, Cocker J, Teedon P, Cherrie JW, van Tongeren M, Comparison of
residents' pesticide exposure with predictions obtained using the UK regulatory exposure assessment
40 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

approach, Regul Toxicol Pharmacol, 2015 Nov; 73(2):634–43, doi: 10.1016/j.yrtph.2015.09.012, Epub 2015
Sep 11, PubMed PMID: 26364754.

151. Gambardella C, Nichino D, Iacometti C, Ferrando S, Falugi C, Faimali M, Long term exposure to low dose
neurotoxic pesticides affects hatching, viability and cholinesterase activity of Artemia sp, Aquat Toxicol, 2018
Mar; 196:79–89, doi: 10.1016/j.aquatox.2018.01.006, Epub 2018 Jan 12, PubMed PMID: 29358113.

152. Gao J, Naughton SX, Beck WD, Hernandez CM, Wu G, Wei Z, Yang X, Bartlett MG, Terry AV Jr, Chlorpyrifos
and chlorpyrifos oxon impair the transport of membrane bound organelles in rat cortical axons,
Neurotoxicology, 2017 Sep; 62:111–23, doi: 10.1016/j.neuro.2017.06.003, Epub 2017 Jun 12, PubMed PMID:
28600141.

153. Gao J, Naughton SX, Wulff H, Singh V, Beck WD, Magrane J, Thomas B, Kaidery NA, Hernandez CM, Terry
AV Jr, Diisopropylfluorophosphate Impairs the Transport of Membrane-Bound Organelles in Rat Cortical
Axons, J Pharmacol Exp Ther, 2016 Mar; 356(3):645–55, doi: 10.1124/jpet.115.230839, Epub 2015 Dec 30,
PubMed PMID: 26718240; PubMed Central PMCID: PMC4767389.

154. Garcia-Reyero N, Escalon L, Prats E, Faria M, Soares AM, Raldúa D, Targeted Gene Expression in Zebrafish
Exposed to Chlorpyrifos-Oxon Confirms Phenotype-Specific Mechanisms Leacceptable daily intakeng to
Adverse Outcomes, Bull Environ Contam Toxicol, 2016 Jun; 96(6):707–13, doi: 10.1007/s00128-016-1798-3,
Epub 2016 Apr 16, PubMed PMID: 27086301; PubMed Central PMCID: PMC4882348.

155. Gawdzik JC, Yue MS, Martin NR, Elemans LMH, Lanham KA, Heideman W, Rezendes R, Baker TR, Taylor
MR, Plavicki JS, sox9b is required in cardiomyocytes for cardiac morphogenesis and function, Sci Rep, 2018
Sep 17; 8(1):13906, doi: 10.1038/s41598-018-32125-7, PubMed PMID: 30224706; PubMed Central PMCID:
PMC6141582.

156. Ginsberg G, Vulimiri SV, Lin YS, Kancherla J, Foos B, Sonawane B, A framework and case studies for
evaluation of enzyme ontogeny in children's health risk evaluation, J Toxicol Environ Health A, 2017; 80(10–
12):569–93, doi:10.1080/15287394.2017.1369915, Epub 2017 Sep 11, PubMed PMID: 28891786.

157. Glazer L, Wells CN, Drastal M, Odamah KA, Galat RE, Behl M, Levin ED, Developmental exposure to low
concentrations of two brominated flame retardants, BDE-47 and BDE-99, causes life-long behavioural
alterations in zebrafish, Neurotoxicology, 2018 May; 66:221–32, doi: 10.1016/j.neuro.2017.09.007, Epub 2017
Sep 19, PubMed PMID: 28935585; PubMed Central PMCID: PMC5858967.

158. Gómez-Canela C, Prats E, Tauler R, Raldúa D, Analysis of neurobehavioural data by chemometric methods in
ecotoxicological studies, Ecotoxicol Environ Saf, 2017 Nov; 145:583–90, doi: 10.1016/j.ecoenv.2017.08.013,
Epub 2017 Aug 9, PubMed PMID: 28802139.

159. Gómez-Giménez B, Felipo V, Cabrera-Pastor A, Agustí A, Hernández-Rabaza V, Llansola M, Developmental


Exposure to Pesticides Alters Motor Activity and Coordination in Rats: Sex Differences and Underlying
Mechanisms, Neurotox Res, 2018 Feb; 33(2):247–58, doi: 10.1007/s12640-017-9823-9, Epub 2017 Oct 3,
PubMed PMID: 28975519.

160. Gómez-Giménez B, Llansola M, Hernández-Rabaza V, Cabrera-Pastor A, Malaguarnera M, Agusti A, Felipo V,


Sex-dependent effects of developmental exposure to different pesticides on spatial learning, The role of
induced neuroinflammation in the hippocampus, Food Chem Toxicol, 2017 Jan; 99:135–48, doi:
10.1016/j.fct.2016.11.028, Epub 2016 Nov 29, PubMed PMID: 27908700.

161. Gong T, Liu R, Che Y, Xu X, Zhao F, Yu H, Song C, Liu Y, Yang C, Engineering Pseudomonas putida KT2440
for simultaneous degradation of carbofuran and chlorpyrifos, Microb Biotechnol, 2016 Nov; 9(6):792–800, doi:
APPENDIX 1 41

10.1111/1751-7915.12381, Epub 2016 Jul 15, PubMed PMID: 27418102; PubMed Central PMCID:
PMC5072195.

162. Gorecki L, Korabecny J, Musilek K, Malinak D, Nepovimova E, Dolezal R, JunvD, Soukup O, Kuca K, SAR
study to find optimal cholinesterase reactivator against organophosphorous nerve agents and pesticides, Arch
Toxicol, 2016 Dec; 90(12):2831–59, Epub 2016 Aug 31, Review, PubMed PMID: 27582056.

163. Grabovska S, Salyha Y, ADHD-like behaviour in the offspring of female rats exposed to low chlorpyrifos doses
before pregnancy, Arh Hig Rada Toksikol, 2015 Jun; 66(2):121–7, doi: 10.1515/aiht-2015-66-2624, PubMed
PMID: 26110473.

164. Grabovska SV, Salyha YT, The effect of chronic intoxication with low doses of chlorpyrifos on the behavioural
parameters of female rats, Fiziol Zh. 2015; 61(2):94–101, Ukrainian, PubMed PMID: 26387166.

165. Grimalt S, Harbeck S, Shegunova P, Seghers J, Sejerøe-Olsen B, Emteborg H, Dabrio M, Development of a


new cucumber reference material for pesticide residue analysis: feasibility study for material processing,
homogeneity and stability assessment, Anal Bioanal Chem, 2015 Apr; 407(11):3083–91, doi:10.1007/s00216-
015-8476-x, Epub 2015 Jan 28, PubMed PMID: 25627789; PubMed Central PMCID: PMC4383829.

166. Guardia-Escote L, Basaure P, Blanco J, Cabré M, Pérez-Fernández C, Sánchez-Santed F, Domingo JL,


Colomina MT, Postnatal exposure to chlorpyrifos produces long-term effects on spatial memory and the
cholinergic system in mice in a sex- and APOE genotype-dependent manner, Food Chem Toxicol, 2018 Dec;
122:1–10, doi: 10.1016/j.fct.2018.09.069, Epub 2018 Sep 29, PubMed PMID: 30278244.

167. Guijarro C, Fuchs K, Bohrn U, Stütz E, Wölfl S, Simultaneous detection of multiple bioactive pollutants using a
multiparametric biochip for water quality monitoring, Biosens Bioelectron, 2015 Oct 15; 72:71–9, doi:
10.1016/j.bios.2015.04.092, Epub 2015 Apr 28, PubMed PMID: 25957833.

168. Guo W, Yan X, Zhao G, Han R, Increased Efficacy of Entomopathogenic Nematode-Insecticide Combinations
Against Holotrichia oblita (Coleoptera: Scarabaeidae), J Econ Entomol, 2017 Feb 1; 110(1):41–51, doi:
10.1093/jee/tow241, PubMed PMID: 28017929.

169. Guo-Ross SX, Meek EC, Chambers JE, Carr RL, Effects of Chlorpyrifos or Methyl Parathion on Regional
Cholinesterase Activity and Muscarinic Receptor Subtype Binding in Juvenile Rat Brain, J Toxicol Pharmacol,
2017; 1. pii: 018, Epub 2017 Dec 30, PubMed PMID: 30035273; PubMed Central PMCID: PMC6052801.

170. Hagstrom D, Hirokawa H, Zhang L, Racceptable daily intakec Z, Taylor P, Collins ES, Planarian
cholinesterase: in vitro characterization of an evolutionarily ancient enzyme to study organophosphorus
pesticide toxicity and reactivation, Arch Toxicol, 2017 Aug; 91(8):2837–47, doi: 10.1007/s00204-016-1908-3,
Epub 2016 Dec 18, PubMed PMID: 27990564.

171. Halimah M, Ismail BS, Nashriyah M, Maznah Z, Mobility Studies of (14)C-Chlorpyrifos in Malaysian Oil Palm
Soils, Bull Environ Contam Toxicol, 2016 Jan; 96(1):120–4, doi: 10.1007/s00128-015-1685-3, Epub 2015 Nov
6, PubMed PMID: 26546229.

172. Hasenbein S, Connon RE, Lawler SP, Geist J, A comparison of the sublethal and lethal toxicity of four
pesticides in Hyalella azteca and Chironomus dilutes, Environ Sci Pollut Res Int, 2015 Aug; 22(15):11327–39,
doi: 10.1007/s11356-015-4374-1, Epub 2015 Mar 26, PubMed PMID: 25804662.

173. Hasenbein S, Lawler SP, Geist J, Connon RE, The use of growth and behavioural endpoints to assess the
effects of pesticide mixtures upon aquatic organisms, Ecotoxicology, 2015 May; 24(4):746–59, doi:
10.1007/s10646-015-1420-1, Epub 2015 Jan 29, PubMed PMID: 25630500.
42 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

174. Hassani S, Sepand MR, Jafari A, Jaafari J, Rezaee R, Zeinali M, Tavakoli F,Razavi-Azarkhiavi K, Protective
effects of curcumin and vitamin E against chlorpyrifos-induced lung oxidative damage, Hum Exp Toxicol, 2015
Jun; 34(6):668–76, doi: 10.1177/0960327114550888, Epub 2014 Sep 17.,PubMed PMID: 25233897.

175. Hayat K, Afzal M, Aqueel MA, Ali S, Saeed MF, Khan QM, Ashfaq M, Damalas CA, Insecticide exposure
affects DNA and antioxidant enzymes activity in honey bee species Apis florea and A. dorsata: Evidence from
Punjab, Pakistan, Sci Total Environ, 2018 Sep 1; 635:1292–301, doi: 10.1016/j.scitotenv.2018.04.221, Epub
2018 Apr 24, PubMed PMID: 29710582.

176. He L, Liu B, Tian J, Lu F, Li X, Tian Y, Culturable epiphytic bacteria isolated from Teleogryllus occipitalus
crickets metabolize insecticides, Arch Insect Biochem Physiol, 2018 Oct; 99(2):e21501, doi:
10.1002/arch.21501, Epub 2018 Aug 17, PubMed PMID: 30120789.

177. He Y, Xu B, Li W, Yu H, Silver nanoparticle-based chemiluminescent sensor array for pesticide discrimination,


J Agric Food Chem, 2015 Mar 25; 63(11):2930–4, doi: 10.1021/acs.jafc.5b00671, Epub 2015 Mar 12, PubMed
PMID: 25751408.

178. Heffernan AL, English K, Toms L, Calafat AM, Valentin-Blasini L, Hobson P, Broomhall S, Ware RS, Jagals P,
Sly PD, Mueller JF, Cross-sectional biomonitoring study of pesticide exposures in Queensland, Australia, using
pooled urine samples, Environ Sci Pollut Res Int, 2016 Dec; 23(23):23436–48, Epub 2016 Sep 10, PubMed
PMID: 27613627.

179. Helali I, Ferchichi S, Maaouia A, Aouni M, Harizi H, Modulation of macrophage functionality induced in vitro by
chlorpyrifos and carbendazim pesticides, J Immunotoxicol, 2016 Sep; 13(5):745–50, doi:
10.1080/1547691X.2016.1181124, Epub 2016 Jul 18, PubMed PMID: 27429139.

180. Hernandez CM, Beck WD, Naughton SX, Poddar I, Adam BL, Yanasak N, Middleton C, Terry AV Jr, Repeated
exposure to chlorpyrifos leads to prolonged impairments of axonal transport in the living rodent brain,
Neurotoxicology, 2015 Mar; 47:17–26, doi: 10.1016/j.neuro.2015.01.002, Epub 2015 Jan 19, PubMed PMID:
25614231.

181. Hertz-Picciotto I, Sass JB, Engel S, Bennett DH, Bradman A, Eskenazi B, Lanphear B, Whyatt R,
Organophosphate exposures during pregnancy and child neurodevelopment: Recommendations for essential
policy reforms, PLoS Med, 2018 Oct 24; 15(10):e1002671, doi: 10.1371/journal.pmed.1002671, eCollection
2018 Oct, PubMed PMID: 30356230.

182. Hickmann F, de Morais AF, Bronzatto ES, Giacomelli T, Guedes JVC, Bernardi O, Susceptibility of the Lesser
Mealworm, Alphitobius diaperinus (Coleoptera: Tenebrionidae), From Broiler Farms of Southern Brazil to
Insecticides, J Econ Entomol, 2018 Apr 2; 111(2):980–5, doi: 10.1093/jee/toy059, PubMed PMID: 29534190.

183. Holzer G, Besson M, Lambert A, François L, Barth P, Gillet B, Hughes S, Piganeau G, Leulier F, Viriot L,
Lecchini D, Laudet V, Fish larval recruitment to reefs is a thyroid hormone-mediated metamorphosis sensitive
to the pesticide chlorpyrifos, Elife, 2017 Oct 30; 6, pii: e27595, doi: 10.7554/eLife.27595, PubMed PMID:
29083300; PubMed Central PMCID: PMC5662287.

184. Hou W, Jiang C, Zhou X, Qian K, Wang L, Shen Y, Zhao Y, Increased Expression of P-Glycoprotein Is
Associated With Chlorpyrifos Resistance in the German Cockroach (Blattodea: Blattellidae), J Econ Entomol,
2016 Dec 1; 109(6):2500–5, doi: 10.1093/jee/tow141, PubMed PMID: 27634281.

185. Houbraken M, Senaeve D, Fevery D, Spanoghe P, Influence of adjuvants on the dissipation of fenpropimorph,
pyrimethanil, chlorpyrifos and lindane on the solid/gas interface, Chemosphere, 2015 Nov; 138:357–63, doi:
10.1016/j.chemosphere.2015.06.040, Epub 2015 Jun 29, PubMed PMID: 26133697.
APPENDIX 1 43

186. Howell GE 3rd, Kondakala S, Holdridge J, Lee JH, Ross MK, Inhibition of cholinergic and non-cholinergic
targets following subacute exposure to chlorpyrifos in normal and high fat fed male C57BL/6J mice, Food
Chem Toxicol, 2018 Aug; 118:821–9, doi: 10.1016/j.fct.2018.06.051, Epub 2018 Jun 21, PubMed PMID:
29935250.

187. Howell GE 3rd, Mulligan C, Young D, Kondakala S, Exposure to chlorpyrifos increases neutral lipid
accumulation with accompanying increased de novo lipogenesis and decreased triglyceride secretion in
McArdle-RH7777 hepatoma cells, Toxicol In Vitro, 2016 Apr; 32:181–9, doi: 10.1016/j.tiv.2016.01.002, Epub
2016 Jan 7, PubMed PMID: 26773343.

188. Hunt L, Bonetto C, Marrochi N, Scalise A, Fanelli S, Liess M, Lydy MJ, Chiu MC, Resh VH, Species at Risk
(SPEAR) index indicates effects of insecticides on stream invertebrate communities in soy production regions
of the Argentine Pampas, Sci Total Environ, 2017 Feb 15; 580:699–709, doi: 10.1016/j.scitotenv.2016.12.016,
Epub 2016 Dec 14, PubMed PMID: 27986319.

189. 195. Ismail AA, Bonner MR, Hendy O, Abdel Rasoul G, Wang K, Olson JR, Rohlman DS, Comparison of
neurological health outcomes between two adolescent cohorts exposed to pesticides in Egypt, PLoS One,
2017 Feb 23; 12(2):e0172696, doi: 10.1371/journal.pone.0172696, eCollection 2017, PubMed PMID:
28231336; PubMed Central PMCID: PMC5322908.

190. Ismail AA, Wang K, Olson JR, Bonner MR, Hendy O, Abdel Rasoul G, Rohlman DS, The impact of repeated
organophosphorus pesticide exposure on biomarkers and neurobehavioural outcomes among adolescent
pesticide applicators, J Toxicol Environ Health A, 2017; 80(10–12):542–55, doi:
10.1080/15287394.2017.1362612, Epub 2017 Sep 7, PubMed PMID: 28880741; PubMed Central PMCID:
PMC5648326.

191. Jaabiri Kamoun I, Jegede OO, Owojori OJ, Bouzid J, Gargouri R, Römbke J, Effects of deltamethrin,
dimethoate, and chlorpyrifos on survival and reproduction of the collembolan Folsomia candida and the
predatory mite Hypoaspis aculeifer in two African and two European soils, Integr Environ Assess Manag, 2018
Jan; 14(1):92–104, doi: 10.1002/ieam.1966, Epub 2017 Aug 22, PubMed PMID: 28755498.

192. Jang Y, Lee AY, Jeong SH, Park KH, Paik MK, Cho NJ, Kim JE, Cho MH, Chlorpyrifos induces NLRP3
inflammasome and pyroptosis/apoptosis via mitochondrial oxidative stress in human keratinocyte HaCaT cells,
Toxicology, 2015 Dec 2; 338:37–46, doi: 10.1016/j.tox.2015.09.006, Epub 2015 Oct 3, PubMed PMID:
26435000.

193. Janssens L, Op de Beeck L, Stoks R, Stoichiometric Responses to an Agricultural Pesticide Are Modified by
Predator Cues, Environ Sci Technol, 2017 Jan 3; 51(1):581–88, doi: 10.1021/acs.est.6b03381, Epub 2016
Dec 21, PubMed PMID: 27936640.

194. Janssens L, Stoks R, Chlorpyrifos-induced oxidative damage is reduced under warming and predation risk:
Explaining antagonistic interactions with a pesticide, Environ Pollut, 2017 Jul; 226:79–88, doi:
10.1016/j.envpol.2017.04.012, Epub 2017 Apr 12, PubMed PMID: 28411497.

195. Janssens L, Verberk W, Stoks R, A widespread morphological antipredator mechanism reduces the sensitivity
to pesticides and increases the susceptibility to warming, Sci Total Environ, 2018 Jun 1; 626:1230–5, doi:
10.1016/j.scitotenv.2018.01.179, Epub 2018 Feb 19, PubMed PMID: 29898530.

196. Jantunen LM, Wong F, Gawor A, Kylin H, Helm PA, Stern GA, Strachan WM, Burniston DA, Bidleman TF, 20
Years of Air-Water Gas Exchange Observations for Pesticides in the Western Arctic Ocean, Environ Sci
44 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Technol, 2015 Dec 1; 49(23):13844–52, doi: 10.1021/acs.est.5b01303, Epub 2015 Aug 18, PubMed PMID:
26196214.

197. Jayawardena UA, Rohr JR, Navaratne AN, Amerasinghe PH, Rajakaruna RS, Combined Effects of Pesticides
and Trematode Infections on Hourglass Tree Frog Polypedates cruciger, Ecohealth, 2016 Mar; 13(1):111–22,
doi: 10.1007/s10393-016-1103-2, Epub 2016 Feb 24, PubMed PMID: 26911919; PubMed Central PMCID:
PMC4852980.

198. Jeon HJ, Lee YH, Kim MJ, Choi SD, Park BJ, Lee SE, Integrated biomarkers induced by chlorpyrifos in two
different life stages of zebrafish (Danio rerio) for environmental risk assessment, Environ Toxicol Pharmacol,
2016 Apr; 43:166–74, doi: 10.1016/j.etap.2016.03.010, Epub 2016 Mar 15, PubMed PMID: 26998704.

199. Jin JX, Jin DC, Li WH, Cheng Y, Li FL, Ye ZC, Monitoring Trends in Insecticide Resistance of Field Populations
of Sogatella furcifera (Hemiptera: Delphacidae) in Guizhou Province, China, 2012–15, J Econ Entomol, 2017
Apr 1; 110(2):641–50, doi: 10.1093/jee/tox027, PubMed PMID: 28334150; PubMed Central PMCID:
PMC5387993.

200. Jin Y, Liu Z, Peng T, Fu Z, The toxicity of chlorpyrifos on the early life stage of zebrafish: a survey on the
endpoints at development, locomotor behaviour, oxidative stress and immunotoxicity, Fish Shellfish Immunol,
2015 Apr; 43(2):405–14, doi: 10.1016/j.fsi.2015.01.010, Epub 2015 Jan 26, PubMed PMID: 25634256.

201. Joly Condette C, Bach V, Mayeur C, Gay-Quéheillard J, Khorsi-Cauet H, Chlorpyrifos Exposure During
Perinatal Period Affects Intestinal Microbiota Associated with Delay of Maturation of Digestive Tract in Rats, J
Pediatr Gastroenterol Nutr, 2015 Jul; 61(1):30–40, doi: 10.1097/MPG.0000000000000734, PubMed PMID:
25643018.

202. Joly Condette C, Elion Dzon B, Hamdad F, Biendo M, Bach V, Khorsi-Cauet H, Use of molecular typing to
investigate bacterial translocation from the intestinal tract of chlorpyrifos-exposed rats, Gut Pathog, 2016 Nov
5; 8:50, eCollection 2016, PubMed PMID: 27826358; PubMed Central PMCID: PMC5097847.

203. Joseph SV, Repellent Effects of Insecticides Against Protaphorura fimata (Collembola: Poduromorpha:
Onychiuridae), J Econ Entomol, 2018 Apr 2; 111(2):747–54, doi: 10.1093/jee/tox375, PubMed PMID:
29361113.

204. Judge SJ, Savy CY, Campbell M, Dodds R, Gomes LK, Laws G, Watson A, Blain PG, Morris CM, Gartside SE,
Mechanism for the acute effects of organophosphate pesticides on the adult 5-HT system, Chem Biol Interact,
2016 Feb 5; 245:82–9, doi: 10.1016/j.cbi.2015.12.014, Epub 2015 Dec 22, PubMed PMID: 26721196; PubMed
Central PMCID: PMC4732990.

205. Kang XL, Zhang M, Wang K, Qiao XF, Chen MH, Molecular cloning, expression pattern of multidrug resistance
associated protein 1 (MRP1, ABCC1) gene, and the synergistic effects of verapamil on toxicity of two
insecticides in the bird cherry-oat aphid, Arch Insect Biochem Physiol, 2016 May; 92(1):65–84, doi:
10.1002/arch.21334, PubMed PMID: 27110952.

206. Karami A, Goh YM, Jahromi MF, Lazorchak JM, Abdullah M, Courtenay SC, Diploid and triploid African catfish
(Clarias gariepinus) differ in biomarker responses to the pesticide chlorpyrifos, Sci Total Environ, 2016 Jul 1;
557–8:204–11, doi: 10.1016/j.scitotenv.2016.03.030, Epub 2016 Mar 18, PubMed PMID: 26994807.

207. Khalid S, Hashmi I, Jamal Khan S, Qazi IA, Nasir H, Effect of metal ions and petrochemicals on
bioremediation of chlorpyrifos in aerobic sequencing batch bioreactor (ASBR), Environ Sci Pollut Res Int, 2016
Oct; 23(20):20646–60, Epub 2016 Jul 28, PubMed PMID: 27470246.
APPENDIX 1 45

208. Khalid S, Hashmi I, Khan SJ, Bacterial assisted degradation of chlorpyrifos: The key role of environmental
conditions, trace metals and organic solvents, J Environ Manage, 2016 Mar 1; 168:1–9, doi:
10.1016/j.jenvman.2015.11.030, Epub 2015 Dec 12, PubMed PMID: 26692411.

209. Khalil F, Qiu X, Kang IJ, Abo-Ghanema I, Shimasaki Y, Oshima Y, Comparison of social behaviour responses
of Japanese medaka (Oryzias latipes) to lethal and sublethal chlorpyrifos concentrations at different exposure
times, Ecotoxicol Environ Saf, 2017 Nov; 145:78–82, doi: 10.1016/j.ecoenv.2017.07.007, Epub 2017 Jul 12,
PubMed PMID: 28708984.

210. Khan HA, Akram W, Iqbal N, Selection and Preliminary Mechanism of Resistance to Profenofos in a Field
Strain of Musca domestica (Diptera: Muscidae) from Pakistan, J Med Entomol, 2015 Sep; 52(5):1013–7, doi:
10.1093/jme/tjv095, Epub 2015 Jul 8, PubMed PMID: 26336235.

211. Khan N, Athar T, Fouad H, Umar A, Ansari ZA, Ansari SG, Application of pristine and doped SnO(2)
nanoparticles as a matrix for agro-hazardous material (organophosphate) detection, Sci Rep, 2017 Feb 14;
7:42510, doi: 10.1038/srep42510, PubMed PMID: 28195202; PubMed Central PMCID: PMC5307345.

212. Kim BM, Lee JW, Seo JS, Shin KH, Rhee JS, Lee JS, Modulated expression and enzymatic activity of the
monogonont rotifer Brachionus koreanus Cu/Zn- and Mn-superoxide dismutase (SOD) in response to
environmental biocides, Chemosphere, 2015 Feb; 120:470–8, doi: 10.1016/j.chemosphere.2014.08.042, Epub
2014 Sep 26, PubMed PMID: 25260044.

213. Kim HY, Wegner SH, Van Ness KP, Park JJ, Pacheco SE, Workman T, Hong S, Griffith W, Faustman EM,
Differential epigenetic effects of chlorpyrifos and arsenic in proliferating and differentiating human neural
progenitor cells, Reprod Toxicol, 2016 Oct; 65:212–23, doi: 10.1016/j.reprotox.2016.08.005, Epub 2016 Aug
11, PubMed PMID: 27523287; PubMed Central PMCID: PMC5067221.

214. Kim RO, Kim BM, Jeong CB, Lee JS, Rhee JS, Effects of chlorpyrifos on life cycle parameters, cytochrome
P450S expression, and antioxidant systems in the monogonont rotifer Brachionus koreanus, Environ Toxicol
Chem, 2016 Jun; 35(6):1449–57, doi: 10.1002/etc.3288, Epub 2016 Mar 15, PubMed PMID: 26496856.

215. Kokushi E, Uno S, Pal S, Koyama J, Effects of chlorpyrifos on the metabolome of the freshwater carp,
Cyprinus carpio, Environ Toxicol, 2015 Mar; 30(3):253–60, doi: 10.1002/tox.21903, Epub 2013 Aug 30,
PubMed PMID: 23997021.

216. Kondakala S, Lee JH, Ross MK, Howell GE 3rd, Effects of acute exposure to chlorpyrifos on cholinergic and
non-cholinergic targets in normal and high-fat fed male C57BL/6J mice, Toxicol Appl Pharmacol, 2017 Dec 15;
337:67–75, doi: 10.1016/j.taap.2017.10.019, Epub 2017 Oct 31, PubMed PMID: 29097212.

217. Kopjar N, Žunec S, Mendaš G, Micek V, Kašuba V, Mikolić A, Lovaković BT, Milić M, Pavičić I, Čermak AMM,
Pizent A, Lucić Vrdoljak A, Želježić D, Evaluation of chlorpyrifos toxicity through a 28-day study:
Cholinesterase activity, oxidative stress responses, parent compound/metabolite levels, and primary DNA
damage in blood and brain tissue of adult male Wistar rats, Chem Biol Interact, 2018 Jan 5; 279:51–63, doi:
10.1016/j.cbi.2017.10.029, Epub 2017 Nov 4, PubMed PMID: 29108776.

218. Krief S, Berny P, Gumisiriza F, Gross R, Demeneix B, Fini JB, Chapman CA, Chapman LJ, Seguya A, Wasswa
J, Agricultural expansion as risk to endangered wildlife: Pesticide exposure in wild chimpanzees and baboons
displaying facial dysplasia, Sci Total Environ, 2017 Nov 15; 598:647–56, doi: 10.1016/j.scitotenv.2017.04.113,
Epub 2017 Apr 25, PubMed PMID: 28454037.

219. Kwon DH, Kang TJ, Kim YH, Lee SH, Phenotypic- and Genotypic-Resistance Detection for Adaptive
Resistance Management in Tetranychus urticae Koch, PLoS One, 2015 Nov 6; 10(11):e0139934, doi:
46 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

10.1371/journal.pone.0139934, eCollection 2015, PubMed PMID: 26545209; PubMed Central PMCID:


PMC4636269.

220. Lahouel A, Kebieche M, Lakroun Z, Rouabhi R, Fetoui H, Chtourou Y, Djamila Z, Soulimani R,


Neurobehavioural deficits and brain oxidative stress induced by chronic low dose exposure of persistent
organic pollutants mixture in adult female rat, Environ Sci Pollut Res Int, 2016 Oct; 23(19):19030–40, doi:
10.1007/s11356-016-6913-9, Epub 2016 May 30, Erratum in: Environ Sci Pollut Res Int. 2016 Oct; 23
(19):19041–2, PubMed PMID: 27240828.

221. Lamb T, Selvarajah LR, Mohamed F, Jayamanne S, Gawarammana I, Mostafa A, Buckley NA, Roberts MS,
Eddleston M, High lethality and minimal variation after acute self-poisoning with carbamate insecticides in Sri
Lanka - implications for global suicide prevention, Clin Toxicol (Phila), 2016 Sep; 54(8):624–31, doi:
10.1080/15563650.2016.1187735, Epub 2016 Jun 2, PubMed PMID: 27252029; PubMed Central PMCID:
PMC4950420.

222. Lan A, Kalimian M, Amram B, Kofman O, Prenatal chlorpyrifos leads to autism-like deficits in C57Bl6/J mice.
Environ Health, 2017 May 2; 16(1):43, doi: 10.1186/s12940-017-0251-3, PubMed PMID: 28464876; PubMed
Central PMCID: PMC5414283.

223. Laporte B, Gay-Quéheillard J, Bach V, Villégier AS, Developmental neurotoxicity in the progeny after maternal
gavage with chlorpyrifos, Food Chem Toxicol, 2018 Mar; 113:66–72, doi: 10.1016/j.fct.2018.01.026, Epub
2018 Feb 3, PubMed PMID: 29421768.

224. Latifi AM, Karami A, Khodi S, Efficient Surface Display of Diisopropylfluorophosphatase (DFPase) in E. coli for
Biodegradation of Toxic Organophosphorus Compounds (DFP and Cp), Appl Biochem Biotechnol, 2015 Oct;
177(3):624–36, doi: 10.1007/s12010-015-1766-0, Epub 2015 Aug 4, PubMed PMID: 26239441.

225. Latini LA, Indaco MM, Aguiar MB, Monza LB, Parolo ME, Melideo CF, Savini MC, Loewy RM, An integrated
approach for assessing the migration behaviour of chlorpyrifos and carbaryl in the unsaturated soil zone, J
Environ Sci Health B, 2018; 53(7):469–75, doi: 0.1080/03601234.2018.1455353, Epub 2018 Apr 6, PubMed
PMID: 29624471.

226. Lee CW, Su H, Lee RH, Lin YP, Tsai YD, Wu DC, Shiea J, Point-of-care identification of organophosphates in
gastric juice by ambient mass spectrometry in emergency settings, Clin Chim Acta, 2018 Oct; 485:288–97, doi:
10.1016/j.cca.2018.07.002, Epub 2018 Jul 3, PubMed PMID: 30018012.

227. Lee I, Eriksson P, Fredriksson A, Buratovic S, Viberg H, Developmental neurotoxic effects of two pesticides:
Behaviour and biomolecular studies on chlorpyrifos and carbaryl, Toxicol Appl Pharmacol, 2015 Nov 1;
288(3):429–38, doi: 10.1016/j.taap.2015.08.014, Epub 2015 Aug 24, PubMed PMID: 26314619.

228. Lee YH, Kang HM, Kim MS, Lee JS, Jeong CB, Lee JS, The protective role of multixenobiotic resistance
(MXR)-mediated ATP-binding cassette (ABC) transporters in biocides-exposed rotifer Brachionus koreanus,
Aquat Toxicol, 2018 Feb; 195:129–36, doi: 10.1016/j.aquatox.2017.12.016, Epub 2017 Dec 30, PubMed
PMID: 29306793.

229. Lee YS, Lewis JA, Ippolito DL, Hussainzada N, Lein PJ, Jackson DA, Stallings JD, Repeated exposure to
neurotoxic levels of chlorpyrifos alters hippocampal expression of neurotrophins and neuropeptides,
Toxicology, 2016 Jan 18; 340:53–62, doi: 10.1016/j.tox.2016.01.001, Epub 2016 Jan 13, PubMed PMID:
26775027; PubMed Central PMCID: PMC4777610.
APPENDIX 1 47

230. Levin ED, Learning about cognition risk with the racceptable daily intakeal-arm maze in the developmental
neurotoxicology battery, Neurotoxicol Teratol, 2015 Nov-Dec; 52(Pt A):88–92, doi: 10.1016/j.ntt.2015.05.007,
Epub 2015 May 23, PubMed PMID: 26013674; PubMed Central PMCID: PMC4656139.

231. Li D, Huang Q, Lu M, Zhang L, Yang Z, Zong M, Tao L, The organophosphate insecticide chlorpyrifos confers
its genotoxic effects by inducing DNA damage and cell apoptosis, Chemosphere, 2015 Sep; 135:387–93, doi:
10.1016/j.chemosphere.2015.05.024, Epub 2015 May 22, PubMed PMID: 26002045.

232. Li D, Wu X, Yu X, Huang Q, Tao L, Synergistic effect of non-ionic surfactants Tween 80 and PEG6000 on
cytotoxicity of insecticides, Environ Toxicol Pharmacol, 2015 Mar; 39(2):677–82, doi:
10.1016/j.etap.2014.12.015, Epub 2014 Dec 31, PubMed PMID: 25699500.

233. Li F, Yuan Y, Meng P, Wu M, Li S, Chen B, Probabilistic acute risk assessment of cumulative exposure to
organophosphorus and carbamate pesticides from dietary vegetables and fruits in Shanghai populations,
Food Addit Contam Part A Chem Anal Control Expo Risk Assess, 2017 May; 34(5):819–31, doi:
10.1080/19440049.2017.1279350, Epub 2017 Feb 3, PubMed PMID: 28077027.

234. Li JX, Shi SM, Xue J, Residual status and risk assessment of forbidden and restricted pesticides of
organophosphorus in Loincerae Japonicae Flos, Zhongguo Zhong Yao Za Zhi, 2018 Jun; 43(11):2333–38, doi:
10.19540/j.cnki.cjcmm.20180309.002, Chinese, PubMed PMID: 29945387.

235. Li R, Zhang L, Shi Q, Guo Y, Zhang W, Zhou B, A protective role of autophagy in TDCIPP-induced
developmental neurotoxicity in zebrafish larvae, Aquat Toxicol, 2018 Jun; 199:46–54, doi:
10.1016/j.aquatox.2018.03.016, Epub 2018 Mar 15, PubMed PMID: 29605586.

236. Lian N, Gou L, Wang Q, Peng S, Xu P, Combined cytotoxicity mechanism of chlorpyrifos and carbofuran
pesticides in vitro, Wei Sheng Yan Jiu, 2017 Jul; 46(4):621–7, Chinese, PubMed PMID: 29903186.

237. Liendro N, Ferrari A, Mardirosian M, Lascano CI, Venturino A, Toxicity of the insecticide chlorpyrifos to the
South American toad Rhinella arenarum at larval developmental stage, Environ Toxicol Pharmacol, 2015 Mar;
39(2):525–35, doi: 10.1016/j.etap.2014.12.022, Epub 2015 Jan 8, PubMed PMID: 25681703.

238. Liu H, Jiang G, Zhang Y, Chen F, Li X, Yue J, Ran C, Zhao Z, Effect of six insecticides on three populations of
Bactrocera (Tetradacus) minax (Diptera:Tephritidae), Curr Pharm Biotechnol, 2015; 16(1):77–83, PubMed
PMID: 25564253.

239. Liu J, Parsons L, Pope C, Comparative effects of parathion and chlorpyrifos on endocannabinoid and
endocannabinoid-like lipid metabolites in rat striatum, Neurotoxicology, 2015 Sep; 50:20–7, doi:
10.1016/j.neuro.2015.07.006, Epub 2015 Jul 26, PubMed PMID: 26215119; PubMed Central PMCID:
PMC4767275.

240. Liu J, Pope C, The cannabinoid receptor antagonist AM251 increases paraoxon and chlorpyrifos oxon toxicity
in rats, Neurotoxicology, 2015 Jan; 46:12–8, doi: 10.1016/j.neuro.2014.11.001, Epub 2014 Nov 20, PubMed
PMID: 25447325; PubMed Central PMCID: PMC4448943.

241. Liu J, Zhang X, Yang M, Hu M, Zhong G, Toxicity assessment of chlorpyrifos-degracceptable daily intakeng
fungal bio-composites and their environmental risks, Sci Rep, 2018 Feb 1; 8(1):2152, doi: 10.1038/s41598-
018-20265-9, PubMed PMID: 29391422; PubMed Central PMCID: PMC5794795.

242. Liu J, Zhang X, Zhang Y, Preparation and release behaviour of chlorpyrifos adsolubilized into layered zinc
hydroxide nitrate intercalated with dodecylbenzenesulfonate, ACS Appl Mater Interfaces, 2015 Jun 3;
7(21):11180–8, doi: 10.1021/acsami.5b01068, Epub 2015 May 22, PubMed PMID: 25970564.
48 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

243. Liu Y, Li X, Zhou C, Liu F, Mu W,Toxicity of nine insecticides on fourvnatural enemies of Spodoptera exigua,
Sci Rep, 2016 Dec 13; 6:39060, doi:v10.1038/srep39060, PubMed PMID: 27958333; PubMed Central PMCID:
PMC5153630.

244. Liu Y, Mo R, Tang F, Fu Y, Guo Y, Influence of different formulations on chlorpyrifos behaviour and risk
assessment in bamboo forest of China, Environ Sci Pollut Res Int, 2015 Dec; 22(24):20245–54, doi:
10.1007/s11356-015-5272-2, Epub 2015 Aug 27, PubMed PMID: 26308925.

245. Liu YQ, Liu B, Ali A, Luo SP, Lu YH, Liang GM, Insecticide Toxicity to Adelphocoris lineolatus (Hemiptera:
Miridae) and its Nymphal Parasitoid Peristenus spretus (Hymenoptera: Braconidae), J Econ Entomol, 2015
Aug; 108(4):1779–85, doi: 10.1093/jee/tov144, Epub 2015 Jun 6, PubMed PMID: 26470319.

246. Locker AR, Michalovicz LT, Kelly KA, Miller JV, Miller DB, O'Callaghan JP, Corticosterone primes the
neuroinflammatory response to Gulf War Illness-relevant organophosphates independently of
acetylcholinesterase inhibition, J Neurochem, 2017 Aug; 142(3):444–55, doi: 10.1111/jnc.14071, Epub 2017
Jun 14, PubMed PMID: 28500787; PubMed Central PMCID: PMC5575502.

247. López-Granero C, Ruiz-Muñoz AM, Nieto-Escámez FA, Colomina MT, Aschner M, Sánchez-Santed F, Chronic
dietary chlorpyrifos causes long-term spatial memory impairment and thigmotaxic behaviour, Neurotoxicology,
2016 Mar; 53:85–92, doi: 10.1016/j.neuro.2015.12.016, Epub 2015 Dec 31, PubMed PMID: 26748072.

248. Lou Y, Li Y, Effects of Chlorpyrifos on Dopaminergic Neuronal Viability with Activation of Microglia, Sheng Wu
Yi Xue Gong Cheng Xue Za Zhi, 2016 Jun; 33(3):506–11, Chinese, PubMed PMID: 29709151.

249. Lozowicka B, Health risk for children and adults consuming apples with pesticide residue, Sci Total Environ,
2015 Jan 1; 502:184–98, doi: 10.1016/j.scitotenv.2014.09.026, Epub 2014 Sep 26, PubMed PMID: 25260164.

250. Lu C, Liu X, Liu C, Wang J, Li C, Liu Q, Li Y, Li S, Sun S, Yan J, Shao J, Chlorpyrifos Induces MLL
Translocations Through Caspase 3-Dependent Genomic Instability and Topoisomerase II Inhibition in Human
Fetal Liver Hematopoietic Stem Cells, Toxicol Sci, 2015 Oct; 147(2):588–606, doi: 10.1093/toxsci/kfv153,
Epub 2015 Jul 20, PubMed PMID: 26198043.

251. Lu K, Wang Y, Chen X, Zhang Z, Li Y, Li W, Zhou Q, Characterization and functional analysis of a


carboxylesterase gene associated with chlorpyrifos resistance in Nilaparvata lugens (Stål), Comp Biochem
Physiol C Toxicol Pharmacol, 2017 Dec; 203:12–20, doi: 10.1016/j.cbpc.2017.10.005, Epub 2017 Oct 18,
PubMed PMID: 29054582.

252. Lukowicz C, Ellero-Simatos S, Régnier M, Polizzi A, Lasserre F, Montagner A, Lippi Y, Jamin EL, Martin JF,
Naylies C, Canlet C, Debrauwer L, Bertrand-Michel J, Al Saati T, Théodorou V, Loiseau N, Mselli-Lakhal L,
Guillou H, Gamet-Payrastre L, Metabolic Effects of a Chronic Dietary Exposure to a Low-Dose Pesticide
Cocktail in Mice: Sexual Dimorphism and Role of the Constitutive Androstane Receptor, Environ Health
Perspect, 2018 Jun 25; 126(6):067007, doi: 10.1289/EHP2877, eCollection 2018 Jun, PubMed PMID:
29950287; PubMed Central PMCID: PMC6084886.

253. Ma XY, Wang XC, Wang D, Ngo HH, Zhang Q, Wang Y, Dai D, Function of a landscape lake in the reduction of
biotoxicity related to trace organic chemicals from reclaimed water, J Hazard Mater, 2016 Nov 15; 318:663–70,
doi: 10.1016/j.jhazmat.2016.07.050, Epub 2016 Jul 21, PubMed PMID: 27475464.

254. Ma Z, Li J, Zhang Y, Shan C, Gao X, Inheritance mode and mechanisms of resistance to imidacloprid in the
house fly Musca domestica (Diptera:Muscidae) from China, PLoS One, 2017 Dec 11; 12(12):e0189343, doi:
10.1371/journal.pone.0189343, eCollection 2017, PubMed PMID: 29228021; PubMed Central PMCID:
PMC5724887.
APPENDIX 1 49

255. Maciel W, Lopes WD, Cruz B, Teixeira W, Felippelli G, Sakamoto CA, Fávero FC, Buzzulini C, Soares V,
Gomes LV, Bichuette M, da Costa AJ, Effects of Haematobia irritans infestation on weight gain of Nelore
calves assessed with different antiparasitic treatment schemes, Prev Vet Med, 2015 Jan 1; 118(1):182–6, doi:
10.1016/j.prevetmed.2014.11.006, Epub 2014 Nov 15, PubMed PMID: 25465474.

256. Malagnoux L, Capowiez Y, Rault M, Impact of insecticide exposure on the predation activity of the European
earwig Forficula auricularia, Environ Sci Pollut Res Int, 2015 Sep; 22(18):14116–26, doi: 10.1007/s11356-015-
4520–9, Epub 2015 May 12, PubMed PMID: 25963069.

257. Malagnoux L, Marliac G, Simon S, Rault M, Capowiez Y, Management strategies in apple orchards influence
earwig community, Chemosphere, 2015 Apr; 124:156–62, doi: 10.1016/j.chemosphere.2014.12.024, Epub
2015 Jan 7, PubMed PMID: 25577700.

258. Malinowska E, Jankowski K, Pesticide residues in some herbs growing in agricultural areas in Poland, Environ
Monit Assess, 2015 Dec; 187(12):775, doi:10.1007/s10661-015-4997-1, Epub 2015 Nov 26, PubMed PMID:
26612566; PubMed Central PMCID: PMC4661190.

259. Mamczarz J, Pescrille JD, Gavrushenko L, Burke RD, Fawcett WP, DeTolla LJ Jr, Chen H, Pereira EFR,
Albuquerque EX, Spatial learning impairment in prepubertal guinea pigs prenatally exposed to the
organophosphorus pesticide chlorpyrifos: Toxicological implications, Neurotoxicology, 2016 Sep; 56:17–28,
doi: 10.1016/j.neuro.2016.06.008, Epub 2016 Jun 11, PubMed PMID: 27296654; PubMed Central PMCID:
PMC5673496.

260. Mansoor MM, Afzal M, Raza AB, Akram Z, Waqar A, Afzal MB, Post-exposure temperature influence on the
toxicity of conventional and new chemistry insecticides to green lacewing Chrysoperla carnea (Stephens)
(Neuroptera: Chrysopidae), Saudi J Biol Sci, 2015 May; 22(3):317–21, doi: 10.1016/j.sjbs.2014.10.008, Epub
2014 Oct 30, PubMed PMID: 25972753; PubMed Central PMCID: PMC4423650.

261. Mansour SA, Gamet-Payrastre L, Ameliorative effect of vitamin E to mouse dams and their pups following
exposure of mothers to chlorpyrifos during gestation and lactation periods, Toxicol Ind Health, 2016 Jul;
32(7):1179–96, doi: 10.1177/0748233714548207, Epub 2014 Sep 18, PubMed PMID: 25234640.

262. Mao XL, Liu J, Li XK, Chi JJ, Liu YJ, Resistance risk, cross-resistance and biochemical resistance mechanism
of Laodelphax striatellus to buprofezin, Ying Yong Sheng Tai Xue Bao, 2016 Jan; 27(1):255–62, Chinese.
PubMed PMID: 27228617.

263. Maqbool Z, Hussain S, Imran M, Mahmood F, Shahzad T, Ahmed Z, Azeem F, Muzammil S, Perspectives of
using fungi as bioresource for bioremediation of pesticides in the environment: a critical review, Environ Sci
Pollut Res Int, 2016 Sep; 23(17):16904–25, doi: 10.1007/s11356-016-7003-8, Epub 2016 Jun 8. Review,
PubMed PMID: 27272922.

264. Maryoung LA, Blunt B, Tierney KB, Schlenk D, Sublethal toxicity of chlorpyrifos to salmonid olfaction after
hypersaline acclimation, Aquat Toxicol, 2015 Apr; 161:94–101, doi: 10.1016/j.aquatox.2015.01.026, Epub
2015 Jan 29, PubMed PMID: 25697678.

265. Masiá A, Vásquez K, Campo J, Picó Y, Assessment of two extraction methods to determine pesticides in soils,
sediments and sludges, Application to the Túria River Basin, J Chromatogr A. 2015 Jan 23; 1378:19–31, doi:
10.1016/j.chroma.2014.11.079, Epub 2014 Dec 18, PubMed PMID: 25573188.

266. Maslova OV, Senko OV, Efremenko EN, The influence of enzymatic removal of chlorpyrifos from feed grain-
mixture on the biochemical parameters of rat blood, Biomed Khim, 2017 Nov; 63(6):559–64, doi:
10.18097/PBMC20176306559, Russian, PubMed PMID: 29251619.
50 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

267. McClelland SJ, Bendis RJ, Relyea RA, Woodley SK, Insecticide-induced changes in amphibian brains: How
sublethal concentrations of chlorpyrifos directly affect neurodevelopment, Environ Toxicol Chem, 2018 Oct;
37(10):2692–98, doi: 10.1002/etc.4240, Epub 2018 Sep 5, PubMed PMID: 30187530.

268. Meijer M, Brandsema JA, Nieuwenhuis D, Wijnolts FM, Dingemans MM, Westerink RH, Inhibition of Voltage-
Gated Calcium Channels After Subchronic and Repeated Exposure of PC12 Cells to Different Classes of
Insecticides, Toxicol Sci, 2015 Oct; 147(2):607–17, doi: 10.1093/toxsci/kfv154, Epub 2015 Jul 17, PubMed
PMID: 26187449.

269. Mekonnen TF, Panne U, Koch M, Electrochemistry coupled online to liquid chromatography-mass
spectrometry for fast simulation of biotransformation reactions of the insecticide chlorpyrifos, Anal Bioanal
Chem, 2017 May; 409(13):3359–68, doi: 10.1007/s00216-017-0277-y, Epub 2017 Mar 10, PubMed PMID:
28283716.

270. Michalovicz LT, Locker AR, Kelly KA, Miller JV, Barnes Z, Fletcher MA, Miller DB, Klimas NG, Morris M, Lasley
SM, O'Callaghan JP, Corticosterone and pyridostigmine/DEET exposure attenuate peripheral cytokine
expression: Supporting a dominant role for neuroinflammation in a mouse model of Gulf War Illness,
Neurotoxicology, 2018 Oct 16; 70:26–32, doi: 10.1016/j.neuro.2018.10.006, PubMed PMID: 30339781.

271. Mie A, Rudén C, Grandjean P, Safety of Safety Evaluation of Pesticides: developmental neurotoxicity of
chlorpyrifos and chlorpyrifos-methyl, Environ Health, 2018 Nov 16; 17(1):77, doi: 10.1186/s12940-018-0421-y,
PubMed PMID:30442131.

272. Mikhail MW, Abd El-Halim AS, Soliman MI, Efficacy of three insecticides on rat flea (Xenopsylla cheopis)
infesting rodents in giza governorate, Egypt, J Egypt Soc Parasitol, 2016 Apr; 46(1):131–4, PubMed PMID:
27363049.

273. Miller JV, LeBouf RF, Kelly KA, Michalovicz LT, Ranpara A, Locker AR, Miller DB, O'Callaghan JP, The
Neuroinflammatory Phenotype in a Mouse Model of Gulf War Illness is Unrelated to Brain Regional Levels of
Acetylcholine as Measured by Quantitative HILIC-UPLC-MS/MS, Toxicol Sci, 2018 Oct 1; 165(2):302–13, doi:
10.1093/toxsci/kfy130, PubMed PMID: 29846716.

274. Mohamed WR, Mehany ABM, Hussein RM, Alpha lipoic acid protects against chlorpyrifos-induced toxicity in
Wistar rats via modulating the apoptotic pathway, Environ Toxicol Pharmacol, 2018 Apr; 59:17–23, doi:
10.1016/j.etap.2018.02.007, Epub 2018 Feb 23, PubMed PMID: 29500983.

275. Moore DR, Priest CD, Olson AD, Teed RS, A probabilistic risk assessment for the Kirtland's warbler potentially
exposed to chlorpyrifos and malathion during the breeding season and migration, Integr Environ Assess
Manag, 2018 Mar; 14(2):252–69, doi: 10.1002/ieam.2004, Epub 2017 Dec 28, PubMed PMID: 29105950.

276. Morgado RG, Gomes PA, Ferreira NG, Cardoso DN, Santos MJ, Soares AM, Loureiro S, Toxicity interaction
between chlorpyrifos, mancozeb and soil moisture to the terrestrial isopod Porcellionides pruinosus,
Chemosphere, 2016 Feb; 144:1845–53, doi: 10.1016/j.chemosphere.2015.10.034, Epub 2015 Nov 11,
PubMed PMID: 26539709.

277. Morselli M, Terzaghi E, Galimberti F, Di Guardo A, Pesticide fate in cultivated mountain basins: The improved
DynAPlus model for predicting peak exposure and directing sustainable monitoring campaigns to protect
aquatic ecosystems, Chemosphere, 2018 Nov; 210:204–14, doi: 10.1016/j.chemosphere.2018.06.181, Epub
2018 Jul 4, PubMed PMID: 30005341.
APPENDIX 1 51

278. Morteza Z, Mousavi SB, Baghestani MA, Aitio A, An assessment of agricultural pesticide use in Iran, 2012–14,
J Environ Health Sci Eng, 2017 Apr 24; 15:10, doi: 10.1186/s40201-017-0272-4, eCollection 2017, PubMed
PMID: 28451437; PubMed Central PMCID: PMC5404313.

279. Mosbah R, Yousef MI, Maranghi F, Mantovani A, Protective role of Nigella sativa oil against reproductive
toxicity, hormonal alterations, and oxidative damage induced by chlorpyrifos in male rats, Toxicol Ind Health,
2016 Jul; 32(7):1266–77, doi: 10.1177/0748233714554675, Epub 2014 Nov 25, PubMed PMID: 25425536.

280. Moser VC, Pacceptable daily intakella S, Esterase detoxication of acetylcholinesterase inhibitors using human
liver samples in vitro, Toxicology, 2016 Apr 15; 353–4:11-20, doi: 10.1016/j.tox.2016.04.006, Epub 2016 Apr
27, PubMed PMID: 27132127.

281. Mosqueira B, Soma DD, Namountougou M, Poda S, Diabaté A, Ali O, Fournet F, Baldet T, Carnevale P, Dabiré
RK, Mas-Coma S, Pilot study on the combination of an organophosphate-based insecticide paint and
pyrethroid-treated long lasting nets against pyrethroid resistant malaria vectors in Burkina Faso, Acta Trop,
2015 Aug; 148:162–9, doi: 10.1016/j.actatropica.2015.04.010, Epub 2015 May 7, PubMed PMID: 25959771.

282. Mousavi MM, Arefhosseini S, Alizadeh Nabili AA, Mahmoudpour M, Nemati M, Development of an ultrasound-
assisted emulsification microextraction method for the determination of chlorpyrifos and organochlorine
pesticide residues in honey samples using gas chromatography with mass spectrometry, J Sep Sci, 2016 Jul;
39(14):2815–22, doi: 10.1002/jssc.201600197, Epub 2016 Jun 17, PubMed PMID: 27214344.

283. Moyano P, Del Pino J, Anadon MJ, Díaz MJ, Gómez G, Frejo MT, Toxicogenomic profile of apoptotic and
necrotic SN56 basal forebrain cholinergic neuronal loss after acute and long-term chlorpyrifos exposure,
Neurotoxicol Teratol, 2017 Jan-Feb; 59:68–73, doi: 10.1016/j.ntt.2016.10.002, Epub 2016 Oct 11, PubMed
PMID: 27737797.

284. Moyano P, Frejo MT, Anadon MJ, García JM, Díaz MJ, Lobo M, Sola E, García J, Del Pino J, SN56 neuronal
cell death after 24 h and 14 days chlorpyrifos exposure through glutamate transmission dysfunction, increase
of GSK-3β enzyme, β-amyloid and tau protein levels, Toxicology, 2018 Jun 1; 402–3:17–27, doi:
10.1016/j.tox.2018.04.003, Epub 2018 Apr 14, PubMed PMID: 29665406.

285. Muangphra P, Tharapoom K, Euawong N, Namchote S, Gooneratne R, Chronic toxicity of commercial


chlorpyrifos to earthworm Pheretima peguana, Environ Toxicol, 2016 Nov; 31(11):1450–9, doi:
10.1002/tox.22150, Epub 2015 Apr 30, PubMed PMID: 25926403.

286. Mullen BR, Ross B, Chou JW, Khankan R, Khialeeva E, Bui K, Carpenter EM, A Complex Interaction Between
Reduced Reelin Expression and Prenatal Organophosphate Exposure Alters Neuronal Cell Morphology, ASN
Neuro, 2016 Jun 30; 8(3). pii: 1759091416656253, doi: 10.1177/1759091416656253, Print 2016 Jun, PubMed
PMID: 27364165; PubMed Central PMCID: PMC4962342.

287. Mullins RJ, Xu S, Pereira EF, Pescrille JD, Todd SW, Mamczarz J, Albuquerque EX, Gullapalli RP, Prenatal
exposure of guinea pigs to the organophosphorus pesticide chlorpyrifos disrupts the structural and functional
integrity of the brain, Neurotoxicology, 2015 May; 48:9–20, doi: 10.1016/j.neuro.2015.02.002, Epub 2015 Feb
19, PubMed PMID: 25704171; PubMed Central PMCID: PMC4442734.

288. Muslim M, Ansari MS, Hasan F, Non-target toxicity of synthetic insecticides on the biological performance and
population growth of Bracon hebetor Say, Ecotoxicology, 2018 Sep; 27(7):1019–31, doi: 10.1007/s10646-018-
1947-z. Epub 2018 May 24, PubMed PMID: 29797170.

289. Nagarajan A, Bodhicharla R, Winter J, Anbalagan C, Morgan K, Searle M, Nazir A, Adenle A, Fineberg A,
Brady D, Vere K, Richens J, O'Shea P, Bell D, de-Pomerai D, A Fluorescence Resonance Energy Transfer
52 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Assay for Monitoring α- Synclein Aggregation in a Caenorhabditis Elegans Model for Parkinson's Disease,
CNS Neurol Disord Drug Targets, 2015; 14(8):1054–68, PubMed PMID: 26295817.

290. Napon G, Dafferner AJ, Saxena A, Lockridge O, Identification of Carboxylesterase, Butyrylcholinesterase,


Acetylcholinesterase, Paraoxonase, and Albumin Pseudoesterase in Guinea Pig Plasma through
Nondenaturing Gel Electrophoresis, Comp Med, 2018 Oct 1; 68(5):367–74, doi:10.30802/AALAS-CM-18-
000047, Epub 2018 Oct 2, PubMed PMID: 30278860; PubMed Central PMCID: PMC6200033.

291. Nasr HM, El-Demerdash FM, El-Nagar WA, Neuro and renal toxicity induced by chlorpyrifos and abamectin in
rats: Toxicity of insecticide mixture, Environ Sci Pollut Res Int, 2016 Jan; 23(2):1852–9, doi: 10.1007/s11356-
015-5448-9, Epub 2015 Sep 24, PubMed PMID: 26403246.

292. Ng TK, Gahan LR, Schenk G, Ollis DL, Altering the substrate specificity of methyl parathion hydrolase with
directed evolution, Arch Biochem Biophys, 2015 May 1; 573:59–68, doi: 10.1016/j.abb.2015.03.012, Epub
2015 Mar 19, PubMed PMID: 25797441.

293. Nganchamung T, Robson MG, Siriwong W, Association between blood cholinesterase activity,
organophosphate pesticide residues on hands, and health effects among chili farmers in Ubon Ratchathani
Province, northeastern Thailand, Rocz Panstw Zakl Hig, 2017; 68(2):175–83, PubMed PMID: 28646835.

294. Nutter TJ, Johnson RD, Cooper BY, A delayed chronic pain like condition with decreased Kv channel activity in
a rat model of Gulf War Illness pain syndrome, Neurotoxicology, 2015 Dec; 51:67–79, doi:
10.1016/j.neuro.2015.09.010, Epub 2015 Sep 26, PubMed PMID: 26409647.

295. Ojha A, Gupta YK, Evaluation of genotoxic potential of commonly used organophosphate pesticides in
peripheral blood lymphocytes of rats, Hum Exp Toxicol, 2015 Apr; 34(4):390–400, doi:
10.1177/0960327114537534, Epub 2014 Sep 8, PubMed PMID: 25205738.

296. Ojha A, Gupta YK, Study of commonly used organophosphate pesticides that induced oxidative stress and
apoptosis in peripheral blood lymphocytes of rats, Hum Exp Toxicol, 2017 Nov; 36(11):1158–68, doi:
10.1177/0960327116680273, Epub 2016 Dec 9, PubMed PMID: 27941166.

297. Oliveri AN, Bailey JM, Levin ED, Developmental exposure to organophosphate flame retardants causes
behavioural effects in larval and adult zebrafish, Neurotoxicol Teratol, 2015 Nov-Dec; 52(Pt B):220–7, doi:
10.1016/j.ntt.2015.08.008, Epub 2015 Sep 5, PubMed PMID: 26344674; PubMed Central PMCID:
PMC4769864.

298. Olmos V, Marro M, Loza-Alvarez P, Raldúa D, Prats E, Padrós F, Piña B, Tauler R, de Juan A, Combining
hyperspectral imaging and chemometrics to assess and interpret the effects of environmental stressors on
zebrafish eye images at tissue level, J Biophotonics, 2018 Mar; 11(3), doi: 10.1002/jbio.201700089, Epub
2017 Oct 4, PubMed PMID: 28766927.

299. Onder S, Dafferner AJ, Schopfer LM, Xiao G, Yerramalla U, Tacal O, Blake TA, Johnson RC, Lockridge O,
Monoclonal Antibody That Recognizes Diethoxyphosphotyrosine-Modified Proteins and Peptides Independent
of Surrounding Amino Acids, Chem Res Toxicol, 2017 Dec 18; 30(12):2218–28, doi:
10.1021/acs.chemrestox.7b00288, Epub 2017 Nov 28, PubMed PMID: 29137457; PubMed Central PMCID:
PMC6006506.

300. Op de Beeck L, Verheyen J, Stoks R, Competition magnifies the impact of a pesticide in a warming world by
reducing heat tolerance and increasing autotomy, Environ Pollut, 2018 Feb; 233:226–34, doi:
10.1016/j.envpol.2017.10.071, Epub 2017 Nov 5, PubMed PMID: 29096295.
APPENDIX 1 53

301. Oriel S, Kofman O, Strain dependent effects of conditioned fear in adult C57Bl/6 and Balb/C mice following
postnatal exposure to chlorpyrifos: relation to expression of brain acetylcholinesterase mRNA, Front Behav
Neurosci, 2015 Apr 29; 9:110, doi: 10.3389/fnbeh.2015.00110, eCollection 2015, PubMed PMID: 25972795;
PubMed Central PMCID: PMC4413781.

302. Pallotta MM, Barbato V, Pinton A, Acloque H, Gualtieri R, Talevi R, Jammes H, Capriglione T, In vitro exposure
to CPF affects bovine sperm epigenetic gene methylation pattern and the ability of sperm to support
fertilization and embryo development, Environ Mol Mutagen, 2018 Oct 26, doi: 10.1002/em.22242, PubMed
PMID: 30365181.

303. Pallotta MM, Ronca R, Carotenuto R, Porreca I, Turano M, Ambrosino C, Capriglione T, Specific Effects of
Chronic Dietary Exposure to Chlorpyrifos on Brain Gene Expression-A Mouse Study, Int J Mol Sci, 2017 Nov
20; 18(11), pii: E2467, doi: 10.3390/ijms18112467, PubMed PMID: 29156651; PubMed Central PMCID:
PMC5713433.

304. Palumbo JC, Prabhaker N, Reed DA, Perring TM, Castle SJ, Huang TI, Susceptibility of Bagrada hilaris
(Hemiptera: Pentatomidae) to Insecticides in Laboratory and Greenhouse Bioassays, J Econ Entomol, 2015
Apr; 108(2):672–82, doi: 10.1093/jee/tov010, Epub 2015 Mar 5, PubMed PMID: 26470178.

305. Papadakis EN, Tsaboula A, Kotopoulou A, Kintzikoglou K, Vryzas Z, Papadopoulou-Mourkidou E, Pesticides in


the surface waters of Lake Vistonis Basin, Greece: Occurrence and environmental risk assessment, Sci Total
Environ, 2015 Dec 1; 536:793–802, doi: 10.1016/j.scitotenv.2015.07.099, Epub 2015 Aug 4, PubMed PMID:
26254079.

306. Papadakis EN, Tsaboula A, Vryzas Z, Kotopoulou A, Kintzikoglou K, Papadopoulou-Mourkidou E, Pesticides in


the rivers and streams of two river basins in northern Greece, Sci Total Environ, 2018 May 15; 624:732–43,
doi: 10.1016/j.scitotenv.2017.12.074, Epub 2017 Dec 27, PubMed PMID: 29272842.

307. Papadopoulou ES, Lagos S, Spentza F, Vidiadakis E, Karas PA, Klitsinaris T, Karpouzas DG, The dissipation
of fipronil, chlorpyrifos, fosthiazate and ethoprophos in soils from potato monoculture areas: first evidence for
the enhanced biodegradation of fosthiazate, Pest Manag Sci, 2016 May; 72(5):1040–50, doi: 10.1002/ps.4092,
Epub 2015 Sep 7, PubMed PMID: 26261048.

308. Pardo S, Martínez AM, Figueroa JI, Chavarrieta JM, Viñuela E, Rebollar-Alviter Á, Miranda MA, Valle J, Pineda
S, Insecticide resistance of adults and nymphs of Asian citrus psyllid populations from Apatzingán Valley,
Mexico, Pest Manag Sci, 2018 Jan; 74(1):135–140, doi: 10.1002/ps.4669, Epub 2017 Sep 11, PubMed PMID:
28719016.

309. Patnaik R, Padhy RN. Human Umbilical Cord Blood-Derived Neural Stem Cell Line as a Screening Model for
Toxicity, Neurotox Res, 2017 Apr; 31(3):319–26, doi: 10.1007/s12640-016-9681-x, Epub 2016 Nov 2, PubMed
PMID: 27807796.

310. Patra RW, Chapman JC, Lim RP, Gehrke PC, Sunderam RM, Interactions betweenwater temperature and
contaminant toxicity to freshwater fish, Environ Toxicol Chem, 2015 Aug; 34(8):1809–17, doi:
10.1002/etc.2990, Epub 2015 Jun 1, PubMed PMID: 26033197.

311. Peiris DC, Dhanushka T, Low doses of chlorpyrifos interfere with spermatogenesis of rats through reduction of
sex hormones, Environ Sci Pollut Res Int, 2017 Sep; 24(26):20859–67, doi: 10.1007/s11356-017-9617-x,
Epub 2017 Jul 18, PubMed PMID: 28721614.

312. Pérez-Rodríguez J, Martínez-Blay V, Soto A, Selfa J, Monzó C, Urbaneja A, Tena A, Aggregation Patterns,
Sampling Plan, and Economic Injury Levels for the New Citrus Pest Delottococcus aberiae (Hemiptera:
54 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Pseudococcidae), J Econ Entomol, 2017 Dec 5; 110(6):2699–706, doi: 10.1093/jee/tox258, PubMed PMID:
29220519.

313. Peris-Sampedro F, Basaure P, Reverte I, Cabré M, Domingo JL, Colomina MT, Chronic exposure to
chlorpyrifos triggered body weight increase and memory impairment depending on human apoE
polymorphisms in a targeted replacement mouse model, Physiol Behav, 2015 May 15; 144:37–45, doi:
10.1016/j.physbeh.2015.03.006, Epub 2015 Mar 5, PubMed PMID: 25747767.

314. Peris-Sampedro F, Blanco J, Cabré M, Basaure P, Guardia-Escote L, Domingo JL, Sánchez DJ, Colomina MT,
New mechanistic insights on the metabolic-disruptor role of chlorpyrifos in apoE mice: a focus on insulin- and
leptin-signalling pathways, Arch Toxicol, 2018 May; 92(5):1717–28, doi: 10.1007/s00204-018-2174-3, Epub
2018 Feb 5, PubMed PMID: 29404631.

315. Peris-Sampedro F, Cabré M, Basaure P, Reverte I, Domingo JL, Teresa Colomina M, Adulthood dietary
exposure to a common pesticide leads to an obese-like phenotype and a diabetic profile in apoE3 mice,
Environ Res, 2015 Oct; 142:169–76, doi: 10.1016/j.envres.2015.06.036, Epub 2015 Jul 7, PubMed PMID:
26162960.

316. Peris-Sampedro F, Reverte I, Basaure P, Cabré M, Domingo JL, Colomina MT, Apolipoprotein E (APOE)
genotype and the pesticide chlorpyrifos modulate attention, motivation and impulsivity in female mice in the 5-
choice serial reaction time task, Food Chem Toxicol, 2016 Jun; 92:224–35, doi: 10.1016/j.fct.2016.03.029,
Epub 2016 Apr 20, PubMed PMID: 27106138.

317. Phung DT, Connell D, Chu C, A new method for setting guidelines to protect human health from agricultural
exposure by using chlorpyrifos as an example, Ann Agric Environ Med, 2015; 22(2):275–80, doi:
10.5604/12321966.1152080, PubMed PMID: 26094523.

318. Poet TS, Timchalk C, Bartels MJ, Smith JN, McDougal R, Juberg DR, Price PS, Use of a probabilistic
PBPK/PD model to calculate Data Derived Extrapolation Factors for chlorpyrifos, Regul Toxicol Pharmacol,
2017 Jun; 86:59–73, doi: 10.1016/j.yrtph.2017.02.014, Epub 2017 Feb 24, PubMed PMID: 28238854.

319. Porreca I, D'Angelo F, De Franceschi L, Mattè A, Ceccarelli M, Iolascon A, Zamò A, Russo F, Ravo M, Tarallo
R, Scarfò M, Weisz A, De Felice M, Mallardo M, Ambrosino C, Pesticide toxicogenomics across scales: in vitro
transcriptome predicts mechanisms and outcomes of exposure in vivo, Sci Rep, 2016 Dec 1; 6:38131, doi:
10.1038/srep38131, PubMed PMID: 27905518; PubMed Central PMCID: PMC5131489.

320. Potter TL, Coffin AW, Assessing pesticide wet deposition risk within a small agricultural watershed in the
Southeastern Coastal Plain (USA), Sci Total Environ, 2017 Feb 15; 580:158–67, doi:
10.1016/j.scitotenv.2016.11.020, Epub 2016 Dec 24, PubMed PMID: 28027802.

321. Proskocil BJ, Bruun DA, Garg JA, Villagomez CC, Jacoby DB, Lein PJ, Fryer AD, The influence of
sensitization on mechanisms of organophosphorus pesticide-induced airway hyperreactivity, Am J Respir Cell
Mol Biol, 2015 Nov; 53(5):738–47, doi: 10.1165/rcmb.2014-0444OC, PubMed PMID: 25897622; PubMed
Central PMCID: PMC4742952.

322. Purschke B, Scheibelberger R, Axmann S, Adler A, Jäger H, Impact of substrate contamination with
mycotoxins, heavy metals and pesticides on the growth performance and composition of black soldier fly
larvae (Hermetia illucens) for use in the feed and food value chain, Food Addit Contam Part A Chem Anal
Control Expo Risk Assess, 2017 Aug; 34(8):1410–20, doi: 10.1080/19440049.2017.1299946, PubMed PMID:
28278126.
APPENDIX 1 55

323. Qiao L, Guo Y, Sun X, Jiao Y, Wang X, Electrochemical immunosensor with NiAl-layered double
hydroxide/graphene nanocomposites and hollow gold nanospheres double-assisted signal amplification,
Bioprocess Biosyst Eng, 2015 Aug; 38(8):1455–68, doi: 10.1007/s00449-015-1388-5, Epub 2015 Mar 24,
PubMed PMID: 25801002.

324. Qiu X, Nomichi S, Chen K, Honda M, Kang IJ, Shimasaki Y, Oshima Y, Short-term and persistent impacts on
behaviours related to locomotion, anxiety, and startle responses of Japanese medaka (Oryzias latipes)
induced by acute, sublethal exposure to chlorpyrifos, Aquat Toxicol, 2017 Nov; 192:148–54, doi:
10.1016/j.aquatox.2017.09.012, Epub 2017 Sep 14, PubMed PMID: 28957716.

325. Radford SA, Panuwet P, Hunter RE Jr, Barr DB, Ryan PB, Degradation of Organophosphorus and Pyrethroid
Insecticides in Beverages: Implications for Risk Assessment, Toxics, 2018 Feb 2; 6(1), pii: E11, doi:
10.3390/toxics6010011, PubMed PMID: 29393904; PubMed Central PMCID: PMC5874784.

326. Radford SA, Panuwet P, Hunter RE Jr, Barr DB, Ryan PB, Production of Insecticide Degradates in Juices:
Implications for Risk Assessment, J Agric Food Chem, 2016 Jun 8; 64(22):4633–8, doi:
10.1021/acs.jafc.6b01143, Epub 2016 May 31, PubMed PMID: 27213611.

327. Rafique N, Nazir S, Akram S, Ahad K, Gohar A, Abbasi ST, Ahmed I, Rafique K, Screening of multiclass
pesticide residues in honey by SPE-GC/MSD: a pilot study, Environ Monit Assess, 2018 Oct 22; 190(11):666,
doi: 10.1007/s10661-018-7041-4, PubMed PMID: 30350126.

328. Rafique N, Tariq SR, Ahmed D, Monitoring and distribution patterns of pesticide residues in soil from
cotton/wheat fields of Pakistan, Environ Monit Assess, 2016 Dec; 188(12):695, Epub 2016 Nov 26, PubMed
PMID: 27889893.

329. Raina R, Baba NA, Verma PK, Sultana M, Singh M, Hepatotoxicity Induced by Subchronic Exposure of
Fluoride and Chlorpyrifos in Wistar Rats: Mitigating Effect of Ascorbic Acid, Biol Trace Elem Res, 2015 Aug;
166(2):157–62, doi: 10.1007/s12011-015-0263-1, Epub 2015 Feb 12, PubMed PMID: 25669166.

330. Raj P, Singh A, Kaur K, Aree T, Singh A, Singh N, Fluorescent Chemosensors for Selective and Sensitive
Detection of Phosmet/Chlorpyrifos with Octahedral Ni(2+) Complexes, Inorg Chem, 2016 May 16;
55(10):4874–83, doi: 10.1021/acs.inorgchem.6b00332, Epub 2016 Apr 26, PubMed PMID: 27115348.

331. Ramasubramanian T, Paramasivam M, Development and validation of a multiresidue method for the
simultaneous determination of organophosphorus insecticides and their toxic metabolites in sugarcane juice
and refined sugar by gas chromatography with flame photometric detection, J Sep Sci, 2016 Jun;
39(11):2164–71, doi: 10.1002/jssc.201600154, Epub 2016 May 17, PubMed PMID: 27061678.

332. Raszewski G, Lemieszek MK, Łukawski K, Juszczak M, Rzeski W, Chlorpyrifos and cypermethrin induce
apoptosis in human neuroblastoma cell line SH-SY5Y, Basic Clin Pharmacol Toxicol, 2015 Feb; 116(2):158–
67, doi: 10.1111/bcpt.12285, Epub 2014 Jul 25, PubMed PMID: 24975276.

333. Rathod AL, Garg RK, Chlorpyrifos poisoning and its implications in human fatal cases: A forensic perspective
with reference to Indian scenario, J Forensic Leg Med, 2017 Apr; 47:29–34, doi: 10.1016/j.jflm.2017.02.003,
Epub 2017 Feb 20, Review, PubMed PMID: 28259020.

334. Rattanaselanon P, Lormphongs S, Chanvaivit S, Morioka I, Sanprakhon P, An Occupational Health Education


Program for Thai Farmers Exposed to Chlorpyrifos, Asia Pac J Public Health, 2018 Oct
11:1010539518806042, doi: 10.1177/1010539518806042, PubMed PMID: 30306796.
56 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

335. Rauh VA, Polluting Developing Brains—EPA Failure on Chlorpyrifos, N Engl J Med, 2018 Mar 29;
378(13):1171–4, doi: 10.1056/NEJMp1716809, PubMed PMID: 29590550.

336. Renick VC, Weinersmith K, Vidal-Dorsch DE, Anderson TW, Effects of a pesticide and a parasite on
neurological, endocrine, and behavioural responses of an estuarine fish, Aquat Toxicol, 2016 Jan; 170:335–
43, doi: 10.1016/j.aquatox.2015.09.010, Epub 2015 Sep 25, PubMed PMID: 26454718.

337. Reygner J, Joly Condette C, Bruneau A, Delanaud S, Rhazi L, Depeint F, Abdennebi-Najar L, Bach V, Mayeur
C, Khorsi-Cauet H, Changes in Composition and Function of Human Intestinal Microbiota Exposed to
Chlorpyrifos in Oil as Assessed by the SHIME(®) Model, Int J Environ Res Public Health, 2016 Nov 4; 13(11),
pii: E1088, PubMed PMID: 27827942; PubMed Central PMCID: PMC5129298.

338. Reygner J, Lichtenberger L, Elmhiri G, Dou S, Bahi-Jaber N, Rhazi L, Depeint F, Bach V, Khorsi-Cauet H,
Abdennebi-Najar L, Inulin Supplementation Lowered the Metabolic Defects of Prolonged Exposure to
Chlorpyrifos from Gestation to Young Adult Stage in Offspring Rats, PLoS One, 2016 Oct 19;
11(10):e0164614, doi: 10.1371/journal.pone.0164614, eCollection 2016, PubMed PMID: 27760213; PubMed
Central PMCID: PMC5070743.

339. Richendrfer H, Creton R. Chlorpyrifos and malathion have opposite effects on behaviours and brain size that
are not correlated to changes in AChE activity, Neurotoxicology, 2015 Jul; 49:50–8, doi:
10.1016/j.neuro.2015.05.002, Epub 2015 May 14, PubMed PMID: 25983063; PubMed Central PMCID:
PMC4523399.

340. Rissato SR, Galhiane MS, Fernandes JR, Gerenutti M, Gomes HM, Ribeiro R, de Almeida MV, Evaluation of
Ricinus communis L. for the Phytoremediation of Polluted Soil with Organochlorine Pesticides, Biomed Res
Int, 2015; 2015:549863, doi: 10.1155/2015/549863, Epub 2015 Aug 2, PubMed PMID: 26301249; PubMed
Central PMCID: PMC4537713.

341. Rivero A, Niell S, Cerdeiras MP, Heinzen H, Cesio MV, Development of analytical methodologies to assess
recalcitrant pesticide bioremediation in biobeds at laboratory scale, Talanta, 2016 Jun 1; 153:17–22, doi:
10.1016/j.talanta.2016.02.025, Epub 2016 Feb 12, PubMed PMID: 27130084.

342. Rodríguez-Fuentes G, Rubio-Escalante FJ, Noreña-Barroso E, Escalante-Herrera KS, Schlenk D, Impacts of


oxidative stress on acetylcholinesterase transcription, and activity in embryos of zebrafish (Danio rerio)
following Chlorpyrifos exposure, Comp Biochem Physiol C Toxicol Pharmacol, 2015 Jun-Jul; 172–3:19–25,
doi: 10.1016/j.cbpc.2015.04.003, Epub 2015 Apr 30, PubMed PMID: 25937383.

343. Rohlman DS, Ismail AA, Rasoul GA, Bonner MR, Hendy O, Mara K, Wang K, Olson JR, A 10-month
prospective study of organophosphorus pesticide exposure and neurobehavioural performance among
adolescents in Egypt, Cortex, 2016 Jan; 74:383–95, doi: 10.1016/j.cortex.2015.09.011, Epub 2015 Oct 20,
PubMed PMID: 26687929; PubMed Central PMCID: PMC4786370.

344. Rosalovsky VP, Grabovska SV, Salyha YT, Changes in glutathione system and Lipid peroxidation in rat blood
during the first hour after chlorpyrifos Exposure, Ukr Biochem J, 2015 Sep-Oct; 87(5):124–32, PubMed PMID:
26717603.

345. Ross MK, Pluta K, Bittles V, Borazjani A, Allen Crow J, Interaction of the serine hydrolase KIAA1363 with
organophosphorus agents: Evaluation of potency and kinetics, Arch Biochem Biophys, 2016 Jan 15; 590:72–
81, doi: 10.1016/j.abb.2015.11.034, Epub 2015 Nov 23, PubMed PMID: 26617293; PubMed Central PMCID:
PMC4727981.
APPENDIX 1 57

346. Ruggieri F, D'Archivio AA, Di Camillo D, Lozzi L, Maggi MA, Mercorio R, Santucci S, Development of
molecularly imprinted polymeric nanofibers by electrospinning and applications to pesticide adsorption, J Sep
Sci, 2015 May; 38(8):1402–10, doi: 10.1002/jssc.201500033, Epub 2015 Mar 14, PubMed PMID: 25677172.

347. Rugno GR, Zanardi OZ, Parra JRP, Yamamoto PT, Lethal and Sublethal Toxicity of Insecticides to the
Lacewing Ceraeochrysa Cubana, Neotrop Entomol, 2018 Aug 31, doi: 10.1007/s13744-018-0626-3, PubMed
PMID: 30168012.

348. Saeed Q, Saeed S, Ahmad F, Switching among natal and auxiliary hosts increases vulnerability of Spodoptera
exigua (Hübner) (Lepidoptera: Noctuidae) to insecticides, Ecol Evol, 2017 Mar 19; 7(8):2725–34, doi:
10.1002/ece3.2908, eCollection 2017 Apr, PubMed PMID: 28428863; PubMed Central PMCID: PMC5395457.

349. Salamzadeh J, Shakoori A, Moracceptable daily intake V. Occurrence of multiclass pesticide residues in
tomato samples collected from different markets of Iran, J Environ Health Sci Eng, 2018 May 7; 16(1):55–63,
doi: 10.1007/s40201-018-0296-4, eCollection 2018 Jun, PubMed PMID: 29983989; PubMed Central PMCID:
PMC6021478.

350. Salgado Costa C, Ronco AE, Trudeau VL, Marino D, Natale GS, Tadpoles of the horned frog Ceratophrys
ornata exhibit high sensitivity to chlorpyrifos for conventional ecotoxicological and novel bioacoustic variables,
Environ Pollut, 2018 Apr; 235:938–47, doi: 10.1016/j.envpol.2017.12.096, Epub 2018 Feb 21, PubMed PMID:
29751398.

351. Sanchez-Hernandez JC, Notario del Pino J, Domínguez J, Earthworm-induced carboxylesterase activity in
soil: Assessing the potential for detoxification and monitoring organophosphorus pesticides, Ecotoxicol
Environ Saf, 2015 Dec; 122:303–12, doi: 10.1016/j.ecoenv.2015.08.012, Epub 2015 Sep 20, PubMed PMID:
26300118.

352. Sandhu HS, Bhanwer AJ, Puri S, Retinoic acid exacerbates chlorpyrifos action in ensuing acceptable daily
intakepogenic differentiation of C3H10T½ cells in a GSK3β dependent pathway, PLoS One, 2017 Mar 14;
12(3):e0173031, doi: 10.1371/journal.pone.0173031, eCollection 2017, Erratum in: PLoS One, 2017 May 30;
12 (5):e0178999, PubMed PMID: 28291828; PubMed Central PMCID: PMC5349446.

353. Sang S, Shu B, Yi X, Liu J, Hu M, Zhong G, Cross-resistance and baseline susceptibility of Spodoptera litura
(Fabricius) (Lepidoptera: Noctuidae) to cyantraniliprole in the south of China, Pest Manag Sci, 2016 May;
72(5):922–8, doi: 10.1002/ps.4068, Epub 2015 Jul 23, PubMed PMID: 26118543.

354. Santillo MF, Liu Y, A fluorescence assay for measuring acetylcholinesterase activity in rat blood and a human
neuroblastoma cell line (SH-SY5Y), J Pharmacol Toxicol Methods, 2015 Nov-Dec; 76:15–22, doi:
10.1016/j.vascn.2015.07.002, Epub 2015 Jul 9, PubMed PMID: 26165232.

355. Santos KFA, Zanuzo Zanardi O, de Morais MR, Jacob CRO, de Oliveira MB, Yamamoto PT, The impact of six
insecticides commonly used in control of agricultural pests on the generalist predator Hippodamia convergens
(Coleoptera: Coccinellidae), Chemosphere, 2017 Nov; 186:218–26, doi: 10.1016/j.chemosphere.2017.07.165,
Epub 2017 Aug 1, PubMed PMID: 28780449.

356. Saoudi M, Hmida IB, Kammoun W, Rebah FB, Jamoussi K, Feki AE, Protective effects of oil of Sardinella
pilchardis against subacute chlorpyrifos-induced oxidative stress in female rats, Arch Environ Occup Health,
2018 Mar 4; 73(2):128–35, doi: 10.1080/19338244.2017.1317627, Epub 2017 May 3, PubMed PMID:
28394715.

357. Sastre S, Fernández Torija C, Atiénzar Pertusa I, Beltrán EM, Pablos MV, González-Doncel M, Stage-
dependent effects of chlorpyrifos on medaka (Oryzias latipes) swimming behaviour using a miniaturized swim
58 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

flume, Aquat Toxicol, 2018 Jul; 200:37–49, doi: 10.1016/j.aquatox.2018.04.008, Epub 2018 Apr 21, PubMed
PMID: 29723761.

358. Sastre S, Fernández Torija C, Carbonell G, Rodríguez Martín JA, Beltrán EM, González-Doncel M, Effects of
dietary 2,2', 4,4'-tetrabromodiphenyl ether (BDE-47) exposure on medaka (Oryzias latipes) swimming
behaviour, Environ Pollut, 2018 Feb; 233:540–51, doi: 10.1016/j.envpol.2017.08.122, Epub 2017 Nov 5,
PubMed PMID: 29102884.

359. Savy CY, Fitchett AE, McQuade R, Gartside SE, Morris CM, Blain PG, Judge SJ, Low-level repeated exposure
to diazinon and chlorpyrifos decrease anxiety-like behaviour in adult male rats as assessed by marble burying
behaviour, Neurotoxicology, 2015 Sep; 50:149–56, doi: 10.1016/j.neuro.2015.08.010, Epub 2015 Aug 19,
PubMed PMID: 26297601.

360. Schmidt HR, Radić Z, Taylor P, Quaternary and tertiary aldoxime antidotes for organophosphate exposure in a
zebrafish model system, Toxicol Appl Pharmacol, 2015 Apr 15; 284(2):197–203, doi:
10.1016/j.taap.2015.02.011, Epub 2015 Feb 17, PubMed PMID: 25701203.

361. Schmidt RJ, Kogan V, Shelton JF, Delwiche L, Hansen RL, Ozonoff S, Ma CC, McCanlies EC, Bennett DH,
Hertz-Picciotto I, Tancredi DJ, Volk HE, Combined Prenatal Pesticide Exposure and Folic Acid Intake in
Relation to Autism Spectrum Disorder, Environ Health Perspect, 2017 Sep 8; 125(9):097007, doi:
10.1289/EHP604, PubMed PMID: 28934093; PubMed Central PMCID: PMC5915192.

362. Scholes RC, Hageman KJ, Closs GP, Stirling CH, Reid MR, Gabrielsson R, Augspurger JM, Predictors of
pesticide concentrations in freshwater trout – The role of life history, Environ Pollut, 2016 Dec; 219:253–61,
doi: 10.1016/j.envpol.2016.10.017, Epub 2016 Oct 28, PubMed PMID: 27814542.

363. Scotta AV, Bongiovanni GA, Soria EA, In vitro Modulating Activity of aqueous extracts from American Plants on
Chlorpyrifos-induced toxicity on Murine Splenocytes, Rev Fac Cien Med Univ Nac Cordoba, 2017 Dec 21;
74(4):325–30, Spanish, PubMed PMID: 29902138.

364. Seifertová M, Čechová E, Llansola M, Felipo V, Vykoukalová M, Kočan A, Determination of selected


neurotoxic insecticides in small amounts of animal tissue utilizing a newly constructed mini-extractor, Anal
Bioanal Chem, 2017 Oct; 409(25):6015–26, doi: 10.1007/s00216-017-0533-1, Epub 2017 Aug 10, PubMed
PMID: 28799107.

365. Sgobbi LF, Machado SAS, Functionalized polyacrylamide as an acetylcholinesterase-inspired biomimetic


device for electrochemical sensing of organophosphorus pesticides, Biosens Bioelectron, 2018 Feb 15;
100:290–7, doi: 10.1016/j.bios.2017.09.019, Epub 2017 Sep 14, PubMed PMID: 28942211.

366. Shaffo FC, Grodzki AC, Schelegle ES, Lein PJ, The Organophosphorus Pesticide Chlorpyrifos Induces Sex-
Specific Airway Hyperreactivity in Adult Rats, Toxicol Sci, 2018 Sep 1; 165(1):244–53, doi:
10.1093/toxsci/kfy158, PubMed PMID: 29939342; PubMed Central PMCID: PMC6135637.

367. Shah PC, Kumar VR, Dastager SG, Khire JM, Phytase production by Aspergillus niger NCIM 563 for a novel
application to degrade organophosphorus pesticides, AMB Express, 2017 Dec; 7(1):66, doi: 10.1186/s13568-
017-0370-9, Epub 2017 Mar 21, PubMed PMID: 28321795; PubMed Central PMCID: PMC5359262.

368. Shah RM, Shad SA, Abbas N, Methoxyfenozide resistance of the housefly, Muscavdomestica L. (Diptera:
Muscidae): cross-resistance patterns, stability andvassociated fitness costs, Pest Manag Sci, 2017 Jan;
73(1):254–61, doi:v10.1002/ps.4296, Epub 2016 May 25, PubMed PMID: 27098995.
APPENDIX 1 59

369. Sharma S, Chadha P, Induction of neurotoxicity by organophosphate pesticide chlorpyrifos and modulating
role of cow urine, Springerplus, 2016 Aug 12; 5(1):1344, doi: 10.1186/s40064-016-3004-9, eCollection 2016,
PubMed PMID: 27588237; PubMed Central PMCID: PMC4987744.

370. Sharma S, Singh PB, Chadha P, Saini HS, Chlorpyrifos pollution: its effect on brain acetylcholinesterase
activity in rat and treatment of polluted soil by indigenous Pseudomonas sp, Environ Sci Pollut Res Int, 2017
Jan; 24(1):381–7, doi: 10.1007/s11356-016-7799-2, Epub 2016 Oct 8, PubMed PMID: 27722883.

371. Sharma S, Uggini GK, Patel V, Desai I, Balakrishnan S, Exposure to sub-lethal dose of a combination
insecticide during early embryogenesis influences the normal patterning of mesoderm resulting in incomplete
closure of ventral body wall of chicks of domestic hen, Toxicol Rep, 2018 Feb 21; 5:302–8, doi:
10.1016/j.toxrep.2018.02.005, eCollection 2018, PubMed PMID: 29556477; PubMed Central PMCID:
PMC5856662.

372. Shin HS, Seo JH, Jeong SH, Park SW, Park YI, Son SW, Kang HG, Kim JS, Effect on the H19 gene
methylation of sperm and organs of offspring after chlorpyrifos-methyl exposure during organogenesis period,
Environ Toxicol, 2015 Dec; 30(12):1355–63, doi: 10.1002/tox.21923, Epub 2015 Mar 17, PubMed PMID:
25782373.

373. Sieke C, Michalski B, Kuhl T, Probabilistic dietary risk assessment of pesticide residues in foods for the
German population based on food monitoring data from 2009 to 2014, J Expo Sci Environ Epidemiol, 2018
Jan; 28(1):46–54, doi: 10.1038/jes.2017.7, Epub 2017 Jul 5, PubMed PMID: 28677675.

374. Silva JG, Boareto AC, Schreiber AK, Redivo DD, Gambeta E, Vergara F, Morais H, Zanoveli JM, Dalsenter PR,
Chlorpyrifos induces anxiety-like behaviour in offspring rats exposed during pregnancy, Neurosci Lett, 2017
Feb 22; 641:94–100, doi: 10.1016/j.neulet.2017.01.053, Epub 2017 Jan 24, PubMed PMID: 28130185.

375. Silver MK, Shao J, Ji C, Zhu B, Xu L, Li M, Chen M, Xia Y, Kaciroti N, Lozoff B, Meeker JD, Prenatal
organophosphate insecticide exposure and infant sensory function, Int J Hyg Environ Health, 2018 Apr;
221(3):469–78, doi: 10.1016/j.ijheh.2018.01.010, PubMed PMID: 29402694; PubMed Central PMCID:
PMC5902422.

376. Silver MK, Shao J, Zhu B, Chen M, Xia Y, Kaciroti N, Lozoff B, Meeker JD, Prenatal naled and chlorpyrifos
exposure is associated with deficits in infant motor function in a cohort of Chinese infants, Environ Int, 2017
Sep; 106:248–56, doi: 10.1016/j.envint.2017.05.015, Epub 2017 Jun 8, PubMed PMID: 28602489; PubMed
Central PMCID: PMC5533622.

377. Sindi RA, Harris W, Arnott G, Flaskos J, Lloyd Mills C, Hargreaves AJ, Chlorpyrifos- and chlorpyrifos oxon-
induced neurite retraction in pre-differentiated N2a cells is associated with transient hyperphosphorylation of
neurofilament heavy chain and ERK 1/2, Toxicol Appl Pharmacol, 2016 Oct 1; 308:20–31, doi:
10.1016/j.taap.2016.08.008, Epub 2016 Aug 10, PubMed PMID: 27521977.

378. Singh N, Lawana V, Luo J, Phong P, Abdalla A, Palanisamy B, Rokad D, Sarkar S, Jin H, Anantharam V,
Kanthasamy AG, Kanthasamy A, Organophosphate pesticide chlorpyrifos impairs STAT1 signaling to induce
dopaminergic neurotoxicity: Implications for mitochondria mediated oxidative stress signaling events,
Neurobiol Dis, 2018 Sep; 117:82–113, doi: 10.1016/j.nbd.2018.05.019, Epub 2018 May 31, PubMed PMID:
29859868; PubMed Central PMCID: PMC6108448.

379. Singh S, Gupta R, Kumari M, Sharma S, Nontarget effects of chemical pesticides and biological pesticide on
rhizospheric microbial community structure and function in Vigna racceptable daily intakeata, Environ Sci
60 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Pollut Res Int, 2015 Aug; 22(15):11290–300, doi: 10.1007/s11356-015-4341-x, Epub 2015 Mar 24, PubMed
PMID: 25801369.

380. Singh S, Gupta R, Sharma S, Effects of chemical and biological pesticides on plant growth parameters and
rhizospheric bacterial community structure in Vigna racceptable daily intakeata, J Hazard Mater, 2015 Jun 30;
291:102–10, doi: 10.1016/j.jhazmat.2015.02.053, Epub 2015 Feb 19, PubMed PMID: 25791643.

381. Singh S, Prakash A, Kaur S, Ming LC, Mani V, Majeed AB, The role of multifunctional drug therapy as an
antidote to combat experimental subacute neurotoxicity induced by organophosphate pesticides, Environ
Toxicol, 2016 Aug; 31(8):1017–26, doi: 10.1002/tox.22111, Epub 2015 Apr 10, PubMed PMID: 25864908.

382. Singleton ST, Lein PJ, Dadson OA, McGarrigle BP, Farahat FM, Farahat T, Bonner MR, Fenske RA, Galvin K,
Lasarev MR, Anger WK, Rohlman DS, Olson JR, Longitudinal assessment of occupational exposures to the
organophosphorous insecticides chlorpyrifos and profenofos in Egyptian cotton field workers, Int J Hyg
Environ Health, 2015 Mar; 218(2):203–11, doi: 10.1016/j.ijheh.2014.10.005, Epub 2014 Nov 7, PubMed PMID:
25466362; PubMed Central PMCID: PMC4323691.

383. Skaljac M, Kirfel P, Grotmann J, Vilcinskas A, Fitness costs of infection with Serratia symbiotica are associated
with greater susceptibility to insecticides in the pea aphid Acyrthosiphon pisum, Pest Manag Sci, 2018 Aug;
74(8):1829–36, doi: 10.1002/ps.4881, Epub 2018 Mar 7, PubMed PMID: 29443436.

384. Skovgaard M, Encinas SR, Jensen OC, Andersen JH, Condarco G, Jørs E, Pesticide Residues in Commercial
Lettuce, Onion, and Potato Samples From Bolivia-A Threat to Public Health? Environ Health Insights, 2017
Apr 18; 11:1178630217704194, doi: 10.1177/1178630217704194, eCollection 2017, Review, PubMed PMID:
28469451; PubMed Central PMCID: PMC5400016.

385. Slotkin TA, Seidler FJ, Prenatal nicotine alters the developmental neurotoxicity of postnatal chlorpyrifos
directed toward cholinergic systems: better, worse, or just "different?", Brain Res Bull, 2015 Jan; 110:54–67,
doi: 10.1016/j.brainresbull.2014.12.003, Epub 2014 Dec 12, PubMed PMID: 25510202; PubMed Central
PMCID: PMC4297693.

386. Slotkin TA, Skavicus S, Card J, Levin ED, Seidler FJ, Diverse neurotoxicants target the differentiation of
embryonic neural stem cells into neuronal and glial phenotypes, Toxicology, 2016 Nov 30; 372:42–51, doi:
10.1016/j.tox.2016.10.015, Epub 2016 Nov 2, PubMed PMID: 27816694; PubMed Central PMCID:
PMC5137195.

387. Slotkin TA, Skavicus S, Levin ED, Seidler FJ, Prenatal nicotine changes the response to postnatal chlorpyrifos:
Interactions targeting serotonergic synaptic function and cognition, Brain Res Bull, 2015 Feb;
111:84–96, doi: 10.1016/j.brainresbull.2015.01.003, Epub 2015 Jan 12, PubMed PMID: 25592617; PubMed
Central PMCID: PMC4324014.

388. Slotkin TA, Skavicus S, Seidler FJ, Prenatal drug exposures sensitize noradrenergic circuits to subsequent
disruption by chlorpyrifos, Toxicology, 2015 Dec 2; 338:8–16, doi: 10.1016/j.tox.2015.09.005, Epub 2015 Sep
28, PubMed PMID:26419632; PubMed Central PMCID: PMC4658258.

389. Smida A, Ncibi S, Taleb J, Ben Saad A, Ncib S, Zourgui L, Immunoprotective activity and antioxidant properties
of cactus (Opuntia ficus indica) extract against chlorpyrifos toxicity in rats, Biomed Pharmacother, 2017 Apr;
88:844–51, doi: 10.1016/j.biopha.2017.01.105, Epub 2017 Feb 4, PubMed PMID: 28167451.

390. Smith JN, Carver ZA, Weber TJ, Timchalk C, Predicting Transport of 3,5,6-Trichloro-2-Pyridinol Into Saliva
Using a Combination Experimental and Computational Approach, Toxicol Sci, 2017 Jun 1; 157(2):438–50,
doi:10.1093/toxsci/kfx055, PubMed PMID: 28402492.
APPENDIX 1 61

391. Sogorb MA, Fuster E, Del Río E, Estévez J, Vilanova E, Effects of mipafox, paraoxon, chlorpyrifos and its
metabolite chlorpyrifos-oxon on the expression of biomarker genes of differentiation in D3 mouse embryonic
stem cells, Chem Biol Interact, 2016 Nov 25; 259(Pt B):368–73, doi: 10.1016/j.cbi.2016.04.017, Epub 2016
Apr 23, PubMed PMID: 27117976.

392. Sogorb MA, Pamies D, Estevan C, Estévez J, Vilanova E, Roles of NTE protein and encoding gene in
development and neurodevelopmental toxicity, Chem Biol Interact, 2016 Nov 25; 259(Pt B):352–7, doi:
10.1016/j.cbi.2016.07.030, Epub 2016 Jul 28, Review, PubMed PMID: 27475862.

393. Solé M, Rivera-Ingraham G, Freitas R, The use of carboxylesterases as biomarkers of pesticide exposure in
bivalves: A methodological approach, Comp Biochem Physiol C Toxicol Pharmacol, 2018 Oct; 212:18–24, doi:
10.1016/j.cbpc.2018.06.002, Epub 2018 Jun 11, PubMed PMID: 29902568.

394. Soliman MI, Abd El-Halim AS, Mikhail MW, Efficacy of certain insecticides against rat flea (Xenopsylla cheopis)
among rodent species in Cairo Governorate, Egypt, J Egypt Soc Parasitol, 2015 Apr; 45(1):1–6, PubMed
PMID: 26012213.

395. Sotomayor V, Chiriotto TS, Pechen AM, Venturino A, Biochemical biomarkers of sublethal effects in Rhinella
arenarum late gastrula exposed to the organophosphate chlorpyrifos, Pestic Biochem Physiol, 2015 Mar;
119:48–53, doi: 10.1016/j.pestbp.2015.02.006, Epub 2015 Feb 16, PubMed PMID: 25868816.

396. Spaggiari D, Daali Y, Rudaz S, An extensive cocktail approach for rapid risk assessment of in vitro CYP450
direct reversible inhibition by xenobiotic exposure, Toxicol Appl Pharmacol, 2016 Jul 1; 302:41–51, doi:
10.1016/j.taap.2016.04.013, Epub 2016 Apr 20, PubMed PMID: 27105555.

397. Spodniewska A, Barski D, Giżejewska A, Effect of enrofloxacin and chlorpyrifos on the levels of vitamins A and
E in Wistar rats, Environ Toxicol Pharmacol, 2015 Sep; 40(2):587–91, doi: 10.1016/j.etap.2015.08.019, Epub
2015 Aug 19, PubMed PMID: 26356388.

398. Spodniewska A, Barski D, Concentration of hepatic vitamins A and E in rats exposed to chlorpyrifos and/or
enrofloxacin, Pol J Vet Sci, 2016; 19(2):371–8, doi: 10.1515/pjvs-2016-0046, PubMed PMID: 27487512.

399. Struciński P, Ludwicki JK, Góralczyk K, Czaja K, Hernik A, Liszewska M, Risk assessment for pesticides' MRL
non-compliances in Poland in the years 2011–15, Rocz Panstw Zakl Hig, 2015; 66(4):309–17, PubMed PMID:
26656412.

400. Su C, Niu P, Low doses of single or combined agrichemicals induces α-synuclein aggregation in nigrostriatal
system of mice through inhibition of proteasomal and autophagic pathways, Int J Clin Exp Med, 2015 Nov 15;
8(11):20508–15, eCollection 2015. PubMed PMID: 26884967; PubMed Central PMCID: PMC4723812.

401. Suemizu H, Kawai K, Murayama N, Nakamura M, Yamazaki H. Chimeric mice with humanized liver as a model
for testing organophosphate and carbamate pesticide exposure. Pest Manag Sci, 2018 Jun; 74(6):1424–30.
doi: 10.1002/ps.4825, Epub 2018 Feb 22, PubMed PMID: 29235720; PubMed Central PMCID: PMC5969263.

402. Suke SG, Sherekar P, Kahale V, Patil S, Mundhada D, Nanoti VM, Ameliorative effect of nanoencapsulated
flavonoid against chlorpyrifos-induced hepatic oxidative damage and immunotoxicity in Wistar rats, J Biochem
Mol Toxicol, 2018 May; 32(5):e22050, doi: 10.1002/jbt.22050, Epub 2018 Apr 18, PubMed PMID: 29667781.

403. Sumon KA, Rico A, Ter Horst MMS, Van den Brink PJ, Haque MM, Rashid H, Risk assessment of pesticides
used in rice-prawn concurrent systems in Bangladesh, Sci Total Environ, 2016 Oct 15; 568:498–506, doi:
10.1016/j.scitotenv.2016.06.014, Epub 2016 Jun 18, PubMed PMID: 27328394.
62 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

404. Sumon KA, Saha S, van den Brink PJ, Peeters ET, Bosma RH, Rashid H, Acute toxicity of chlorpyrifos to
embryo and larvae of banded gourami Trichogaster fasciata, J Environ Sci Health B, 2017 Feb; 52(2):92–98,
doi: 10.1080/03601234.2016.1239979, Epub 2016 Oct 24, PubMed PMID: 28099091.

405. Sun L, Xu W, Peng T, Chen H, Ren L, Tan H, Xiao D, Qian H, Fu Z, Developmental exposure of zebrafish
larvae to organophosphate flame retardants causes neurotoxicity, Neurotoxicol Teratol, 2016 May-Jun; 55:16–
22, doi: 10.1016/j.ntt.2016.03.003, Epub 2016 Mar 24, PubMed PMID: 27018022.

406. Tam NT, Berg H, Van Cong N, Evaluation of the joint toxicity of chlorpyrifos ethyl and fenobucarb on climbing
perch (Anabas testudineus) from rice fields in the Mekong Delta, Vietnam, Environ Sci Pollut Res Int, 2018
May; 25(14):13226–34, doi: 10.1007/s11356-016-6980-y, Epub 2016 Jun 2, PubMed PMID: 27250094.

407. Tam NT, Berg H, Van Cong N, The combined effect of Bassa 50EC and Vitashield 40EC on the brain
acetylcholinesterase activity in climbing perch (Anabas testudineus), Environ Sci Pollut Res Int, 2018 Jun;
25(17):17207–15, doi: 10.1007/s11356-018-2112-1, Epub 2018 Apr 30, PubMed PMID: 29713976; PubMed
Central PMCID: PMC6015109.

408. Tang X, Yang Y, Huang W, McBride MB, Guo J, Tao R, Dai Y, Transformation of chlorpyrifos in integrated
recirculating constructed wetlands (IRCWs) as revealed by compound-specific stable isotope (CSIA) and
microbial community structure analysis, Bioresour Techno, 2017 Jun; 233:264–70, doi:
10.1016/j.biortech.2017.02.077, Epub 2017 Feb 20, PubMed PMID: 28285217.

409. Tanvir EM, Afroz R, Chowdhury M, Gan SH, Karim N, Islam MN, Khalil MI, A model of chlorpyrifos distribution
and its biochemical effects on the liver and kidneys of rats, Hum Exp Toxicol, 2016 Sep; 35(9):991–1004, doi:
10.1177/0960327115614384, Epub 2015 Oct 30, PubMed PMID: 26519480.

410. Tascone O, Fillâtre Y, Roy C, Meierhenrich UJ, Behaviour of Multiclass Pesticide Residue Concentrations
during the Transformation from Rose Petals to Rose Absolute, J Agric Food Chem, 2015 May 27;
63(20):4922–32, doi: 10.1021/acs.jafc.5b00985, Epub 2015 May 19, PubMed PMID: 25942486.

411. Thaimuangphol W, Kasamesiri P, Lethal effects of pesticides on the nauplii fairy shrimp, Branchinella
thailandensis, Commun Agric Appl Biol Sci, 2015; 80(3):385–92, PubMed PMID: 27141736.

412. Thakur S, Dhiman M, Mantha AK, APE1 modulates cellular responses to organophosphate pesticide-induced
oxidative damage in non-small cell lung carcinoma A549 cells, Mol Cell Biochem, 2018 Apr; 441(1–2):201–16,
doi: 10.1007/s11010-017-3186-7, Epub 2017 Sep 8, PubMed PMID: 28887667.

413. Tian J, Dai H, Deng Y, Zhang J, Li Y, Zhou J, Zhao M, Zhao M, Zhang C, Zhang Y, Wang P, Bing G, Zhao L,
The effect of HMGB1 on sub-toxic chlorpyrifos exposure-induced neuroinflammation in amygdala of neonatal
rats, Toxicology, 2015 Dec 2; 338:95–103, doi: 10.1016/j.tox.2015.10.010, Epub 2015 Oct 30, PubMed PMID:
26524701.

414. Tiethof AK, Richardson JR, Hart RP, Knockdown of Butyrylcholinesterase but Not Inhibition by Chlorpyrifos
Alters Early Differentiation Mechanisms in Human Neural Stem Cells, Toxics, 2018 Sep 1; 6(3), pii: E52, doi:
10.3390/toxics6030052, PubMed PMID: 30200437; PubMed Central PMCID: PMC6160911.

415. Tong Z, Duan J, Wu Y, Liu Q, He Q, Shi Y, Yu L, Cao H, Evaluation of Highly Detectable Pesticides Sprayed in
Brassica napus L.: Degradation Behaviour and Risk Assessment for Honeybees, Molecules, 2018 Sep 27;
23(10), pii: E2482, doi: 10.3390/molecules23102482, PubMed PMID: 30262759.
APPENDIX 1 63

416. Tosi S, Costa C, Vesco U, Quaglia G, Guido G, A 3-year survey of Italian honey bee-collected pollen reveals
widespread contamination by agricultural pesticides, Sci Total Environ, 2018 Feb 15; 615:208–18, doi:
10.1016/j.scitotenv.2017.09.226, Epub 2017 Sep 29, PubMed PMID: 28968582.

417. Trinder M, McDowell TW, Daisley BA, Ali SN, Leong HS, Sumarah MW, Reid G, Probiotic Lactobacillus
rhamnosus Reduces Organophosphate Pesticide Absorption and Toxicity to Drosophila melanogaster, Appl
Environ Microbiol, 2016 Sep 30; 82(20):6204–13, Print 2016 Oct 15, PubMed PMID: 27520820; PubMed
Central PMCID: PMC5068162.

418. Tsaboula A, Papadakis EN, Vryzas Z, Kotopoulou A, Kintzikoglou K, Papadopoulou-Mourkidou E,


Environmental and human risk hierarchy of pesticides: A prioritization method, based on monitoring, hazard
assessment and environmental fate, Environ Int, 2016 May; 91:78–93, doi: 10.1016/j.envint.2016.02.008,
Epub 2016 Feb 22, PubMed PMID: 26915710.

419. Tussellino M, Ronca R, Carotenuto R, Pallotta MM, Furia M, Capriglione T, Chlorpyrifos exposure affects fgf8,
sox9, and bmp4 expression required for cranial neural crest morphogenesis and chondrogenesis in Xenopus
laevis embryos, Environ Mol Mutagen, 2016 Oct; 57(8):630–40, doi: 10.1002/em.22057, Epub 2016 Sep 27,
PubMed PMID: 27669663.

420. Tüzün N, Debecker S, Op de Beeck L, Stoks R, Urbanisation shapes behavioural responses to a pesticide,
Aquat Toxicol, 2015 Jun; 163:81–8, doi: 10.1016/j.aquatox.2015.04.002, Epub 2015 Apr 2, PubMed PMID:
25863029.

421. Uchendu C, Ambali SF, Ayo JO, Esievo KA, The protective role of alpha-lipoic acid on long-term exposure of
rats to the combination of chlorpyrifos and deltamethrin pesticides, Toxicol Ind Health, 2017 Feb; 33(2):159–
70, doi: 10.1177/0748233715616553, Epub 2016 Jul 10, PubMed PMID: 26697883

422. Uchendu C, Ambali SF, Ayo JO, Esievo KAN, Chronic co-exposure to chlorpyrifos and deltamethrin pesticides
induces alterations in serum lipids and oxidative stress in Wistar rats: mitigating role of alpha-lipoic acid,
Environ Sci Pollut Res Int, 2018 May 7, doi: 10.1007/s11356-018-2185-x, PubMed PMID: 29736639.

423. Uniyal S, Sharma RK, Technological advancement in electrochemical biosensor based detection of
Organophosphate pesticide chlorpyrifos in the environment: A review of status and prospects, Biosens
Bioelectron, 2018 Sep 30; 116:37–50, doi: 10.1016/j.bios.2018.05.039, Epub 2018 May 26, Review, PubMed
PMID: 29857260.

424. Urlacher E, Monchanin C, Rivière C, Richard FJ, Lombardi C, Michelsen-Heath S, Hageman KJ, Mercer AR,
Measurements of Chlorpyrifos Levels in Forager Bees and Comparison with Levels that Disrupt Honey Bee
Odor-Mediated Learning Under Laboratory Conditions, J Chem Ecol, 2016 Feb; 42(2):127–38, doi:
10.1007/s10886-016-0672-4, Epub 2016 Feb 12, PubMed PMID: 26872472.

425. Van den Brink PJ, Klein SL, Rico A, Interaction between stress induced by competition, predation, and an
insecticide on the response of aquatic invertebrates, Environ Toxicol Chem, 2017 Sep; 36(9):2485–92, doi:
10.1002/etc.3788, Epub 2017 Apr 4, PubMed PMID: 28295548.

426. van Herk WG, Vernon RS, Waterer DR, Tolman JH, Lafontaine PJ, Prasad RP, Field Evaluation of Insecticides
for Control of Cabbage Maggot (Diptera: Anthomyiidae) in Rutabaga in Canada, J Econ Entomol, 2017 Feb 1;
110(1):177–85, doi: 10.1093/jee/tow238, PubMed PMID: 28031474.

427. van Wendel de Joode B, Mora AM, Lindh CH, Hernández-Bonilla D, Córdoba L, Wesseling C, Hoppin JA,
Mergler D, Pesticide exposure and neurodevelopment in children aged 6-9 years from Talamanca, Costa Rica,
64 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Cortex, 2016 Dec; 85:137–50, doi: 10.1016/j.cortex.2016.09.003, Epub 2016 Sep 15, PubMed PMID:
27773359.

428. Vassallo A, Chiappalone M, De Camargos Lopes R, Scelfo B, Novellino A, Defranchi E, Palosaari T, Weisschu
T, Ramirez T, Martinoia S, Johnstone AFM, Mack CM, Landsiedel R, Whelan M, Bal-Price A, Shafer TJ, A
multi-laboratory evaluation of microelectrode array-based measurements of neural network activity for acute
neurotoxicity testing, Neurotoxicology, 2017 May; 60:280–92, doi: 10.1016/j.neuro.2016.03.019, Epub 2016
Mar 29, PubMed PMID: 27036093.

429. Venerosi A, Tait S, Stecca L, Chiarotti F, De Felice A, Cometa MF, Volpe MT, Calamandrei G, Ricceri L, Effects
of maternal chlorpyrifos diet on social investigation and brain neuroendocrine markers in the offspring—a
mouse study, Environ Health, 2015 Apr 2; 14:32, doi: 10.1186/s12940-015-0019-6, PubMed PMID: 25889763;
PubMed Central PMCID: PMC4448273.

430. Ventura C, Nieto MR, Bourguignon N, Lux-Lantos V, Rodriguez H, Cao G, Randi A, Cocca C, Núñez M,
Pesticide chlorpyrifos acts as an endocrine disruptor in adult rats causing changes in mammary gland and
hormonal balance, J Steroid Biochem Mol Biol, 2016 Feb; 156:1–9, doi: 10.1016/j.jsbmb.2015.10.010, Epub
2015 Oct 27, PubMed PMID: 26518068.

431. Ventura C, Venturino A, Miret N, Randi A, Rivera E, Núñez M, Cocca C, Chlorpyrifos inhibits cell proliferation
through ERK1/2 phosphorylation in breast cancer cell lines, Chemosphere, 2015 Feb; 120:343–50, doi:
10.1016/j.chemosphere.2014.07.088, Epub 2014 Aug 31, PubMed PMID: 25180937.

432. Ventura C, Zappia CD, Lasagna M, Pavicic W, Richard S, Bolzan AD, Monczor F, Núñez M, Cocca C, Effects
of the pesticide chlorpyrifos on breast cancer disease. Implication of epigenetic mechanisms, J Steroid
Biochem Mol Biol, 2018 Oct 2, pii: S0960-0760(18)30187-0, doi: 10.1016/j.jsbmb.2018.09.021, PubMed PMID:
30290214.

433. Villa S, Di Nica V, Pescatore T, Bellamoli F, Miari F, Finizio A, Lencioni V, Comparison of the behavioural
effects of pharmaceuticals and pesticides on Diamesa zernyi larvae (Chironomidae), Environ Pollut, 2018 Jul;
238:130–9, doi: 10.1016/j.envpol.2018.03.029, Epub 2018 Mar 20, PubMed PMID: 29554561.

434. Vivan LM, Torres JB, Fernandes PL, Activity of Selected Formulated Biorational and Synthetic Insecticides
Against Larvae of Helicoverpa armigera (Lepidoptera: Noctuidae), J Econ Entomol, 2017 Feb 1; 110(1):118–
26, doi: 10.1093/jee/tow244, PubMed PMID: 28011685.

435. Wang DM, Zhang BX, Liu XM, Rao XJ, Li SG, Li MY, Liu S, Molecular characterization of two
acetylcholinesterase genes from the rice leaffolder, cnaphalocrocis medinalis (lepidoptera: pyralidae), Arch
Insect Biochem Physiol, 2016 Nov; 93(3):129–42, doi: 10.1002/arch.21347, Epub 2016 Jul 22, PubMed PMID:
27447944.

436. Wang L, Espinoza HM, MacDonald JW, Bammler TK, Williams CR, Yeh A, Louie KW,Marcinek DJ, Gallagher
EP, Olfactory Transcriptional Analysis of Salmon Exposed to Mixtures of Chlorpyrifos and Malathion Reveal
Novel Molecular Pathways of Neurobehavioural Injury, Toxicol Sci, 2016 Jan; 149(1):145–57,
doi:10.1093/toxsci/kfv223, Epub 2015 Oct 22, PubMed PMID: 26494550; PubMed Central PMCID:
PMC4731405.

437. Wang L, Liu Z, Zhang J, Wu Y, Sun H, Chlorpyrifos exposure in farmers and urban adults: Metabolic
characteristic, exposure estimation, and potential effect of oxidative damage, Environ Res, 2016 Aug;
149:164–70, doi: 10.1016/j.envres.2016.05.011, Epub 2016 May 18, PubMed PMID: 27208467.
APPENDIX 1 65

438. Wang P, Dai H, Zhang C, Tian J, Deng Y, Zhao M, Zhao M, Bing G, Zhao L, Evaluation of the effects of
chlorpyrifos combined with lipopolysaccharide stress on neuroinflammation and spatial memory in neonatal
rats, Toxicology, 2018 Dec 1; 410:106–15, doi: 10.1016/j.tox.2018.09.008, Epub 2018 Sep 17, PubMed PMID:
30236991.

439. Wang P, Rashid M, Liu J, Hu M, Zhong G, Identification of multi-insecticide residues using GC-NPD and the
degradation kinetics of chlorpyrifos in sweet corn and soils, Food Chem, 2016 Dec 1; 212:420–6, doi:
10.1016/j.foodchem.2016.05.008, Epub 2016 May 2, PubMed PMID: 27374551.

440. Wang P, Wang J, Sun YJ, Yang L, Wu YJ, Cadmium and chlorpyrifos inhibit cellular immune response in
spleen of rats, Environ Toxicol, 2017 Jul; 32(7):1927–36, doi: 10.1002/tox.22415, Epub 2017 Mar 15, PubMed
PMID: 28296077.

441. Wang Q, Li C, Zheng R, Que X, Phytoremediation of chlorpyrifos in aqueous system by riverine macrophyte,
Acorus calamus: toxicity and removal rate, Environ Sci Pollut Res Int, 2016 Aug; 23(16):16241–8, doi:
10.1007/s11356-016-6673-6, Epub 2016 May 7, PubMed PMID: 27154841.

442. Wang RL, Zhu-Salzman K, Baerson SR, Xin XW, Li J, Su YJ, Zeng RS, Identification of a novel cytochrome
P450 CYP321B1 gene from tobacco cutworm (Spodoptera litura) and RNA interference to evaluate its role in
commonly used insecticides, Insect Sci, 2017 Apr; 24(2):235–47, doi: 10.1111/1744-7917.12315, Epub 2016
May 13, PubMed PMID: 26782704.

443. Wang X, Xiang X, Yu H, Liu S, Yin Y, Cui P, Wu Y, Yang J, Jiang C, Yang Q, Monitoring and biochemical
characterization of beta-cypermethrin resistance in Spodoptera exigua (Lepidoptera: Noctuidae) in Sichuan
Province, China, Pestic Biochem Physiol, 2018 Apr; 146:71–9, doi: 10.1016/j.pestbp.2018.02.008, Epub 2018
Mar 2, PubMed PMID: 29626995.

444. Wanwimolruk S, Kanchanamayoon O, Boonpangrak S, Prachayasittikul V, Food safety in Thailand 1: it is safe


to eat watermelon and durian in Thailand, Environ Health Prev Med, 2015 May; 20(3):204–15, doi:
10.1007/s12199-015-0452-8, Epub 2015 Feb 20, PubMed PMID: 25697579; PubMed Central PMCID:
PMC4434230.

445. Wanwimolruk S, Kanchanamayoon O, Phopin K, Prachayasittikul V, Food safety in Thailand 2: Pesticide


residues found in Chinese kale (Brassica oleracea), a commonly consumed vegetable in Asian countries, Sci
Total Environ, 2015 Nov 1; 532:447–55, doi: 10.1016/j.scitotenv.2015.04.114, Epub 2015 Jun 18, PubMed
PMID: 26093223.

446. Weber TJ, Smith JN, Carver ZA, Timchalk C, Non-invasive saliva human biomonitoring: development of an in
vitro platform, J Expo Sci Environ Epidemiol, 2017 Jan; 27(1):72–7, doi: 10.1038/jes.2015.74, Epub 2015 Nov
11, PubMed PMID: 26555474.

447. Wei W, Wang J, Tian CB, Du SW, Wu KC, A highly hydrolytically stable lanthanide organic framework as a
sensitive luminescent probe for DBP and chlorpyrifos detection, Analyst, 2018 Nov 5; 143(22):5481–6, doi:
10.1039/c8an01606b, PubMed PMID: 30289144.

448. Weichert FG, Floeter C, Meza Artmann AS, Kammann U, Assessing the ecotoxicity of potentially neurotoxic
substances—Evaluation of a behavioural parameter in the embryogenesis of Danio rerio, Chemosphere, 2017
Nov; 186:43–50, doi: 10.1016/j.chemosphere.2017.07.136, Epub 2017 Jul 27, PubMed PMID: 28772184.

449. Williman C, Munitz MS, Montti MIT, Medina MB, Navarro AF, Ronco AE, Pesticide survey in water and
suspended solids from the Uruguay River Basin, Argentina, Environ Monit Assess, 2017 Jun; 189(6):259, doi:
10.1007/s10661-017-5956-9, Epub 2017 May 8, PubMed PMID: 28484956.
66 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

450. Wolf J, Potrich M, Lozano ER, Gouvea A, Pegorini CS, Combined physical and chemical methods to control
lesser mealworm beetles under laboratory conditions, Poult Sci, 2015 Jun; 94(6):1145–9, doi:
10.3382/ps/pev093, Epub 2015 Mar 31, PubMed PMID: 25834245.

451. Woodley SK, Mattes BM, Yates EK, Relyea RA, Exposure to sublethal concentrations of a pesticide or
predator cues induces changes in brain architecture in larval amphibians, Oecologia, 2015 Nov; 179(3):655–
65, doi: 10.1007/s00442-015-3386-3, Epub 2015 Jul 14, PubMed PMID: 26169394.

452. Wu X, Davie-Martin CL, Steinlin C, Hageman KJ, Cullen NJ, Bogdal C, Understanding and Predicting the Fate
of Semivolatile Organic Pesticides in a Glacier-Fed Lake Using a Multimedia Chemical Fate Model, Environ
Sci Technol, 2017 Oct 17; 51(20):11752–60, doi: 10.1021/acs.est.7b03483, Epub 2017 Oct 6, PubMed PMID:
28925251.

453. Wu X, Yang X, Majumder A, Swetenburg R, Goodfellow FT, Bartlett MG, Stice SL, From the Cover:
AstrocytesAre Protective Against Chlorpyrifos Developmental Neurotoxicity in Human Pluripotent Stem Cell-
Derived Astrocyte-Neuron Cocultures, Toxicol Sci, 2017 Jun 1; 157(2):410–20, doi: 10.1093/toxsci/kfx056,
PubMed PMID: 28369648.

454. Xia X, Sun B, Gurr GM, Vasseur L, Xue M, You M, Gut Microbiota Mediate Insecticide Resistance in the
Diamondback Moth, Plutella xylostella (L.), Front Microbiol, 2018 Jan 23; 9:25, doi: 10.3389/fmicb.2018.00025,
eCollection 2018, PubMed PMID: 29410659; PubMed Central PMCID: PMC5787075.

455. Xiao P, Liu F, Liu Y, Yao S, Zhu G, Effects of Pesticide Mixtures on Zooplankton Assemblages in Aquatic
Microcosms Simulating Rice Paddy Fields, Bull Environ Contam Toxicol, 2017 Jul; 99(1):27–32, doi:
10.1007/s00128-017-2105-7, Epub 2017 May 15, PubMed PMID: 28508304.

456. Xu DM, Wang YH, Wang N, Rao GW, Effects of single and co-exposure of Cu and chlorpyrifos on the toxicity
of earthworm, Huan Jing Ke Xue, 2015 Jan; 36(1):280–5, Chinese, PubMed PMID: 25898676.

457. Xu MY, Sun YJ, Wang P, Xu HY, Chen LP, Zhu L, Wu YJ, Metabolomics analysisnand biomarker identification
for brains of rats exposed subchronically to thenmixtures of low-dose cadmium and chlorpyrifos, Chem Res
Toxicol, 2015 Junn15; 28(6):1216–23, doi: 10.1021/acs.chemrestox.5b00054, Epub 2015 Apr 22,
PubMednPMID: 25856237.

458. Xu MY, Wang P, Sun YJ, Wu YJ, Metabolomic analysis for combined hepatotoxicity of chlorpyrifos and
cadmium in rats, Toxicology, 2017 Jun 1; 384:50–8, doi: 10.1016/j.tox.2017.04.008, Epub 2017 Apr 19,
PubMed PMID:28433638.

459. Xu MY, Wang P, Sun YJ, Yang L, Wu YJ, Joint toxicity of chlorpyrifos and cadmium on the oxidative stress and
mitochondrial damage in neuronal cells, Food Chem Toxicol, 2017 May; 103:246–52, doi:
10.1016/j.fct.2017.03.013, Epub 2017 Mar 9, PubMed PMID: 28286310.

460. Xu TF, Peng YK, Li YY, Zhai C, Zheng XC, Qiao L, A Point Raman Scanning Method for Rapidly Detection of
Spinach Quality and Safety Parameters, Guang Pu Xue Yu Guang Pu Fen Xi, 2016 Jun; 36(6):1765–70,
Chinese, PubMed PMID: 30052388.

461. Xu ZB, Zou XP, Zhang N, Feng QL, Zheng SC, Detoxification of insecticides, allechemicals and heavy metals
by glutathione S-transferase SlGSTE1 in the gut of Spodoptera litura, Insect Sci, 2015 Aug; 22(4):503–11, doi:
10.1111/1744-7917.12142, Epub 2014 Nov 6, PubMed PMID: 24863567.
APPENDIX 1 67

462. Yahia D, Ali MF, Assessment of neurohepatic DNA damage in male Sprague-Dawley rats exposed to
organophosphates and pyrethroid insecticides, Environ Sci Pollut Res Int, 2018 Jun; 25(16):15616–29, doi:
10.1007/s11356-018-1776-x, Epub 2018 Mar 23, PubMed PMID: 29572745.

463. Yan DK, Hu M, Tang YX, Fan JQ, Proteomic Analysis Reveals Resistance Mechanism Against Chlorpyrifos in
Frankliniella occidentalis (Thysanoptera: Thripidae), J Econ Entomol, 2015 Aug; 108(4):2000–8, doi:
10.1093/jee/tov139, Epub 2015 May 29, PubMed PMID: 26470346.

464. Yang BJ, Liu ML, Zhang YX, Liu ZW, Effects of temperature on fitness costs in chlorpyrifos-resistant brown
planthopper, Nilaparvata lugens (Hemiptera: Delphacidae), Insect Sci, 2018 Jun; 25(3):409–17, doi:
10.1111/1744-7917.12432, Epub 2017 Feb 16, PubMed PMID: 28026125.

465. Yang C, Yu H, Jiang H, Qiao C, Liu R, An engineered microorganism can simultaneously detoxify cadmium,
chlorpyrifos, and γ-hexachlorocyclohexane, J Basic Microbiol, 2016 Jul; 56(7):820–6, doi:
10.1002/jobm.201500559, Epub 2015 Dec 9, PubMed PMID: 26648050.

466. Yang G, Chen C, Wang Y, Cai L, Kong X, Qian Y, Wang Q, Joint toxicity of chlorpyrifos, atrazine, and cadmium
at lethal concentrations to the earthworm Eisenia fetida, Environ Sci Pollut Res Int, 2015 Jun; 22(12):9307–15,
doi: 10.1007/s11356-015-4097-3, Epub 2015 Jan 18, PubMed PMID: 25595933.

467. Yang G, Chen C, Yu Y, Zhao H, Wang W, Wang Y, Cai L, He Y, Wang X, Combined effects of four pesticides
and heavy metal chromium (Ⅵ) on the earthworm using avoidance behaviour as an endpoint, Ecotoxicol
Environ Saf, 2018 Aug 15; 157:191–200, doi: 10.1016/j.ecoenv.2018.03.067, Epub 2018 Apr 3, PubMed PMID:
29621711.

468. Yang T, Feng S, Lu Y, Yin C, Wang J, Dual-template magnetic molecularly imprinted particles with multi-hollow
structure for the detection of dicofol and chlorpyrifos-methyl, J Sep Sci, 2016 Jun; 39(12):2388–95,
doi:10.1002/jssc.201600258, Epub 2016 May 24, PubMed PMID: 27119595.

469. Yang X, Naughton SX, Han Z, He M, Zheng YG, Terry AV Jr, Bartlett MG, Mass Spectrometric Quantitation of
Tubulin Acetylation from Pepsin-Digested Rat Brain Tissue Using a Novel Stable-Isotope Standard and
Capture by Anti-Peptide Antibody (SISCAPA) Method, Anal Chem, 2018 Feb 6; 90(3):2155–63, doi:
10.1021/acs.analchem.7b04484, Epub 2018 Jan 24, PubMed PMID: 29320166.

470. Yang X, Wu X, Brown KA, Le T, Stice SL, Bartlett MG, Determination of chlorpyrifos and its metabolites in cells
and culture media by liquid chromatography-electrospray ionization tandem mass spectrometry, J Chromatogr
B Analyt Technol Biomed Life Sci, 2017 Sep 15; 1063:112–7, doi: 10.1016/j.jchromb.2017.08.010, Epub 2017
Aug 18, PubMed PMID: 28858752.

471. Yang XQ, Gene expression analysis and enzyme assay reveal a potential role of the carboxylesterase gene
CpCE-1 from Cydia pomonella in detoxification of insecticides, Pestic Biochem Physiol, 2016 May; 129:56–62,
doi: 10.1016/j.pestbp.2015.10.018, Epub 2015 Oct 23, PubMed PMID: 27017882.

472. Yang Y, Kong W, Zhao L, Xiao Q, Liu H, Zhao X, Yang M, A multiresidue method for simultaneous
determination of 44 organophosphorous pesticides in Pogostemon cablin and related products using modified
QuEChERS sample preparation procedure and GC-FPD, J Chromatogr B Analyt Technol Biomed Life Sci,
2015 Jan 1; 974:118–25, doi: 10.1016/j.jchromb.2014.10.023, Epub 2014 Oct 25, PubMed PMID: 25463206.

473. Yang Y, Wang C, Xu H, Lu Z, Sublethal effects of four insecticides on folding and spinning behaviour in the rice
leaffolder, Cnaphalocrocis medinalis (Guenée) (Lepidoptera: Pyralidae), Pest Manag Sci, 2018 Mar;
74(3):658–64, doi: 10.1002/ps.4753, Epub 2017 Dec 1, PubMed PMID: 28984412.
68 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

474. Yao J, Zhu YC, Adamczyk J, Luttrell R, Influences of acephate and mixtures with other commonly used
pesticides on honey bee (Apis mellifera) survival and detoxification enzyme activities, Comp Biochem Physiol
C Toxicol Pharmacol, 2018 Jul; 209:9–17, doi: 10.1016/j.cbpc.2018.03.005, Epub 2018 Mar 18, PubMed
PMID: 29563044.

475. Yao Q, Xu S, Dong Y, Que Y, Quan L, Chen B, Characterization of Vitellogenin and Vitellogenin Receptor of
Conopomorpha sinensis Bradley and Their Responses to Sublethal Concentrations of Insecticide, Front
Physiol, 2018 Sep 11; 9:1250, doi: 10.3389/fphys.2018.01250, eCollection 2018, PubMed PMID: 30279662;
PubMed Central PMCID: PMC6154279.

476. Yazdinezhad A, Abbasian M, Hojjat Hosseini S, Naserzadeh P, Agh-Atabay AH, Hosseini MJ, Protective effects
of Ziziphora tenuior extract against chlorpyrifos induced liver and lung toxicity in rat: Mechanistic approaches
in subchronic study, Environ Toxicol, 2017 Sep; 32(9):2191–202, doi: 10.1002/tox.22432, Epub 2017 Jun 1,
PubMed PMID: 28569040.

477. Ye YH, Li C, Yang J, Ma L, Xiao Y, Hu J, Rajput NA, Gao CF, Zhang YY, Wang MH, Construction of an
immobilised acetylcholinesterase column and its application in screening insecticidal constituents from
Magnolia officinalis, Pest Manag Sci, 2015 Apr; 71(4):607–15, doi: 10.1002/ps.3908, Epub 2014 Oct 20,
PubMed PMID: 25228142.

478. Youssif BGM, Abdelrahman MH, Abdelazeem AH, Abdelgawad MA, Ibrahim HM, Salem OIA, Mohamed MFA,
Treambleau L, Bukhari SNA, Design, synthesis, mechanistic and histopathological studies of small-molecules
of novel indole-2-carboxamides and pyrazino[1,2-a]indol-1(2H)-ones as potential anticancer agents effecting
the reactive oxygen species production, Eur J Med Chem, 2018 Feb 25; 146:260–73, doi:
10.1016/j.ejmech.2018.01.042, Epub 2018 Jan 31, PubMed PMID: 29407956.

479. Yu RX, Wang YH, Hu XQ, Wu SG, Cai LM, Zhao XP, Individual and Joint Acute Toxicities of Selected
Insecticides Against Bombyx mori (Lepidoptera: Bombycidae), J Econ Entomol, 2016 Feb; 109(1):327–33, doi:
10.1093/jee/tov316, Epub 2015 Nov 6, PubMed PMID: 26546487.

480. Yuan H, Li G, Yang L, Yan X, Yang D, Development of melamine-formaldehyde resin microcapsules with low
formaldehyde emission suited for seed treatment, Colloids Surf B Biointerfaces, 2015 Apr 1; 128:149–54, doi:
10.1016/j.colsurfb.2015.02.029, Epub 2015 Feb 23, PubMed PMID: 25734968.

481. Zainudin BH, Salleh S, Mohamed R, Yap KC, Muhamad H, Development, validation and determination of
multiclass pesticide residues in cocoa beans using gas chromatography and liquid chromatography tandem
mass spectrometry, Food Chem, 2015 Apr 1; 172:585–95, doi: 10.1016/j.foodchem.2014.09.123, Epub 2014
Sep 28, PubMed PMID: 25442595.

482. Zarei MH, Soodi M, Qasemian-Lemraski M, Jafarzadeh E, Taha MF, Study of the chlorpyrifos neurotoxicity
using neural differentiation of acceptable daily intakepose tissue-derived stem cells, Environ Toxicol, 2016
Nov; 31(11):1510–9, doi: 10.1002/tox.22155, Epub 2015 May 27, PubMed PMID: 26018426.

483. Zhai C, Xu TF, Peng YK, Li YY, Detection of Chlorpyrifos on Spinach Based on Surface Enhanced Raman
Spectroscopy with Silver Colloids, Guang Pu Xue Yu Guang Pu Fen Xi, 2016 Sep; 36(9):2835–40, Chinese,
PubMed PMID: 30084609.

484. Zhang J, Dai H, Deng Y, Tian J, Zhang C, Hu Z, Bing G, Zhao L, Neonatal chlorpyrifos exposure induces loss
of dopaminergic neurons in young adult rats, Toxicology, 2015 Oct 2; 336:17–25, doi:
10.1016/j.tox.2015.07.014, Epub 2015 Jul 26, PubMed PMID: 26215101.
APPENDIX 1 69

485. Zhang LJ, Jing YP, Li XH, Li CW, Bourguet D, Wu G, Temperature-sensitive fitness cost of insecticide
resistance in Chinese populations of the diamondback moth Plutella xylostella, Mol Ecol, 2015 Apr;
24(7):1611–27, doi: 10.1111/mec.13133, PubMed PMID: 25732547.

486. Zhang N, Liu J, Chen SN, Huang LH, Feng QL, Zheng SC, Expression profiles of glutathione S-transferase
superfamily in Spodoptera litura tolerated to sublethal doses of chlorpyrifos, Insect Sci, 2016 Oct; 23(5):675–
87, doi: 10.1111/1744-7917.12202, Epub 2015 May 21, Erratum in: Insect Sci, 2016 Dec; 23 (6):924, PubMed
PMID: 25641855.

487. Zhang R, Wu W, Zhang Z, Park Y, He L, Xing B, McClements DJ, Effect of the Composition and Structure of
Excipient Emulsion on the Bioaccessibility of Pesticide Residue in Agricultural Products, J Agric Food Chem,
2017 Oct 18; 65(41):9128–38, doi: 10.1021/acs.jafc.7b02607, Epub 2017 Oct 4, PubMed PMID: 28914056.

488. Zhang X, Liao X, Mao K, Zhang K, Wan H, Li J, Insecticide resistance monitoring and correlation analysis of
insecticides in field populations of the brown planthopper Nilaparvata lugens (stål) in China 2012–14, Pestic
Biochem Physiol, 2016 Sep; 132:13–20, doi: 10.1016/j.pestbp.2015.10.003, Epub 2015 Oct 21, PubMed
PMID: 27521908.

489. Zhang Y, Han Y, Yang Q, Wang L, He P, Liu Z, Li Z, Guo H, Fang J, Resistance to cycloxaprid in Laodelphax
striatellus is associated with altered expression of nicotinic acetylcholine receptor subunits, Pest Manag Sci,
2018 Apr; 74(4):837–43, doi: 10.1002/ps.4757, Epub 2017 Dec 14, PubMed PMID: 28991400.

490. Zhang Y, Yang B, Li J, Liu M, Liu Z, Point mutations in acetylcholinesterase 1 associated with chlorpyrifos
resistance in the brown planthopper, Nilaparvata lugens Stål, Insect Mol Biol, 2017 Aug; 26(4):453–60, doi:
10.1111/imb.12309, Epub 2017 Apr 13, PubMed PMID: 28407384.

491. Zhao J, Chen B, Species sensitivity distribution for chlorpyrifos to aquatic organisms: Model choice and
sample size, Ecotoxicol Environ Saf, 2016 Mar; 125:161–9, doi: 10.1016/j.ecoenv.2015.11.039, Epub 2015
Dec 14, PubMed PMID: 26701839.

492. Zhao L, Xie F, Wang TT, Liu MY, Li JL, Shang L, Deng ZX, Zhao PX, Ma XM, Chlorpyrifos Induces the
Expression of the Epstein-Barr Virus Lytic Cycle Activator BZLF-1 via Reactive Oxygen Species, Oxid Med
Cell Longev, 2015; 2015:309125, doi: 10.1155/2015/309125, Epub 2015 Jul 14, PubMed PMID: 26257840;
PubMed Central PMCID: PMC4516845.

493. Zhao Y, Zhang Y, Wang G, Han R, Xie X, Effects of chlorpyrifos on the gut microbiome and urine metabolome
in mouse (Mus musculus), Chemosphere, 2016 Jun; 153:287–93, doi: 10.1016/j.chemosphere.2016.03.055,
Epub 2016 Mar 26, PubMed PMID: 27018521.

494. Zhou XW, Zhao XH, Susceptibility of nine organophosphorus pesticides in skimmed milk towards inoculated
lactic acid bacteria and yogurt starters, J Sci Food Agric, 2015 Jan; 95(2):260–6, doi: 10.1002/jsfa.6710, Epub
2014 May 21, PubMed PMID: 24777955.

495. Zhu J, Dubois A, Ge Y, Olson JA, Ren X, Application of human haploid cell genetic screening model in
identifying the genes required for resistance to environmental toxicants: Chlorpyrifos as a case study, J
Pharmacol Toxicol Methods, 2015 Nov-Dec; 76:76–82, doi: 10.1016/j.vascn.2015.08.154, Epub 2015 Aug 20,
PubMed PMID: 26299976; PubMed Central PMCID: PMC4831576.

496. Zhu J, Li Y, Jiang H, Liu C, Lu W, Dai W, Xu J, Liu F, Selective toxicity of the mesoionic insecticide,
triflumezopyrim, to rice planthoppers and beneficial arthropods, Ecotoxicology, 2018 May; 27(4):411–9, doi:
10.1007/s10646-018-1904-x, Epub 2018 Feb 5, PubMed PMID: 29404868.
70 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

497. Zuo YY, Huang JL, Wang J, Feng Y, Han TT, Wu YD, Yang YH, Knockout of a P-glycoprotein gene increases
susceptibility to abamectin and emamectin benzoate in Spodoptera exigua, Insect Mol Biol, 2018 Feb;
27(1):36–45, doi:10.1111/imb.12338, Epub 2017 Jul 28, PubMed PMID: 28753233.

498. Zurlinden TJ, Reisfeld B, A Novel Method for the Development of Environmental Public Health Indicators and
Benchmark Dose Estimation Using a Health-Based End Point for Chlorpyrifos, Environ Health Perspect, 2018
Apr 20; 126(4):047009, doi:10.1289/EHP1743, Erratum in: Environ Health Perspect, 2018 Sep; 126(9):99001,
PubMed PMID: 29681141; PubMed Central PMCID: PMC6071752.

Zurlinden TJ, Reisfeld B, Erratum: A Novel Method for the Development of Environmental Public Health Indicators
and Benchmark Dose Estimation Using a Health-Based Endpoint for Chlorpyrifos, Environ Health Perspect,
2018 Sep; 126(9):99001, doi: 10.1289/EHP4279, PubMed PMID: 30187771.
APPENDIX 2 71

APPENDIX 2: STUDY ABSTRACTS AND STUDY EVALUATIONS


Summary
Study and published abstract Study evaluation
No.

1 Environ Toxicol Chem. 2018 Jul; 37(7):1898–906. doi: 10.1002/etc.4139. Epub 2018 May 7. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Comparative in vitro and in vivo effects of chlorpyrifos oxon in the outbred CD-1 mouse (Mus additional useful data on low dose chlorpyrifos effects.
musculus) and great plains toad (anaxyrus cognatus).
Anderson T, Liu J, McMurry S, Pope C.
We compared biochemical, functional, and behavioural responses to the organophosphorus
anticholinesterase chlorpyrifos oxon (CPO) in mice (mus musculus, CD-1) and toads (anaxyrus
cognatus, Great Plains toad). Toads were substantially less sensitive to acute lethality of CPO
based on the maximum tolerated (nonlethal) dose (toads, 77 mg/kg; mice, 5.9 mg/kg). Sublethal
exposures led to classical signs of toxicity (increased involuntary movements, autonomic secretions)
in mice but hypoactivity in toads. Motor performance in an inclined plane test was not affected by
CPO in mice but was altered at the highest dosage in toads. Acetylcholinesterase (AChE),
butyrylcholinesterase, monoacylglycerol lipase, and fatty acid amide hydrolase activities in brain
were inhibited in mice but not in toads, and fatty acid amide hydrolase activity in the liver was
inhibited in both species. Toad brain AChE was less sensitive to in vitro inhibition by CPO (50%
inhibitory concentration [IC50; 20 min, 37 °C], 101 vs 7.8 nM; IC50 [20 min, 26 °C], 149 vs 6.2 nM),
and studies of inhibitor kinetics indicated substantially lower anticholinesterase potency of CPO
against the toad brain enzyme. Using an in vitro indirect inhibition assay, preincubation of CPO with
toad brain homogenate was more effective than an equivalent mouse brain homogenate at reducing
CPO potency. These data suggest that the relatively low sensitivity of toads to cholinergic toxicity is
based on the low sensitivity of brain AChE, which in turn may be attributable to more effective
target-site detoxification. Environ Toxicol Chem 2018; 37:1898-1906.
2 Food Chem Toxicol. 2018 Aug; 118:42–52. doi: 10.1016/j.fct.2018.04.065. Epub 2018 May 3. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Postnatal chlorpyrifos exposure and apolipoprotein E (APOE) genotype differentially affect additional useful data on low dose chlorpyrifos effects.
cholinergic expression and developmental parameters in transgenic mice.
Basaure P, Guardia-Escote L, Cabré M, Peris-Sampedro F, Sánchez-Santed F, Domingo JL,
Colomina MT.
Chlorpyrifos (CPF) is one of the most commonly used organophosphate pesticides in the world. Our
previous results described that apolipoprotein E (APOE) polymorphisms are a source of individual
differences in susceptibility to CPF. The aim of this study was to assess the physical and
72 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

biochemical effects of postnatal exposure to CPF in the apoE targeted replacement mouse model.
Mice were exposed to CPF at 0 or 1 mg/kg/day from postnatal day 10–15. Physical development,
plasma and forebrain cholinesterase (ChE) activity and gene expression in liver and forebrain were
evaluated. CPF exposure delays physical maturation and decreases the expression of choline
acetyltransferase, α4-subunit and the α7 receptor. CPF decreases the expression of vesicular
acetylcholine transporter (VAChT) mRNA in the forebrain only in apoE3 mice. The expression of
paraoxonase-2 in the forebrain was also influenced by APOE genotype and CPF. Differences
between genotypes were observed in litter size, ChE activity, expression of butyrylcholinesterase
and paraoxonase-1 in liver and variants of acetylcholinesterase, VAChT and the α7 receptor in the
forebrain. These results support that there are different vulnerabilities to postnatal CPF exposure
according to the APOE polymorphism, which in turn affects the cholinergic system and defenses to
oxidative stress.
3 Behav Brain Res. 2017 Feb 1; 318:1–11. doi: 10.1016/j.bbr.2016.10.014. Epub 2016 Oct 11. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Two cholinesterase inhibitors trigger dissimilar effects on behaviour and body weight in C57BL/6 additional useful data on low dose chlorpyrifos effects.
mice:
The case of chlorpyrifos and rivastigmine.
Basaure P, Peris-Sampedro F, Cabré M, Reverte I, Colomina MT.
Cholinesterases (ChE) are common targets of organophosphate (OP) pesticides and play a critical
role in the pathology of some dementias. While chlorpyrifos (CPF) remains one of the most
commonly used OPs in the world, numerous investigations have reported its neurotoxic potential
and highlighted behavioural disturbances upon its administration. Rivastigmine currently serves to
treat Alzheimer's disease, but it may induce cholinergic overstimulation in non-demented individuals.
The present investigation aimed to compare the acute and delayed effects caused by both ChE
inhibitors in adult C57BL/6 male mice. The animals were daily fed either a standard, a CPF- (5mg/kg
body weight) or a rivastigmine-supplemented diet (1 or 2mg/kg body weight) for 8 weeks. After the
treatment, we established an 8-week washout period to assess recovery. ChE enzyme activity,
biomarkers, physical effects, and behavioural alterations were evaluated at different time points
during the exposure and after the washout period. Both rivastigmine doses induced a time-
dependent weight increase. CPF and rivastigmine inhibited brain acetylcholinesterase following an
isoform-specific pattern. As for behavioural assessment, CPF negatively modulated learning
strategies and impaired memory in a Barnes maze task at the end of the exposure.
On the other hand, the low dose of rivastigmine improved memory recall at the end of the washout
APPENDIX 2 73

Summary
Study and published abstract Study evaluation
No.

period in a Morris water maze. Indeed, our results endorse the positive effects of low doses of
rivastigmine following a drug-free period in young mice. Therefore, doses and periodicity of
treatment to improve cognition in elderly people upon rivastigmine administration should be revised.
4 J Environ Sci Health B. 2017 Feb; 52(2):77–83. doi: 10.1080/03601234.2016.1239973. Epub 2016 Chlorpyrifos dose used exceeds the in vivo NOAELs for
Oct 24. cholinesterase inhibition in mice. The study provides no
additional useful data on low dose chlorpyrifos effects.
Mother gestational exposure to organophosphorus pesticide induces neuron and glia loss in
daughter adult brain.
Chen XP, Chao YS, Chen WZ, Dong JY.
Chlorpyrifos (CPF) is a widely used organophosphorus pesticide with developmental neurotoxicity
such as morphogenesis toxicity. In the present study, we assessed the effects of prenatal CPF
exposure on systemic parameters and cytoarchitecture of medial prefrontal cortex (mPFC) in
adulthood. Gestational dams were exposed to 5mg/kg/d of CPF during gestational days 13–17,
while body weight, organ coefficient, and neuron and glia counts of offspring were determined on
postnatal day 60. Our results showed that CPF treatment induced little or no effects on body weight
and organ coefficients. There were also no significant pathological changes in mPFC. However,
neuron and glia count analysis showed that CPF treatment reduced neuron and glia counts in
anterior cingulate, prelimbic, and infralimbic areas of mPFC. The CPF react pattern was similar in
both sexes, and there was no statistical difference in most of the sub-regions. Thus, our results
revealed an embryonic origin brain deficit induced by gestational mother pesticide exposure.
5 Toxicol Res (Camb). 2016 Jun 14; 5(5):1359–70. doi: 10.1039/c5tx00282f. eCollection 2016 Sep 1. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
An in vivo study in mice: mother's gestational exposure to organophosphorus pesticide retards the additional useful data on low dose chlorpyrifos effects.
division and migration process of neural progenitors in the fetal developing brain.
Chen XP, Wang TT, Wu XZ, Wang DW, Chao YS.
Background: Widely utilized pesticides such as chlorpyrifos (CPF) can cause cognitive
abnormalities, neurotransmitter disruptions and brain cytoarchitecture deficits in adulthood due to
exposure in the prenatal period, but the mechanism underlying the development and maintenance of
such neurotoxicity in embryonic neurogenesis remains largely unclear. Using embryonic neocortex
slices, we investigated mitosis population constituents and characteristic interkinetic nuclear
migration (INM) to evaluate the CPF effects on the proliferation process of neural progenitors.
Methods: Gestational days (GD) 14 and GD 7.5–11.5 ICR dams were exposed to 5 mg kg–1 of CPF
to investigate immediate toxicity and sustained toxicity. Proliferating nuclei were labeled with 50 mg
74 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

kg-1 of Brdu at 1, 3, 6 and 9 hours before samples were collected. The mitoses count and Brdu
positive nuclei (BPN) location were measured and analyzed in standard sections of the embryonic
dorsolateral cortex. Results: CPF reduced the mitoses count in the primary progenitors but not in
the secondary progenitors which are time sustained. CPF retarded BPN migration with a 6–9 μm
delay of the relative location in the immediate groups and a 3–6 μm delay in the sustained ones.
CPF had no or little effects on the global mitoses count and BPN count. Conclusion: Prenatal CPF
exposure disrupts the proliferation process of primary progenitors in the embryonic dorsolateral
cortex immediately and with sustained effects, which may contribute to explain the toxicity
mechanism in early neurogenesis.
6 J Neuroinflammation. 2016 Jun 14; 13(1):149. doi: 10.1186/s12974-016-0617-4. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Prenatal exposure to the organophosphate insecticide chlorpyrifos enhances brain oxidative stress additional useful data on low dose chlorpyrifos effects.
and prostaglandin E2 synthesis in a mouse model of idiopathic autism.
The dose (6 mg/kg bw/d) was selected on the basis of
De Felice A, Greco A, Calamandrei G, Minghetti L. previous multidose studies because it was effective at
BACKGROUND: Autism spectrum disorders (ASD) are emerging as polygenic and multifactorial producing behavioural changes in the absence of overt toxic
disorders in which complex interactions between defective genes and early exposure to symptoms in dams or major effects on pregnancy length,
environmental stressors impact on the correct neurodevelopment and brain processes. number of pups at delivery, sex ratio, pups weight at delivery
Organophosphate insecticides, among which chlorpyrifos (CPF), are widely diffused environmental and growing rate.
toxicants associated with neurobehavioural deficits and increased risk of ASD occurrence in
children. Oxidative stress and dysregulated immune responses are implicated in both
organophosphate neurodevelopmental effects and ASD etiopathogenesis. BTBR T+tf/J mice, a well-
studied model of idiopathic autism, show several behavioural and immunological alterations found in
ASD children, and we recently showed that CPF gestational exposure strengthened some of these
autistic-like traits. In the present study, we aimed at investigating whether the behavioural effects of
gestational CPF administration are associated with brain increased oxidative stress and altered lipid
mediator profile. METHODS: Brain levels of F2-isoprostanes (15-F2t-IsoP), as index of in vivo
oxidative stress, and prostaglandin E2 (PGE2), a major arachidonic acid metabolite released by
immune cells and by specific glutamatergic neuron populations mainly in cortex and hippocampus,
were assessed by specific enzyme-immuno assays in brain homogenates from BTBR T+tf/J and
C57Bl6/J mice, exposed during gestation to either vehicle or CPF. Measures were performed in
mice of both sexes, at different postnatal stages (PNDs 1, 21, and 70). RESULTS: At birth, BTBR
T+tf/J mice exhibited higher baseline
15-F2t-IsoP levels as compared to C57Bl6/J mice, suggestive of greater oxidative stress processes.
Gestational treatment with CPF-enhanced 15-F2t-IsoP and PGE2 levels in strain-and age-
APPENDIX 2 75

Summary
Study and published abstract Study evaluation
No.

dependent manner, with 15-F2t-IsoP increased in BTBR T+tf/J mice at PNDs 1 and 21, and PGE2
elevated in BTBR T+tf/J mice at PNDs 21 and 70. At PND 21, CPF effects were sex-dependent
being the increase of the two metabolites mainly associated with male mice. CPF treatment also
induced a reduction of somatic growth, which reached statistical significance at PND 21.
CONCLUSIONS: These findings indicate that the autistic-like BTBR T+tf/J strain is highly vulnerable
to environmental stressors during gestational period. The results further support the hypothesis that
oxidative stress might be the link between environmental neurotoxicants such as CPF and ASD. The
increased levels of oxidative stress during early postnatal life could result in delayed and long-
lasting alterations in specific pathways relevant to ASD, of which PGE2 signaling represents an
important one.
7 PLoS One. 2015 Mar 24; 10(3):e0121663. doi: 10.1371/journal.pone.0121663. eCollection 2015. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Prenatal exposure to a common organophosphate insecticide delays motor development in a mouse additional useful data on low dose chlorpyrifos effects.
model of idiopathic autism.
The dose (6 mg/kg bw/d) was selected on the basis of
De Felice A, Scattoni ML, Ricceri L, Calamandrei G. previous multidose studies because it was effective at
Autism spectrum disorders are characterized by impaired social and communicative skills and producing behavioural changes in the absence of overt toxic
repetitive behaviours. Emerging evidence supported the hypothesis that these neurodevelopmental symptoms in dams or major effects on pregnancy length,
disorders may result from a combination of genetic susceptibility and exposure to environmental number of pups at delivery, sex ratio, pups weight at delivery
toxins in early developmental phases. This study assessed the effects of prenatal exposure to and growing rate.
chlorpyrifos (CPF), a widely diffused organophosphate insecticide endowed with developmental
neurotoxicity at sub-toxic doses, in the BTBR T+tf/J mouse strain, a validated model of idiopathic
autism that displays several behavioural traits relevant to the autism spectrum. To this aim, pregnant
BTBR mice were administered from gestational day 14 to 17 with either vehicle or CPF at a dose of
6 mg/kg/bw by oral gavages. Offspring of both sexes underwent assessment of early developmental
milestones, including somatic growth, motor behaviour and ultrasound vocalization. To evaluate the
potential long-term effects of CPF, two different social behaviour patterns typically altered in the
BTBR strain (free social interaction with a same-sex companion in females, or interaction with a
sexually receptive female in males) were also examined in the two sexes at adulthood. Our findings
indicate significant effects of CPF on somatic growth and neonatal motor patterns. CPF treated pups
showed reduced weight gain, delayed motor maturation (ie, persistency of immature patterns such
as pivoting at the expenses of coordinated locomotion) and a trend to enhanced ultrasound
vocalization. At adulthood, CPF associated alterations were found in males only: the altered pattern
of investigation of a sexual partner, previously described in BTBR mice, was enhanced in CPF
males, and associated to increased ultrasonic vocalization rate. These findings strengthen the need
76 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

of future studies
to evaluate the role of environmental chemicals in the etiology of neurodevelopment disorders.
8 Pestic Biochem Physiol. 2018 Jan; 144:36–41. doi: 10.1016/j.pestbp.2017.11.002. Epub 2017 Nov Study used chlorpyrifos-ethyl and provides no additional
10. useful data on low dose chlorpyrifos neurological effects.
Chlorpyrifos-induced parkinsonian model in mice: Behaviour, histopathology and biochemistry.
Deveci HA, Karapehlivan M.
INTRODUCTION: The aim of this study was to investigate the protective effect of caffeic acid
phenethyl ester (CAPE) on Paraoxonase (PON1) activity, and levels of lipid profile, total sialic acid
(TSA), total antioxidant capacity (TAC) and total oxidant capacity (TOC) in the plasma and brain
tissue of mice with chlorpyrifos-ethyl (CPF)-induced Parkinson. MATERIAL AND METHOD: In the
study, 35 male Swiss albino mice were divided into 5 groups including equal number of mice as
follows; intraperitoneal injection of saline for mice in control (C) group, subcutaneous injection of
80mg/kg CPF for CPF group, intraperitoneal injection of 10μmol/kg CAPE for CAPE group,
subcutaneous injection of 80mg/kg CPF and intraperitoneal injection of 10μmol/kg CAPE for
CPF+CAPE group and intraperitoneal injection of 10% ethanol diluted in physiological saline
solution for 21days for ethanol (E) group. All the mice were fed with normal feed and tap water ad
libitum. At the end of the study, PON1 activity, lipid profile (except for brain), and TSA, TAC and TOC
levels in the plasma and brain tissue were analyzed. Tissue samples of brain substantia nigra were
evaluated histopathologically. RESULTS: Levels of plasma TAC, high density lipoprotein (HDL) and
PON1 activity were statistically lower in CPF group than the other groups (P<0.001). Also, levels of
plasma TOC, TSA, total cholesterol, triglycerides, low density lipoprotein (LDL) and very low density
lipoprotein (VLDL) were statistically higher in CPF group than the other groups(P<0.001). PON1
activity and level of TAC were significantly lower in brain tissue of CPF groups (P<0.001). In
addition, TOC and TSA levels were significantly higher in brain tissue in CPF group (P<0.001).
CONCLUSION: In conclusion, CAPE showed a protective effect on PON1 activity and levels of lipid
profile, TSA, TAC and TOC in plasma and brain tissue and prevented the neurodegenerations in
brain tissue in CPF-induced Parkinson's disease.
9 Food Chem Toxicol. 2018 Dec; 122:1–10. doi: 10.1016/j.fct.2018.09.069. Epub 2018 Sep 29. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Postnatal exposure to chlorpyrifos produces long-term effects on spatial memory and the cholinergic additional useful data on low dose chlorpyrifos effects.
system in mice in a sex- and APOE genotype-dependent manner.
Guardia-Escote L, Basaure P, Blanco J, Cabré M, Pérez-Fernández C, Sánchez-Santed F, Domingo
APPENDIX 2 77

Summary
Study and published abstract Study evaluation
No.

JL, Colomina MT.


Organophosphorus pesticides—and in particular chlorpyrifos (CPF)—are extensively used
worldwide.
They mainly exert their toxicity by targeting the cholinergic system. Several studies suggested that
the gene coding for apolipoprotein E (apoE), which is a risk factor for several diseases, can also
confer different vulnerability to toxic insults. This study was aimed at assessing the long-term effects
of postnatal exposure to CPF on learning and memory as well as the expression levels of several
genes involved in cholinergic neurotransmission in mice. Both male and female apoE4-TR and
C57BL/6 mice were exposed to either 0 or 1 mg/kg/day of CPF by oral gavage using a micropipette
on postnatal days 10–15. At 9 months, they were tested in a Morris Water Maze (MWM) and the
gene expression in the frontal cortex and hippocampus was evaluated. Our results show that, in
males, CPF had an effect on the spatial retention, while in females, it altered the expression levels
of nicotinic receptors. Furthermore, apoE4-TR mice performed the worst during the MWM retention
and presented low expression levels in a considerable number of cholinergic genes. Taken together,
the current results reveal long-term effects in mice nine months after postnatal exposure to CPF,
which are modulated by sex and apoE4 genotype.
10 Food Chem Toxicol. 2018 Aug; 118:821–9. doi: 10.1016/j.fct.2018.06.051. Epub 2018 Jun 21. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Inhibition of cholinergic and non-cholinergic targets following subacute exposure to chlorpyrifos in additional useful data on low dose chlorpyrifos effects.
normal and high fat fed male C57BL/6J mice.
Howell GE 3rd, Kondakala S, Holdridge J, Lee JH, Ross MK.
The effects of obesity on organophosphate pesticide-mediated toxicities, including both cholinergic
and non-cholinergic targets, have not been fully elucidated. Therefore, the present study was
designed to determine if high fat diet intake alters the effects of repeated exposure to chlorpyrifos
(CPS) on the activities of both cholinergic and noncholinergic serine hydrolase targets. Male
C57BL/6J mice were placed on either standard rodent chow or high fat diet for four weeks with CPS
exposure (2.0 mg/kg) for the last 10 days of diet intake. Exposure to CPS did not alter
acetylcholinesterase in the central nervous system, but it did significantly inhibit circulating
cholinesterase activities in both diet groups. CPS significantly inhibited hepatic carboxylesterase
and fatty acid amide hydrolase and this inhibition was significantly greater in high fat fed animals.
Additionally, CPS exposure and high fat diet intake downregulated genes involved in hepatic de
novo lipogenesis as well as cytochrome P450 enzymes involved in hepatic xenobiotic metabolism.
In summary, the present study demonstrates that high fat diet intake potentiates CPS mediated
78 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

inhibition of both carboxylesterase and fatty acid amide hydrolase in the liver of obese animals
following subacute exposure and suggests obesity may be a risk factor for increased non-
cholinergic hepatic CPS toxicity.
11 Toxicol Appl Pharmacol. 2017 Dec 15; 337:67–75. doi: 10.1016/j.taap.2017.10.019. Epub 2017 Oct Chlorpyrifos dose used exceeds the in vivo NOAELs for
31. cholinesterase inhibition in mice. The study provides no
additional useful data on low dose chlorpyrifos effects.
Effects of acute exposure to chlorpyrifos on cholinergic and non-cholinergic targets in normal and
high-fat fed male C57BL/6J mice.
Kondakala S, Lee JH, Ross MK, Howell GE 3rd.
The prevalence of obesity is increasing at an alarming rate in the United States with 36.5% of adults
being classified as obese. Compared to normal individuals, obese individuals have noted
pathophysiological alterations which may alter the toxicokinetics of xenobiotics and therefore alter
their toxicities. However, the effects of obesity on the toxicity of many widely utilized pesticides has
not been established. Therefore, the present study was designed to determine if the obese
phenotype altered the toxicity of the most widely used organophosphate (OP) insecticide,
chlorpyrifos (CPS). Male C57BL/6J mice were fed normal or high-fat diet for 4weeks and
administered a single dose of vehicle or CPS (2.0mg/kg; oral gavage) to assess cholinergic
(acetylcholinesterase activities) and non-cholinergic (carboxylesterase and endocannabinoid
hydrolysis) endpoints. Exposure to CPS significantly decreased red blood cell acetylcholinesterase
(AChE) activity, but not brain AChE activity, in both diet groups. Further, CPS exposure decreased
hepatic carboxylesterase activity and hepatic hydrolysis of a major endocannabinoid, anandamide,
in a diet-dependent manner with high-fat diet fed animals being more sensitive to CPS-mediated
inhibition. These in vivo studies were corroborated by in vitro studies using rat primary hepatocytes,
which demonstrated that fatty acid amide hydrolase and CES activities were more sensitive to CPS-
mediated inhibition than 2-arachidonoylglycerol hydrolase activity. These data demonstrate hepatic
CES and FAAH activities in high-fat diet fed mice were more potently inhibited than those in normal
diet fed mice following CPS exposure, which suggests that the obese phenotype may exacerbate
some of the non-cholinergic effects of CPS exposure.
12 Toxicol Ind Health. 2016 Jul; 32(7):1179–96. doi: 10.1177/0748233714548207. Epub 2014 Sep 18. There are reporting inadequacies regarding the dose actually
used in the study.
Ameliorative effect of vitamin E to mouse dams and their pups following exposure of mothers to
chlorpyrifos during gestation and lactation periods. In the chemicals description in the paper, the following is
stated:
Mansour SA, Gamet-Payrastre L.
APPENDIX 2 79

Summary
Study and published abstract Study evaluation
No.

Pesticides are omnipresent in environment, water, fruits, and vegetables and are considered as risk ‘High-purity PESTANAL1 Analytical Standard of CPF (99.2%
factors for human health. Consumers are mainly exposed to pesticides through diet, and the main (was purchased from Fluka (Riedel-de Haën, France) and
question to be answered concerns the impact of such exposure on health. In this study, we added as a component of rodent nuggets at a dose of 50
developed a mouse model to mimic consumer exposure. During gestation and lactation periods, the mg/kg of food. This dose allowed the mice to ingest the
experimental mouse dams (M) received one of the following treatments: (a) diet-free of pesticides; equivalent of the acceptable daily intake for CPF (0.01 mg/kg
(b) diet enriched with chlorpyrifos (CPF; 44.0 μg kg(-1)); c) diet + oral vitamin E (vit. E; α-tocopherol; body weight (b.w.)/day), based on previous estimate of
200 mg/kg/mouse); and (d) diet enriched with CPF (44.0 μg/kg + oral vit. E (200 mg/kg/mouse). At average food consumption per mouse (Merhi et al, 2010).
weaning, pups (P) and dams were killed, and organs as well as blood samples were collected. The acceptable daily intake was as defined for humans by
Compared with control results, CPF induced alteration of measured parameters (eg organ weight, the Joint Food and Agriculture Organization/World Health
alkaline phosphatase, urea, malondialdehyde, superoxide dismutase, and cholinesterase) either in Organization Meeting on Pesticide Residues (JMPR, 2005)
mouse dams or in their offspring. Also, CPF induced histological impairment in kidney, liver, and and extrapolated to mice on the basis of mean b.w.
ovary. Administration of vit. E in conjunction with CPF clearly alleviated deviation of these Information on the toxicity of the pesticide was obtained from
parameters than those of control ones. In conclusion, a dietary exposure of mice during gestation Extoxnet (1996).’
and lactation to low dose of CPF led to significant changes in the mother but also in the weaned
animals that have not been directly exposed to this pesticide. These biological and histological In the experimental design description in the paper, the
modifications could be reversed by an oral supplementation of vit. E. following is stated:
‘All the 16 individually housed dams with their pups were
segregated into 4 different groups with 4 animals in each
group. During gestation and lactation periods, the
experimental mouse dams (M) received one of the following
treatments: (a) diet free of pesticides C (M); (b) diet enriched
with CPF; 44.0 µg/kg CPF (M); (c) diet free of pesticides +
oral vit. E (α-tocopherol); 200 mg/kg/mouse V (M); and (d)
diet enriched with CPF; 44.0 µg/kg + oral vit. E; 200 mg
/kg/mouse CPF/V (M). The dose of vit. E was divided into two
equal portions and given twice a week. Pups (P) of the
experimental dams were designated as C (P), CPF (P), V (P),
and CPF/V (P), to sign control, CPF, vit. E, and vit. E + CPF-
treated groups, respectively. At the end of weaning, blood
samples were taken from the facial artery of each animal
(mothers and pups).’
Critically, the actual level of food consumption in the study
was not measured. Accordingly the dosimetrics of the study
are considered to be unreliable. Adding weight to this
conclusion was the fact that plasma cholinesterase was
80 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

significantly (p < 0.05) inhibited in the chlorpyrifos only


treated animals (mean 2831 U/L) cf. the negative control
group (3581 U/L).
Overall the study is not regarded as being reliable.
Furthermore, the effects observed were associated with the
presence of inhibition of plasma cholinesterase. The study
has no value for evaluation of low dose chlorpyrifos effects.
13 ASN Neuro. 2016 Jun 30; 8(3). pii: 1759091416656253. doi: 10.1177/1759091416656253. Print Dosimetry: ‘To determine the effects of reduced reelin
2016 Jun. expression and CPO exposure on gross morphology, protein
expression, and dendritic spine distribution, C57Bl/6 female
A Complex Interaction Between Reduced Reelin Expression and Prenatal Organophosphate mice were crossed with heterozygous male reeler mice
Exposure Alters Neuronal Cell Morphology. (B6C3Fe-a/a-RelnRl/+) originally obtained from the Jackson
Mullen BR, Ross B, Chou JW, Khankan R, Khialeeva E, Bui K, Carpenter EM. Laboratory (Bar Harbor, ME, USA); these mice are referred to
as Rl+/− throughout the text. At gestational Day 13.5,
Genetic and environmental factors are both likely to contribute to neurodevelopmental disorders pregnant females were implanted with osmotic minipumps
including schizophrenia, autism spectrum disorders, and major depressive disorders. Prior studies loaded with 6 mg/ml (20 mg/kg) of CPO (or a vehicle control)
from our laboratory and others have demonstrated that the combinatorial effect of two factors- as previously described. This dose corresponds to that used
reduced expression of reelin protein and prenatal exposure to the organophosphate pesticide by other groups (Laviola et al, 2006) and is well below the
chlorpyrifos oxon-gives rise to acute biochemical effects and to morphological and behavioural reported LD50 of 60 mg/kg.’
phenotypes in adolescent and young adult mice. In the current study, we examine the consequences
of these factors on reelin protein expression and neuronal cell morphology in adult mice. While the The authors have also claimed that ‘20 mg/kg corresponds to
cell populations that express reelin in the adult brain appear unchanged in location and distribution, a moderate environmental exposure for humans.’ Such
the levels of full length and cleaved reelin protein show persistent reductions following prenatal claims are not correct, even for mixing, loacceptable daily
exposure to chlorpyrifos oxon. Cell positioning and organization in the hippocampus and cerebellum intakeng and application, under Australian conditions.
are largely normal in animals with either reduced reelin expression or prenatal exposure to Notably this dose is > 6600 times the current Australian
chlorpyrifos oxon, but cellular complexity and dendritic spine organization is altered, with a skewed acceptable daily intake of 0.003 mg/kg bw/day.
distribution of immature dendritic spines in adult animals. Paradoxically, combinatorial exposure to Chlorpyrifos dose used exceeds the in vivo NOAELs for
both factors appears to generate a rescue of the dendritic spine phenotypes, similar to the cholinesterase inhibition in mice. The study provides no
mitigation of behavioural and morphological changes observed in our prior study. Together, our additional useful data on low dose chlorpyrifos effects.
observations support an interaction between reelin expression and chlorpyrifos oxon exposure that
is not simply additive, suggesting a complex interplay between genetic and environmental factors in
regulating brain morphology.
14 Front Behav Neurosci. 2015 Apr 29; 9:110. doi: 10.3389/fnbeh.2015.00110. eCollection 2015. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
APPENDIX 2 81

Summary
Study and published abstract Study evaluation
No.

Strain dependent effects of conditioned fear in adult C57Bl/6 and Balb/C mice following postnatal additional useful data on low dose chlorpyrifos effects.
exposure to chlorpyrifos: relation to expression of brain acetylcholinesterase mRNA.
Oriel S, Kofman O.
Following reports of emotional psychopathology in children and adults exposed to
organophosphates, the effects of postnatal chlorpyrifos (CPF) on fear-conditioning and depression-
like behaviours were tested in adult mice. Concomitant changes in expression of mRNA for synaptic
and soluble splice variants of acetylcholinesterase (AChE) were examined in mouse pups and
adults of the Balb/C and C57Bl/6 (B6) strains, which differ in their behavioural and hormonal stress
response. Mice were injected subcutaneously with 1 mg/kg CPF on postnatal days 4–10 and tested
as adults for conditioned fear, sucrose preference, and forced swim. Acetylcholinesterase activity
was assessed in the brains of pups on the first and last day of treatment. Expression of soluble and
synaptic AChE mRNA was assessed in brains of treated pups and fear-conditioned adults using
real-time PCR. Adult Balb/C mice exposed postnatally to CPF showed exacerbated fear-conditioning
and impaired active avoidance. Adult B6 mice exposed postnatally to CPF showed a more specific
fear response to tones and less freezing in the inter-tone intervals, in contrast to the vehicle-
pretreated mice. Chlorpyrifos also attenuated sweet preference and enhanced climbing in the forced
swim test. Chlorpyrifos-treated mice had increased expression of both synaptic and readthrough
AChE transcripts in the hippocampus of Balb/C mice and decreased expression in the amygdala
following fear-conditioning. In conclusion, postnatal CPF had long-term effects on fear and
depression, as well as on expression of AChE mRNA. These changes may be related to alteration in
the interaction between hippocampus and amygdala in regulating negative emotions.
15 Int J Mol Sci. 2017 Nov 20; 18(11). pii: E2467. doi: 10.3390/ijms18112467. Chlorpyrifos produced a mixed pattern of up-and down-
regulation of genes, which was more evident in the groups
Specific Effects of Chronic Dietary Exposure to Chlorpyrifos on Brain Gene Expression-A Mouse treated with a higher chlorpyrifos concentration (10
Study. mg/kg/day). Effects on gene expression occurred at doses ≥
Pallotta MM, Ronca R, Carotenuto R, Porreca I, Turano M, Ambrosino C, Capriglione T. 1 mg/kg bw/day ie at doses higher than the thresholds for
inhibition of blood cholinesterases in mice. The majority of
Chlorpyrifos (CPF) is an organophosphate insecticide used to control pests on a variety of food and the down-regulated transcripts belonged to proteins involved
feed crops. In mammals, maternal exposure to CPF has been reported to induce cerebral cortex in synaptic transmission and plasticity, GABAergic and
thinning, alteration of long-term brain cognitive function, and Parkinson-like symptoms, but the dopaminergic signalling. The total number of gene networks
mechanisms of these processes are not fully understood. In this study, we aimed to gain a deeper affected sharply reduced at 8 months, and
understanding of the alterations induced in the brains of mice chronically exposed to CPF by dietary acetylcholinesterase values shifted from highly inhibited—
intake. For our purpose, we analysed F1 offspring (sacrificed at 3 and 8 months) of Mus musculus, approximately 80% in 3-month-old mice—to 30% in the eldest
treated in utero and postnatally with 3 different doses of CPF (0.1-1-10 mg/kg/day). Using RT² mice.
82 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

Profiler PCR Arrays, we evaluated the alterations in the expression of 84 genes associated with
neurodegenerative diseases. In the brains of exposed mice, we evidenced a clear dose-response
relationship for AChE inhibition and alterations of gene expression. Some of the genes that were
steacceptable daily intakely down-regulated, such as Pink1, Park 2, Sv2b, Gabbr2, Sept5 and
Atxn2, were directly related to Parkinson's onset. Our experimental results shed light
on the possibility that long-term CPF exposure may exert membrane signalling alterations which
make brain cells more susceptible to develop neurodegenerative diseases.
16 Physiol Behav. 2015 May 15; 144:37–45. doi: 10.1016/j.physbeh.2015.03.006. Epub 2015 Mar 5. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Chronic exposure to chlorpyrifos triggered body weight increase and memory impairment depending additional useful data on low dose chlorpyrifos effects.
on human apoE polymorphisms in a targeted replacement mouse model.
Peris-Sampedro F, Basaure P, Reverte I, Cabré M, Domingo JL, Colomina MT.
Despite restrictions on their use, humans are still constantly exposed to organophosphates (OPs). A
huge number of studies have ratified the neurotoxic effects of chlorpyrifos (CPF) and suggested its
association with neurodegenerative diseases, but data are still scarce. Human apolipoprotein E
(apoE) plays an important role in lipid transport and distribution. In humans, the apoE4 isoform has
been linked to an increased risk of Alzheimer's disease (AD). ApoE3 is the most prevalent isoform
worldwide, and has been often established as the healthful one. The current study, performed in
targeted replacement (TR) adult male mice, aimed to inquire whether genetic variations of the
human apoE respond differently to a chronic dietary challenge with CPF. At four/five months of age,
mice carrying apoE2, apoE3 or apoE4 were pair-fed a diet supplemented with CPF at 0 or 2mg/kg
body weight/day for 13weeks. Cholinergic signs were monitored daily and body weight changes
weekly. In the last week of treatment, learning and memory were assessed in a Barnes maze task.
Dietary CPF challenge increased body weight only in apoE3 mice. Differences in the acquisition and
retention of the Barnes maze were attributed to apoE genetic differences. Our results showed that
apoE4 mice performed worse than apoE2 and apoE3 carriers in the acquisition period of the spatial
task, and that apoE2 mice had poorer retention than the other two genotypes. On the other hand,
CPF increased the search velocity of apoE2 subjects during the acquisition period. Retention was
impaired only in CPF-exposed apoE3 mice. These results underline that gene×environment
interactions need to be taken into account in epidemiological studies. Given that apoE3, the most
common polymorphism in humans, has proved to be the most sensitive to CPF, the potential
implications for human health merit serious thought.
17 Arch Toxicol. 2018 May; 92(5):1717–28. doi: 10.1007/s00204-018-2174-3. Epub 2018 Feb 5. Chlorpyrifos dose used exceeds the in vivo NOAELs for
APPENDIX 2 83

Summary
Study and published abstract Study evaluation
No.

New mechanistic insights on the metabolic-disruptor role of chlorpyrifos in apoE mice: a focus on cholinesterase inhibition in mice. The study provides no
insulin- and leptin-signalling pathways. additional useful data on low dose chlorpyrifos effects.
Peris-Sampedro F, Blanco J, Cabré M, Basaure P, Guardia-Escote L, Domingo JL, Sánchez DJ,
Colomina MT.
Recently, we have provided evidence, suggesting that mice expressing the human apolipoprotein E3
(apoE3) are more prone to develop an obesity-like phenotype and a diabetic profile when
subchronically fed a chlorpyrifos (CPF)-supplemented diet. The aim of the current study was to
examine the underlying mechanisms through which CPF alters both insulin- and leptin-signalling
pathways in an APOE-dependent manner. Both adult apoE3- and E4-targeted replacement and
C57BL/6 mice were exposed to CPF at 0 or 2 mg/kg body weight/day through the diet for 8
consecutive weeks. We determined the expression of JAK2, p-JAK2, STAT3, p-STAT3, SOCS3, IRS-
1, p-IRS-1, AKT, p-AKT, GSK3β, p-GSK3β, and apoE in the liver, as well as hepatic mRNA levels of
pon1, pon2, and pon3. CPF markedly disrupted both leptin and insulin homeostasis, particularly in
apoE3 mice. Indeed, only CPF-fed apoE3 mice exhibited an increased phosphorylation ratio of
STAT3, as well as increased total SOCS3 protein levels. Similarly, the exposure to CPF drastically
reduced the phosphorylation ratio of both AKT and GSK3β, especially in apoE3mice. Overall, CPF
reduced the expression of the three pon genes, principally in C57BL/6 and apoE3 mice. These
results provide notable mechanistic insights on the metabolic effects of the pesticide CPF, and attest
the increased vulnerability of apoE3 carriers to its metabolic-disruptor role.
18 Food Chem Toxicol. 2016 Jun; 92:224–35. doi: 10.1016/j.fct.2016.03.029. Epub 2016 Apr 20. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Apolipoprotein E (APOE) genotype and the pesticide chlorpyrifos modulate attention, motivation and additional useful data on low dose chlorpyrifos effects.
impulsivity in female mice in the 5-choice serial reaction time task.
Peris-Sampedro F, Reverte I, Basaure P, Cabré M, Domingo JL, Colomina MT.
Organophosphate pesticides—and chlorpyrifos (CPF) in particular—contribute to a wide range of
neurobehavioural disorders. Most experimental research focuses on learning and memory
processes, while other behaviours remain understudied. The isoforms of the human apolipoprotein E
(apoE) confer different cognitive skills on their carriers, but data on this topic are still limited. The
current study was performed to assess whether the APOE genotypic variability differently modulates
the effects of CPF on attentional performance, inhibitory control and motivation. Human apoE
targeted replacement adult female mice (apoE2, apoE3 and apoE4) were trained to stably perform
the 5-choice serial reaction time task (5-CSRTT). Animals were then subjected to daily dietary CPF
(3.75 mg/kg body weight) for 4 weeks. After CPF exposure, we established a 4-week CPF-free
84 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

period to assess recovery. All individuals acquired the task, apoE2 mice showed enhanced learning,
while apoE4 mice displayed increased premature and perseverative responding. This genotype-
dependent lack of inhibitory control was reversed by CPF. Overall, the pesticide induced protracted
impairments in sustained attention and motivation, and it reduced anticipatory responding. ApoE3
mice exhibited delayed attentional disruptions throughout the wash-out period. Taken together,
these findings provide notable evidence on the emergence of CPF-related attentional and
motivational deficits.
19 Sci Rep. 2016 Dec 1; 6:38131. doi: 10.1038/srep38131. Genomic thyroid toxicity signatures were only detected in
vivo at doses ≥ 1 mg/kg bw/day ie above the NOEL for
Pesticide toxicogenomics across scales: in vitro transcriptome predicts mechanisms and outcomes inhibition of blood cholinesterases. Furthermore because of
of exposure in vivo. their lower thyroid efficiency, mice are a supersensitive model
Porreca I, D'Angelo F, De Franceschi L, Mattè A, Ceccarelli M, Iolascon A, Zamò A, Russo F, Ravo for these types of effects cf. humans. Overall the study
M, Tarallo R, Scarfò M, Weisz A, De Felice M, Mallardo M, Ambrosino C. provides no additional useful data on low dose chlorpyrifos
effects.
In vitro Omics analysis (ie transcriptome) is suggested to predict in vivo toxicity and adverse effects
in humans, although the causal link between high-throughput data and effects in vivo is not easily
established. Indeed, the chemical-organism interaction can involve processes, such as adaptation,
not established in cell cultures. Starting from this consideration we investigate the transcriptomic
response of immortalized thyrocytes to ethylenthiourea and chlorpyrifos. In vitro data revealed
specific and common genes/mechanisms of toxicity, controlling the proliferation/survival of the
thyrocytes and unrelated hematopoietic cell lineages. These results were phenotypically confirmed
in vivo by the reduction of circulating T4 hormone and the development of pancytopenia after long
exposure. Our data imply that in vitro toxicogenomics is a powerful tool in predicting adverse effects
in vivo, experimentally confirming the vision described as Tox21c (Toxicity Testing in the 21st
century) although not fully recapitulating the biocomplexity of a living animal.
20 Environ Health. 2015 Apr 2; 14:32. doi: 10.1186/s12940-015-0019-6. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Effects of maternal chlorpyrifos diet on social investigation and brain neuroendocrine markers in the additional useful data on low dose chlorpyrifos effects.
offspring—a mouse study.
Venerosi A, Tait S, Stecca L, Chiarotti F, De Felice A, Cometa MF, Volpe MT, Calamandrei G, Ricceri
L.
BACKGROUND: Chlorpyrifos (CPF) is one of the most widely used organophosphate pesticides
worldwide. Epidemiological studies on pregnant women and their children suggest a link between in
utero CPF exposure and delay in psychomotor and cognitive maturation. A large number of studies
APPENDIX 2 85

Summary
Study and published abstract Study evaluation
No.

in animal models have shown adverse effects of CPF on developing brain and more recently on
endocrine targets. Our aim was to determine if developmental exposure to CPF affects social
responsiveness and associated molecular neuroendocrine markers at adulthood. METHOD:
Pregnant CD1 outbred mice were fed from gestational day 15 to lactation day 14 with either a CPF-
added (equivalent to 6 mg/kg/bw/day during pregnancy) or a standard diet. We then assessed in the
offspring the long-term effects of CPF exposure on locomotion, social recognition performances and
gene expression levels of selected neurondocrine markers in amygdala and hypothalamus.
RESULTS: No sign of CPF systemic toxicity was detected. CPF induced behavioural alterations in
adult offspring of both sexes: CPF-exposed males displayed enhanced investigative response to
unfamiliar social stimuli, whereas CPF-exposed females showed a delayed onset of social
investigation and lack of reaction to social novelty. In parallel, molecular effects of CPF were sex
dimorphic: in males CPF increased expression of estrogen receptor beta in hypothalamus and
decreased oxytocin expression in amygdala; CPF increased vasopressin 1a receptor expression in
amygdala in both sexes. CONCLUSIONS: These data indicate that developmental CPF affects
mouse social behaviour and interferes with development of sex-dimorphic neuroendocrine pathways
with potential disruptive effects on neuroendocrine axes homeostasis. The route of exposure
selected in our study corresponds to relevant human exposure scenarios, our data thus supports the
view that neuroendocrine effects, especially in susceptible time windows, should deserve more
attention in risk assessment of OP insecticides.
21 Pest Manag Sci. 2018 Jun; 74(6):1424–30. doi: 10.1002/ps.4825. Epub 2018 Feb 22. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in mice. The study provides no
Chimeric mice with humanized liver as a model for testing organophosphate and carbamate additional useful data on low dose chlorpyrifos effects.
pesticide exposure.
Suemizu H, Kawai K, Murayama N, Nakamura M, Yamazaki H.
BACKGROUND: Diagnosis of acute intoxication with organophosphate (OP) or carbamate (CM)
pesticides in humans is achieved by measuring plasma butyrylcholinesterase (BuChE) activity.
However, BuChE activity is not an ideal biomarker in experimental animal models. The aim of this
study was to establish an experimental mouse model for evaluating exposure to OP and CM
pesticides by monitoring BuChE activity using chimeric mice in which the liver was reconstituted
with human hepatocytes.
RESULTS: A single oral administration of acephate (300 mg/kg), chlorpyrifos (10 mg/kg), fenobucarb
(300 mg/kg) or molinate (250 mg/kg) in chimeric mice led to inhibition of >95%, > 95%, 28% and
60% of plasma BuChE activity after 7, 0.5, 0.5 and 7 h, respectively. Dose-dependent decreases in
86 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

plasma BuChE activity were also observed for acephate and chlorpyrifos. A 5-day repeated-dose
study with 10 or 30 mg/kg acephate found a constitutive reduction in plasma BuChE activity to 80%
and 70% of pre-dose levels, respectively.
CONCLUSION: Changes in plasma BuChE activity in chimeric mice with humanized liver clearly
reflected the exposure levels of OP and CM pesticides. These results suggest that the humanized-
liver mouse model may be suitable for estimating levels of exposure to these pesticides in humans.
Rat
22 Chem Biol Interact. 2018 Sep 27; 296:105–16. doi: 10.1016/j.cbi.2018.09.016. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Diphenyl diselenide abrogates brain oxidative injury and neurobehavioural deficits associated with additional useful data on low dose chlorpyrifos effects.
pesticide chlorpyrifos exposure in rats.
Adedara IA, Owoeye O, Awogbindin IO, Ajayi BO, Rocha JBT, Farombi EO.
Exposure to pesticide chlorpyrifos (CPF) is associated with neurodevelopmental toxicity both in
humans and animals. Diphenyl diselenide (DPDS) is a simple synthetic organoselenium well
reported to possess antioxidant, anti-inflammatory and neuroprotective effects. However, there is
paucity of information on the beneficial effects of DPDS on CPF-mediated brain injury and
neurobehavioural deficits. The present study investigated the neuroprotective mechanism of DPDS
in rats sub-chronically treated with CPF alone at 5 mg/kg body weight or orally co-treated with DPDS
at 2.5 and 5 mg/kg body weight for 35 consecutive days. Endpoint analyses using video-tracking
software in a novel environment revealed that co-treatment with DPDS significantly (p < 0.05)
protected against CPF-mediated locomotor and motor deficits precisely the decrease in maximum
speed, total distance travelled, body rotation, absolute turn angle, forelimb grip strength as well as
the increase in negative geotaxis and incidence of fecal pellets. The enhancement in the
neurobehavioural activities of rats co-treated with DPDS was verified by track plot analyses.
Besides, DPDS assuaged CPF-induced decrease in acetylcholinesterase and antioxidant enzymes
activities and the increase in myeloperoxidase activity and lipid peroxidation level in the mid-brain,
cerebral cortex and cerebellum of the rats. Histologically, DPDS co-treatment abrogated CPF-
mediated neuronal degeneration in the cerebral cortex, dentate gyrus and cornu ammonis3 in the
treated rats. In conclusion, the neuroprotective mechanisms of DPDS is related to the prevention of
oxidative stress, enhancement of redox status and acetylcholinesterase activity in brain regions of
the rats. DPDS may be a promising chemotherapeutic agent against brain injury resulting from CPF
exposure.
APPENDIX 2 87

Summary
Study and published abstract Study evaluation
No.

23 Neurotoxicology. 2017 Mar; 59:183–90. doi: 10.1016/j.neuro.2015.11.016. Epub 2015 Nov 28. Alterations in behaviour (time of emergence from a dark
container into a lighted area) were detected following dosing
Decreased anxiety in juvenile rats following exposure to low levels of chlorpyrifos during at ≥ 0.5 mg/kg bw/d for 7 days over post-natal days 10–16.
development. The LOAEL for this effect was 0.5 mg/kg bw/d which exceeds
Carr RL, Armstrong NH, Buchanan AT, Eells JB, Mohammed AN, Ross MK, Nail CA. the NOELs for inhibition of blood cholinesterases used in this
report (0.1 mg/kg bw/d) and is equal to the NOEL for brain
Exposure to chlorpyrifos (CPF) during the late preweanling period in rats inhibits the cholinesterase in this species (0.5 mg/kg bw/d). Thus the
endocannabinoid metabolizing enzymes fatty acid hydrolase (FAAH) and monoacylglycerol lipase findings of this paper support the use of health based
(MAGL), resulting in accumulation of their respective substrates anandamide (AEA) and 2- guidance values based on the NOELs for inhibition of blood
arachidonylglycerol (2-AG). This occurs at 1.0mg/kg, but at a lower dosage (0.5mg/kg) only FAAH cholinesterases.
and AEA are affected with no measurable inhibition of either cholinesterase (ChE) or MAGL. The
endocannabinoid system plays a vital role in nervous system development and may be an important
developmental target for CPF. The endocannabinoid system plays an important role in the regulation
of anxiety and, at higher dosages, developmental exposure to CPF alters anxiety-like behaviour.
However, it is not clear whether exposure to low dosages of CPF that do not inhibit ChE will cause
any persistent effects on anxiety-like behaviour. To determine if this occurs, 10-day old rat pups
were exposed daily for 7 days to either corn oil or 0.5, 0.75, or 1.0mg/kg CPF by oral gavage. At 12h
following the last CPF administration, 1.0mg/kg resulted in significant inhibition of FAAH, MAGL, and
ChE, whereas 0.5 and 0.75mg/kg resulted in significant inhibition of only FAAH. AEA levels were
significantly elevated in all three treatment groups as were palmitoylethanolamide and
oleoylethanolamide, which are also substrates for FAAH. 2-AG levels were significantly elevated by
0.75 and 1.0mg/kg but not 0.5mg/kg. On day 25, the latency to emerge from a dark container into a
highly illuminated novel open field was measured as an indicator of anxiety. All three CPF treatment
groups spent significantly less time in the dark container prior to emerging as compared to the
control group, suggesting a decreased level of anxiety. This demonstrates that repeated
preweanling exposure to dosages of CPF that do not inhibit brain ChE can induce a decline in the
level of anxiety that is detectable during the early postweanling period.
24 Arh Hig Rada Toksikol. 2015 Jun; 66(2):121–7. doi: 10.1515/aiht-2015-66-2624. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
ADHD-like behaviour in the offspring of female rats exposed to low chlorpyrifos doses before additional useful data on low dose chlorpyrifos effects.
pregnancy.
Grabovska S, Salyha Y.
The aim of this study was to investigate how chronic low-dose chlorpyrifos exposure of female
Wistar rats before and during pregnancy affects behavioural parameters in their offspring. Four
88 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

months before pregnancy, we exposed three groups of rats to chlorpyrifos doses of 5, 10, and 15
mg kg-1 body weight every day for 30 days, whereas one group received a single 30 mg kg-1 dose
on gestational day 6. When the offspring of the exposed rats grew up, we studied their anxiety rate,
motor activity, and cognitive abilities using the respective behavioural tests: open field test,
dark/light box, and the extrapolation escape test. The offspring of rats exposed before pregnancy
had significantly higher activity rate than controls, and even showed motor agitation and
hyperactivity signs. The offspring of rats exposed to the single dose had difficulties solving the
extrapolation escape test and showed poorer short- and long-term memory performance. This
confirmed that even pre-pregnancy chlorpyrifos exposure can cause neurobehavioural
consequences in offspring. Even though the mechanisms of the observed changes remain unclear
and need further investigation, these data seem alarming and may serve as an important argument
for revising the terms of safe pesticide use.
25 Fiziol Zh. 2015; 61(2):94–101. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
The effect of chronic intoxication with low doses of chlorpyrifos on the behavioural parameters of additional useful data on low dose chlorpyrifos effects.
female rats.
[Article in Ukrainian; only the English language abstract has been evaluated]
Grabovska SV, Salyha YT.
We explored the effect of chronic chlorpyrifos intake on the behaviour of female laboratory rats.
The animals were exposed to chlorpyrifos in doses 5 and 10 mg/kg daily for a month long and tested
with behavioural tests: Morris water maze, Open field test, and Dark/light box. Everyday chlorpyrifos
intake in dose 10 mg/kg was revealed to cause long- and short-term spatial memory lesion,
decrease in total activity, sharp changes in the level of anxiety (it significantly increased at the start
of xenobiotic intake and then normalized with time). Similar trends, though less intense, were
observed in the group exposed
to 5 mg/kg chlorpyrifos. From the survey results it can be concluded that long-term (30 days long)
intake
of even low doses of chlorpyrifos can cause adverse effects on the functional state of nervous
system.
26 J Toxicol Pharmacol. 2017; 1. pii: 018. Epub 2017 Dec 30.
Effects of chlorpyrifos or methyl parathion on regional cholinesterase activity and muscarinic
receptor subtype binding in juvenile rat brain.
APPENDIX 2 89

Summary
Study and published abstract Study evaluation
No.

Guo-Ross SX, Meek EC, Chambers JE, Carr RL.


The effects of developmental exposure to two organophosphorus (OP) insecticides, chlorpyrifos
(CPF) and methyl parathion (MPS), on cholinesterase (ChE) activity and muscarinic acetylcholine
receptor (mAChR) binding were investigated in preweanling rat brain. Animals were orally gavaged
daily with low, medium, and high dosages of the insecticides using an incremental dosing regimen
from postnatal day 1 (PND1) to PND20. On PND12, PND17 and PND20, the cerebral cortex, corpus
striatum, hippocampus, and medulla-pons were collected for determination of ChE activity, total
mAChR density, and the density of the individual mAChR subtypes. ChE activity was inhibited by the
medium and high dosages of CPF and MPS at equal levels in all four brain regions at all three ages
examined. Exposure to both compounds decreased the levels of the M1, M2/M4, and M3 subtypes
and the total mAChR level in all brain regions, but the effects varied by dosage group and brain
region. On PND12, only the high dosages induced receptor changes while on PND17 and PND20,
greater effects became evident. In general, the effects on the M1 subtype and total receptor levels
appeared to be greater in the cerebral cortex and hippocampus than in the corpus striatum and
medulla-pons. This did not appear to be the case for the M2/M4 and M3 subtypes effects. The
differences between CPF and MPS were minimal even though in some cases, CPF exerted
statistically greater effects than MPS did. In general, repeated exposure to organophosphorus
insecticides can alter the levels of the various mAChR subtypes in various brain regions which could
induce perturbation in cholinergic neurochemistry during the maturation of the brain regions.

C
hlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
additional useful data on low dose chlorpyrifos effects.
27 Neurotoxicology. 2015 Mar; 47:17–26. doi: 10.1016/j.neuro.2015.01.002. Epub 2015 Jan 19. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Repeated exposure to chlorpyrifos leads to prolonged impairments of axonal transport in the living additional useful data on low dose chlorpyrifos effects.
rodent brain.
Hernandez CM, Beck WD, Naughton SX, Poddar I, Adam BL, Yanasak N, Middleton C, Terry AV Jr.
90 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

The toxicity of the class of chemicals known as the organophosphates (OP) is most commonly
attributed to the inhibition of the enzyme acetylcholinesterase. However, there is significant
evidence that this mechanism may not account for all of the deleterious neurologic and
neurobehavioural symptoms of OP exposure, especially those associated with levels that produce
no overt signs of acute toxicity. In the study described here we evaluated the effects of the
commonly used OP-pesticide, chlorpyrifos (CPF) on axonal transport in the brains of living rats
using manganese (Mn(2+))-enhanced magnetic resonance imaging (MEMRI) of the optic nerve (ON)
projections from the retina to the superior colliculus (SC). T1-weighted MEMRI scans were
evaluated at 6 and 24h after intravitreal injection of Mn(2+). As a positive control for axonal
transport deficits, initial studies were conducted with the tropolone alkaloid colchicine administered
by intravitreal injection. In subsequent studies both single and repeated exposures to CPF were
evaluated for effects on axonal transport using MEMRI. As expected, intravitreal injection of
colchicine (2.5μg) produced a robust decrease in transport of Mn(2+) along the optic nerve (ON)
and to the superior colliculus (SC) (as indicated by the reduced MEMRI contrast). A single
subcutaneous (s.c.) injection of CPF (18.0mg/kg) was not associated with significant alterations in
the transport of Mn(2+). Conversely, 14-days of repeated s.c. exposure to CPF (18.0mg/kg/day) was
associated with decreased transport of Mn(2+) along the ONs and to the SC, an effect that was also
present after a 30-day (CPF-free) washout period. These results indicate that repeated exposures to
a commonly used pesticide, CPF can result in persistent alterations in axonal transport in the living
mammalian brain. Given the fundamental importance of axonal transport to neuronal function, these
observations may (at least in part) explain some of the long term neurological deficits that have
been observed in humans who have been repeatedly exposed to doses of OPs not associated with
acute toxicity.
28 Toxicology. 2016 Jan 18; 340:53–62. doi: 10.1016/j.tox.2016.01.001. Epub 2016 Jan 13. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Repeated exposure to neurotoxic levels of chlorpyrifos alters hippocampal expression of additional useful data on low dose chlorpyrifos effects.
neurotrophins and neuropeptides.
Lee YS, Lewis JA, Ippolito DL, Hussainzada N, Lein PJ, Jackson DA, Stallings JD.
Chlorpyrifos (CPF), an organophosphorus pesticide (OP), is one of the most widely used pesticides
in the world. Subchronic exposures to CPF that do not cause cholinergic crisis are associated with
problems in cognitive function (ie, learning and memory deficits), but the biological mechanism(s)
underlying this association remain speculative. To identify potential mechanisms of subchronic CPF
neurotoxicity, adult male Long Evans (LE) rats were administered CPF at 3 or 10mg/kg/d (s.c.) for
21 days. We quantified mRNA and non-coding RNA (ncRNA) expression profiles by RNA-seq,
APPENDIX 2 91

Summary
Study and published abstract Study evaluation
No.

microarray analysis and small ncRNA sequencing technology in the CA1 region of the hippocampus.
Hippocampal slice immunohistochemistry was used to determine CPF-induced changes in protein
expression and localization patterns. Neither dose of CPF caused overt clinical signs of cholinergic
toxicity, although after 21 days of exposure, cholinesterase activity was decreased to 58% or 13% of
control levels in the hippocampus of rats in the 3 or 10mg/kg/d groups, respectively. Differential
gene expression in the CA1 region of the hippocampus was observed only in the 10mg/kg/d dose
group relative to controls. Of the 1382 differentially expressed genes identified by RNA-seq and
microarray analysis, 67 were common to both approaches. Differential expression of six of these
genes (Bdnf, Cort, Crhbp, Nptx2, Npy and Pnoc) was verified in an independent CPF exposure
study; immunohistochemistry demonstrated that CRHBP and NPY were elevated in the CA1 region
of the hippocampus at 10mg/kg/d CPF. Gene ontology enrichment analysis suggested association of
these genes with receptor-mediated cell survival signaling pathways. miR132/212 was also elevated
in the CA1 hippocampal region, which may play a role in the disruption of neurotrophin-mediated
cognitive processes after CPF administration. These findings identify potential mediators of CPF-
induced neurobehavioural deficits following subchronic exposure to CPF at a level that inhibits
hippocampal cholinesterase to less than 20% of control. An equally significant finding is that
subchronic exposure to CPF at a level that produces more moderate inhibition of hippocampal
cholinesterase (approximately 50% of control) does not produce a discernable change in gene
expression.
29 Neurotoxicology. 2015 Jan; 46:12–8. doi: 10.1016/j.neuro.2014.11.001. Epub 2014 Nov 20. Chlorpyrifos oxon dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats by chlorpyrifos. The study
The cannabinoid receptor antagonist AM251 increases paraoxon and chlorpyrifos oxon toxicity in provides no additional useful data on low dose chlorpyrifos
rats. effects.
Liu J, Pope C.
Organophosphorus anticholinesterases (OPs) elicit acute toxicity by inhibiting acetylcholinesterase
(AChE), leacceptable daily intakeng to acetylcholine accumulation and overstimulation of cholinergic
receptors. Endocannabinoids (eCBs, eg, arachidonoyl ethanolamide [AEA] and 2-arachidonoyl
glycerol [2-AG]) are neuromodulators that regulate neurotransmission by reducing neurotransmitter
release. The eCBs are degraded by the enzymes fatty acid amide hydrolase (FAAH, primarily
involved in hydrolysis of AEA) and monoacylglycerol lipase (MAGL, primarily responsible for
metabolism of 2-AG). We previously reported that the cannabinoid receptor agonist WIN 55,212-2
reduced cholinergic toxicity after paraoxon exposure. This study compared the effects of the
cannabinoid receptor antagonist AM251 on acute toxicity following either paraoxon (PO) or
chlorpyrifos oxon (CPO). CPO was more potent in vitro than PO at inhibiting AChE (≈ 2 fold), FAAH
92 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

(≈ 8 fold), and MAGL (≈ 19 fold). Rats were treated with vehicle, PO (0.3 and 0.6 mg/kg, sc) or CPO
(6 and 12 mg/kg, sc) and subsets treated with AM251 (3mg/kg, ip; 30 min after OP). Signs of toxicity
were recorded for 4h and rats were then sacrificed. OP-treated rats showed dose-related involuntary
movements, with AM251 increasing signs of toxicity with the lower dosages. PO and CPO elicited
excessive secretions, but AM251 had no apparent effect with either OP. Lethality was increased by
AM251 with the higher dosage of PO, but no lethality was noted with either dosage of CPO, with or
without AM251. Both OPs caused extensive inhibition of hippocampal AChE and FAAH (>80–90%),
but only CPO inhibited MAGL (37–50%). These results provide further evidence that eCB signaling
can influence acute OP toxicity. The selective in vivo inhibition of MAGL by CPO may be important
in the differential lethality noted between PO and CPO with AM251 co-administration.
30 Neurotoxicology. 2016 Mar; 53:85–92. doi: 10.1016/j.neuro.2015.12.016. Epub 2015 Dec 31. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Chronic dietary chlorpyrifos causes long-term spatial memory impairment and thigmotaxic additional useful data on low dose chlorpyrifos effects.
behaviour.
López-Granero C, Ruiz-Muñoz AM, Nieto-Escámez FA, Colomina MT, Aschner M, Sánchez-Santed
F.
Little is known about the long-term effects of chronic exposure to low-level organophosphate (OP)
pesticides, and the role of neurotransmitter systems, other than the cholinergic system, in mediating
OP neurotoxicity. In this study, rats were administered 5mg/kg/day of chlorpyrifos (CPF) for 6
months commencing at 3-months-of-age. The animals were examined 7 months later (at 16-months-
of-age) for spatial learning and memory in the Morris water maze (MWM) and locomotor activity. In
addition, we assessed the chronic effects of CPF on glutamatergic and gamma-aminobutyric acid
(GABAergic) function using pharmacological challenges with dizocilpine (MK801) and diazepam.
Impaired performance related to altered search patterns, including thigmotaxis and long-term spatial
memory was noted in the MWM in animals exposed to CPF, pointing to dietary CPF-induced
behavioural disturbances, such as anxiety. Twenty-four hours after the 31st session of repeated
acquisition task, 0.1mg/kg MK801, an N-methyl-d-aspartate (NMDA) antagonist was
intraperitoneally (i.p.) injected for 4 consecutive days. Decreased latencies in the MWM in the
control group were noted after two sessions with MK801 treatment. Once the MWM assessment was
completed, animals were administered 0.1 or 0.2mg/kg of MK801 and 1 or 3mg/kg of diazepam i.p.,
and tested for locomotor activity. Both groups, the CPF dietary and control, displayed analogous
performance in motor activity. In conclusion, our data point to a connection between the long-term
spatial memory, thigmotaxic response and CPF long after the exposure ended.
APPENDIX 2 93

Summary
Study and published abstract Study evaluation
No.

31 Environ Toxicol Pharmacol. 2018 Apr; 59:17–23. doi: 10.1016/j.etap.2018.02.007. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Epub 2018 Feb 23. additional useful data on low dose chlorpyrifos effects.
Alpha lipoic acid protects against chlorpyrifos-induced toxicity in Wistar rats via modulating the
apoptotic pathway.
Mohamed WR, Mehany ABM, Hussein RM.
The chronic exposure to chlorpyrifos (CPF) pesticide induces several human disorders including
hepatotoxicity. Alpha-lipoic acid (ALA) is a natural antioxidant compound found in plants and
animals.
The present study aimed to investigate the possible protective effect of ALA against CPF-induced
hepatotoxicity and the possible underlying molecular mechanism. Thirty-two male Wistar rats were
divided into: Normal rats received only vehicle; ALA group received ALA (10 mg/kg, i.p.); CPF group
received CPF (18 mg/kg, s.c.) and CPF-ALA group received CPF (18 mg/kg, s.c.) once daily for 14
days. The present results demonstrated that administration of ALA significantly improved liver
functions (p < 0.05) and limited the histopathological lesions induced by CPF in liver tissues.
Furthermore, ALA decreased hepatic malondialdehyde contents while increased the glutathione
peroxidase, catalase, superoxide dismutase and acetylcholinesterase activities. Interestingly, ALA
showed significant antiapoptotic effects through downregulation of Bax and Caspase-3 expression
levels. In conclusion, ALA possess protective effects against CPF-induced liver injury through
attenuation of apoptosis and oxidative stress.
32 Environ Sci Pollut Res Int. 2016 Jan; 23(2):1852–9. doi: 10.1007/s11356-015-5448-9. Epub 2015 Chlorpyrifos dose used exceeds the in vivo NOAELs for
Sep 24. cholinesterase inhibition in rats. The study provides no
additional useful data on low dose chlorpyrifos effects.
Neuro and renal toxicity induced by chlorpyrifos and abamectin in rats: Toxicity of insecticide
mixture.
Nasr HM, El-Demerdash FM, El-Nagar WA.
Oxidative stress by increased production of reactive oxygen species has been implicated in
pesticides toxicity. This study focused on the toxicological effects of chlorpyrifos, an
organophosphate insecticide and abamectin, a biocide each alone or in combination on antioxidant
status, and oxidative stress biomarkers in brain and kidney. Animals were divided into four groups.
The first group was used as control while groups 2, 3, and 4 were treated with chlorpyrifos (CPF;
14.9 mg/kg bw), abamectin (ABM; 30 mg/kg bw), and chlorpyrifos plus abamectin, respectively. Rats
were treated daily with the tested compounds by oral gavages for 30  days. Results revealed that
94 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

thiobarbituric acid-reactive substances (TBARS) levels were significantly increased in brain and
kidney due to insecticides administration. On the other hand, reduced glutathione (GSH) and protein
contents in addition to the activities of antioxidant enzymes, alkaline phosphatase (ALP), and
acetylcholinesterase (AChE) were significantly decreased in rat organs. A significant induction in
lactate dehydrogenase (LDH) activity, urea, and creatinine levels were also observed. The response
was more pronounced in rats treated with both CPF and ABM. Results showed that the used
insecticides had the propensity to cause significant oxidative damage in rat brain and kidney which
is associated with marked perturbations in antioxidant defense system. It can be concluded that
antioxidant enzymes can be used as potential biomarkers of toxicity associated with pesticides
exposure.
33 PLoS One. 2016 Oct 19; 11(10):e0164614. doi: 10.1371/journal.pone.0164614. eCollection 2016. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Inulin Supplementation Lowered the Metabolic Defects of Prolonged Exposure to Chlorpyrifos from additional useful data on low dose chlorpyrifos effects.
Gestation to Young Adult Stage in Offspring Rats.
Reygner J, Lichtenberger L, Elmhiri G, Dou S, Bahi-Jaber N, Rhazi L, Depeint F, Bach V, Khorsi-
Cauet H, Abdennebi-Najar L.
Increasing evidence indicates that chlorpyrifos (CPF), an organophosphorus insecticide, is involved
in metabolic disorders. We assess the hypothesis whether supplementation with prebiotics from
gestation to adulthood, through a modulation of microbiota composition and fermentative activity,
alleviates CPF induced metabolic disorders of 60 days old offspring. 5 groups of Wistar rats, from
gestation until weaning, received two doses of CPF pesticide: 1 mg/kg/day (CPF1) or 3.5 mg/kg/day
(CPF3.5) with free access to inulin (10g/L in drinking water). Then male pups received the same
treatment as dams. Metabolic profile, leptin sensitivity, insulin receptor (IR) expression in liver, gut
microbiota composition and short chain fatty acid composition (SCFAs) in the colon, were analyzed
at postnatal day 60 in the offspring (PND 60). CPF3.5 increased offspring's birth body weight (bw)
but decreased bw at PND60. Inulin supplementation restored the bw at PND 60 to control levels.
Hyperinsulinemia and decrease in insulin receptor β in liver were seen in CPF1 exposed rats. In
contrast, hyperglycemia and decrease in insulin level were found in CPF3.5 rats. Inulin restored the
levels of some metabolic parameters in CPF groups to ranges comparable with the controls. The
total bacterial population, short chain fatty acid (SCFA) production and butyrate levels were
enhanced in CPF groups receiving inulin. Our data indicate that developmental exposure to CPF
interferes with metabolism with dose related effects evident at adulthood. By modulating microbiota
population and fermentative activity, inulin corrected adult metabolic disorders of rats exposed to
CPF during development. Prebiotics supply may be thus considered as a novel nutritional strategy
APPENDIX 2 95

Summary
Study and published abstract Study evaluation
No.

to counteract insulin resistance and diabetes induced by a continuous pesticide exposure.


34 Ukr Biochem J. 2015 Sep-Oct; 87(5):124–32. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Changes in glutathione system and lipid peroxidation in rat blood during the first hour after additional useful data on low dose chlorpyrifos effects.
chlorpyrifos exposure.
Rosalovsky VP, Grabovska SV, Salyha YT.
Chlorpyrifos (CPF) is a highly toxic organophosphate compound, widely used as an active
substance of many insecticides. Along with the anticholinesterase action, CPF may affect other
biochemical mechanisms, particularly through disrupting pro- and antioxidant balance and inducing
free-racceptable daily intakecal oxidative stress. Origins and occurrence of these phenomena are
still not fully understood. The aim of our work was to investigate the effects of chlorpyrifos on key
parameters of glutathione system and on lipid peroxidation in rat blood in the time dynamics during
one hour after exposure. We found that
a single exposure to 50 mg/kg chlorpyrifos caused a linear decrease in butyryl cholinesterase
activity, increased activity of glutathione peroxidase and glutathione reductase, alterations in the
levels of glutathione, TBA-active products and lipid hydroperoxides during 1 hour after poisoning.
The most significant changes in studied parameters were detected at the 15–30th minutes after
chlorpyrifos exposure.
35 Neurotoxicology. 2015 Sep; 50:149–56. doi: 10.1016/j.neuro.2015.08.010. Epub 2015 Aug 19. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Low-level repeated exposure to diazinon and chlorpyrifos decrease anxiety-like behaviour in adult additional useful data on low dose chlorpyrifos effects.
male rats as assessed by marble burying behaviour.
Savy CY, Fitchett AE, McQuade R, Gartside SE, Morris CM, Blain PG, Judge SJ.
Occupational exposure to organophosphate (OPs) pesticides is reported to increase in the risk of
developing anxiety and depression. Preclinical studies using OP levels, which inhibit
acetylcholinesterase activity, support the clinical observations, but little is known of the effects of
exposure below this threshold. We examined the effects of low level OP exposure on behaviours
and neurochemistry associated with affective disorders. Adult rats were administered either diazinon
(1 mg/kg i.p.) which is present in sheep dip and flea collars, chlorpyrifos (1 mg/kg i.p.) which is
present in crop sprays, or vehicle for 5 days. OP exposure did not affect acetylcholinesterase
activity (blood, cerebellum, caudate putamen, hippocampus, prefrontal cortex), anhedonia-like
behaviour (sucrose preference), working memory (novel object recognition), locomotor activity or
anxiety-like behaviour in the open field arena. In contrast OP exposure attenuated marble burying
96 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

behaviour, an ethological measure of anxiety. The diazinon-induced reduction in marble burying


persisted after exposure cessation. In comparison to vehicle, dopamine levels were lowered by
chlorpyrifos, but not diazinon. 5-HT levels and turnover were unaffected by OP exposure. However,
5-HT transporter expression was reduced by diazinon suggesting subtle changes in 5-HT
transmission. These data indicate exposure to occupational and domestic OPs, below the threshold
to inhibit acetylcholinesterase, can subtly alter behaviour and neurochemistry.
36 Toxicol Sci. 2018 Sep 1; 165(1):244–53. doi: 10.1093/toxsci/kfy158. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
The Organophosphorus Pesticide Chlorpyrifos Induces Sex-Specific Airway Hyperreactivity in Adult additional useful data on low dose chlorpyrifos effects.
Rats.
Shaffo FC, Grodzki AC, Schelegle ES, Lein PJ.
Occupational and environmental exposures to organophosphorus pesticides (OPs) are associated
with increased incidence of asthma and other pulmonary diseases. Although the canonical
mechanism of OP neurotoxicity is inhibition of acetylcholinesterase (AChE), it was previously
reported that the OP chlorpyrifos (CPF) causes airway hyperreactivity (AHR) in guinea pigs at levels
that do not inhibit lung or brain AChE. The guinea pig is considered to have inherently
hyperresponsive airways, thus, cross-species validation is needed to confirm relevance to humans.
Additionally, sex differences in asthma incidence have been demonstrated in the human population,
but whether OP-induced AHR is sex-dependent has not been systematically studied in a preclinical
model. In this study, 8-week old male and female Sprague Dawley rats were administered CPF at
doses causing comparable AChE inhibition in whole lung homogenate (30  mg/kg in males, 7 mg/kg
in females, sc) prior to assessing pulmonary mechanics in response to electrical stimulation of the
vagus nerves at 24 h, 48 h, 72 h, 7 d or 14 d post-exposure in males, and 24 h or 7 d post-exposure
in females. CPF significantly potentiated vagally induced airway resistance and tissue elastance at 7
d post-exposure in males, and at 24 h and 7 d post-exposure in females. These effects occurred
independent of significant AChE inhibition in cerebellum, blood, trachealis, or isolated airway,
suggesting that AChE independent OP-induced airway hyperreactivity is a cross-species
phenomenon. These findings have significant implications for assessing the risk posed by CPF, and
potentially other OPs, to human health and safety.
37 Springerplus. 2016 Aug 12; 5(1):1344. doi: 10.1186/s40064-016-3004-9. eCollection 2016. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Induction of neurotoxicity by organophosphate pesticide chlorpyrifos and modulating role of cow additional useful data on low dose chlorpyrifos effects.
urine.
APPENDIX 2 97

Summary
Study and published abstract Study evaluation
No.

Sharma S, Chadha P.
INTRODUCTION: Organophosphate pesticides are among the most widely used synthetic chemicals
for controlling a wide variety of pests and for domestic purposes. Among these chlorpyrifos (CPF) is
the most extensively used pesticide throughout the world, including India. OBJECTIVE: The present
study was undertaken to examine the neurotoxicity induced by CPF and modulatory effect of cow
urine as a natural antioxidant alternative to reduce the neurotoxic effects of CPF. DESIGN: For this
purpose LD50 was determined and one fourth of LD50 was selected (38 mg/kg body weight (b.wt))
for treatment of rats. The antioxidant level of cow urine was determined by ABTS assay. RESULTS:
Exposure to pesticides resulted in significant reduction in the acetylcholinestrase (AChE) activity (P
≤ 0.01). However, groups pretreated with cow urine had improved levels of AChE activity as
compared to CPF treated groups. CONCLUSION: Thus, the present findings clearly show that oral
CPF has the propensity to cause significant neurotoxicity in rat brains while cow urine treatment
alleviates CPF induced toxicity to a greater extent. In addition, AChE can be used as a potential
biomarker of toxicity associated with pesticide exposure.
38 Environ Sci Pollut Res Int. 2017 Jan; 24(1):381–7. doi: 10.1007/s11356-016-7799-2. Epub 2016 Oct Chlorpyrifos dose used exceeds the in vivo NOAELs for
8. cholinesterase inhibition in rats. The study provides no
additional useful data on low dose chlorpyrifos effects.
Chlorpyrifos pollution: its effect on brain acetylcholinesterase activity in rat and treatment of polluted
soil by indigenous Pseudomonas sp.
Sharma S, Singh PB, Chadha P, Saini HS.
The study was aimed to evaluate the levels of chlorpyrifos (CPF) pollution in agricultural soil of
Punjab, India, its detrimental effects on acetylcholinesterase (AChE) activity in rat brain and
bioremediation of soils polluted with CPF using indigenous and adapted bacterial lab isolate. The
analysis revealed that soil samples of Bathinda and Amritsar regions are highly contaminated with
chlorpyrifos showing 19 to 175 mg/kg concentrations of CPF. The non-targeted animals may get
poisoned with CPF by its indirect dermal absorption, inhalation of toxic fumes and regular
consumption of soiled food grains. The study indicated that even the lowermost concentrations of
CPF, 19 and 76 mg/kg of soil found in the Amritsar and Bathinda regions respectively can
significantly inhibit the AChE activity in rat brain within 24 h of its treatment. This represents the
antagonistic effect of CPF on AChE which is a prime neurotransmitter present in all living beings
including humans. In light of this, an attempt was made to remediate the polluted soil, a major
reservoir of CPF, using Pseudomonas sp. (ChlD), an indigenous bacterial isolate. The culture
efficiently degraded 10 to 100 mg/kg chlorpyrifos supplemented in the soil and utilized it as sole
98 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

source of carbon and energy for its growth. Thus, this study provides a detailed insight regarding the
level of CPF pollution in Punjab, its detrimental effects on mammals and bio-based solution to
remediate the sites polluted with CPF.
39 Neurosci Lett. 2017 Feb 22; 641:94–100. doi: 10.1016/j.neulet.2017.01.053. Epub 2017 Jan 24. None of these studies are regarded as being of regulatory
quality due to inadequacies of the statistical methods used,
Chlorpyrifos induces anxiety-like behaviour in offspring rats exposed during pregnancy. lack of blinding of the observer and/or other techniques to
Silva JG, Boareto AC, Schreiber AK, Redivo DD, Gambeta E, Vergara F, Morais H, Zanoveli JM, reduce observer bias, lack of assay validation by
Dalsenter PR. incorporation of appropriate control groups and lack of
historical control data. Because of the regulatory unreliability
Chlorpyrifos is a pesticide, member of the organophosphate class, widely used in several countries of the study a NOAEL was not determined.
to manage insect pests on many agricultural crops. Currently, chlorpyrifos health risks are being
reevaluated due to possible adverse effects, especially on the central nervous system. The aim of
this study was to investigate the possible action of this pesticide on the behaviours related to
anxiety and depression of offspring rats exposed during pregnancy. Wistar rats were treated orally
with chlorpyrifos (0.01, 0.1, 1 and 10mg/kg/day) on gestational days 14–20. Male offspring
behaviour was evaluated on post-natal days 21 and 70 by the elevated plus-maze test, open field
test and forced swimming test. The results demonstrated that exposure to 0.1, 1 or 10mg/kg/day of
chlorpyrifos could induce anxiogenic-like, but not depressive-like behaviour at post-natal day 21,
without causing fetal toxicity. This effect was reversed on post-natal day 70.
40 Environ Toxicol. 2016 Aug; 31(8):1017–26. doi: 10.1002/tox.22111. Epub 2015 Apr 10. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
The role of multifunctional drug therapy as an antidote to combat experimental subacute additional useful data on low dose chlorpyrifos effects.
neurotoxicity induced by organophosphate pesticides.
Singh S, Prakash A, Kaur S, Ming LC, Mani V, Majeed AB.
Organophosphate pesticides are used in agriculture where they are associated with numerous
cases of intentional and accidental misuse. These toxicants are potent inhibitors of cholinesterases
leacceptable daily intakeng to a massive build-up of acetylcholine which induces an array of
deleterious effects, including convulsions, oxidative damage and neurobehavioural deficits. Antidotal
therapies with atropine and oxime yield a remarkable survival rate, but fail to prevent neuronal
damage and behavioural problems. It has been indicated that multifunction drug therapy with
potassium channel openers, calcium channel antagonists and antioxidants (either single-agent
therapy or combination therapy) may have the potential to prevent cell death and/or slow down the
processes of secondary neuronal damage. The aim of the present study, therefore, was to make a
relative assessment of the potential effects of nicorandil (2 mg/kg), clinidipine (10 mg/kg), and grape
APPENDIX 2 99

Summary
Study and published abstract Study evaluation
No.

seed proanthocyanidin (GSPE) extract (200 mg/kg) individually against subacute chlorpyrifos
induced toxicity. The test drugs were administered to Wistar rats 2 h after exposure to Chlorpyrifos
(CPF). Different behavioural studies and biochemical estimation has been carried in the study. The
results showed that chronic administration of CPF significantly impaired learning and memory, along
with motor coordination, and produced a marked increase in oxidative stress along with significantly
reduced acetylcholine esterase (AChE) activity. Treatment with nicorandil, clinidipine and GSPE was
shown to significantly improve memory performance, attenuate oxidative damage and enhance
AChE activity in rats. The present study also suggests that a combination of nicorandil, clinidipine,
and GSPE has a better neuroprotective effect against subacute CPF induced neurotoxicity than if
applied individually. Environ Toxicol 31: 1017–26, 2016.
41 Brain Res Bull. 2015 Jan; 110:54–67. doi: 10.1016/j.brainresbull.2014.12.003. Epub 2014 Dec 12. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Prenatal nicotine alters the developmental neurotoxicity of postnatal chlorpyrifos directed toward additional useful data on low dose chlorpyrifos effects.
cholinergic systems: better, worse, or just "different?”.
However, the study does illustrate the potentially important
Slotkin TA, Seidler FJ. interaction between nicotine and chlorpyrifos in terms of
This study examines whether prenatal nicotine exposure sensitizes the developing brain to neurodevelopment. Nicotine contained in tobacco smoke
subsequent developmental neurotoxicity evoked by chlorpyrifos, a commonly-used insecticide. We produces abnormalities of brain development leacceptable
gave nicotine to pregnant rats throughout gestation at a dose (3mg/kg/day) producing plasma levels daily intakeng to neurobehavioural deficits such as attention
typical of smokers; offspring were then given chlorpyrifos on postnatal days 1–4, at a dose (1mg/kg) deficit/hyperactivity disorder, conduct disorder and affective
that produces minimally-detectable inhibition of brain cholinesterase activity. We evaluated indices disorders ie the effects of nicotine exposure overlap the
for acetylcholine (ACh) synaptic function throughout adolescence, young adulthood and later claimed neurodevelopmental effects of low dose chlorpyrifos
adulthood, in brain regions possessing the majority of Ach projections and cell bodies; we measured exposure in children.
nicotinic ACh receptor binding, hemicholinium-3 binding to the presynaptic choline transporter and ‘prenatal nicotine administration alters the subsequent
choline acetyltransferase activity, all known targets for the adverse developmental effects of nicotine developmental neurotoxicity of chlorpyrifos in a complex
and chlorpyrifos given individually. By itself nicotine elicited overall upregulation of the ACh markers, manner, comprising sex-specific, regional and temporal
albeit with selective differences by sex, region and age. Likewise, chlorpyrifos alone had highly sex- effects on the expression of the synaptic proteins that most
selective effects. Importantly, all the effects showed temporal progression between adolescence and closely delineate ACh function. Despite that complexity, there
adulthood, pointing to ongoing synaptic changes rather than just persistence after an initial injury. was a clear overall pattern. The combined exposure always
Prenatal nicotine administration altered the responses to chlorpyrifos in a consistent pattern for all lowered values relative to those of the individual treatments
three markers, lowering values relative to those of the individual treatments or to those expected or to the values expected from simple additive effects,
from simple additive effects of nicotine and chlorpyrifos. The combination produced global producing a unique, global interference with emergence of
interference with emergence of the ACh phenotype, an effect not seen with nicotine or chlorpyrifos the ACh phenotype, an effect not seen with nicotine or
alone. Given that human exposures to nicotine and chlorpyrifos are widespread, our results point to chlorpyrifos alone. Depending upon sex, age and region,
the creation of a subpopulation with heightened vulnerability. nicotine or chlorpyrifos by themselves evoked both increases
100 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

and decreases in expression of nAChRs, the high-affinity


presynaptic choline transporter and ChAT. Consequently,
where an individual treatment evoked an increase in a given
parameter, the superimposed reduction from combined
treatment appeared to “protect” by restoring those values
closer to normal. However, our results show quite clearly that
this is probably not protection against the initial brain injury,
but rather a failure of adaptive responses: in every case, the
effects on ACh parameters showed progressive changes
between adolescence and adulthood, not persistence of an
initial effect.’
42 Brain Res Bull. 2015 Feb; 111:84–96. doi: 10.1016/j.brainresbull.2015.01.003. Epub 2015 Jan 12. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Prenatal nicotine changes the response to postnatal chlorpyrifos: Interactions targeting serotonergic additional useful data on low dose chlorpyrifos neurological
synaptic function and cognition. effects.
Slotkin TA, Skavicus S, Levin ED, Seidler FJ. However, the study does illustrate the potentially important
Nicotine and chlorpyrifos are developmental neurotoxicants that target serotonin systems. We interaction between nicoine and chlorpyrifos in terms of
examined whether prenatal nicotine exposure alters the subsequent response to chlorpyrifos given neurodevelopment.
postnatally. Pregnant rats received nicotine throughout gestation at 3mg/kg/day, a regimen designed ‘…….that prenatal nicotine treatment changes the response
to achieve plasma levels seen in smokers; chlorpyrifos was given to pups on postnatal days (PN) 1– of 5HT systems to subsequent chlorpyrifos exposure,
4 at 1mg/kg, just above the detection threshold for brain cholinesterase inhibition. We assessed specifically leacceptable daily intakeng to impairment of
long-term effects from adolescence (PN30) through full adulthood (PN150), measuring the adaptive increases in 5HT receptors and presynaptic activity;
expression of serotonin receptors and serotonin turnover (index of presynaptic impulse activity) in hence the combined treatment produces a global interference
cerebrocortical brain regions encompassing the projections that are known targets for nicotine and with 5HT synaptic function not seen with either agent alone.
chlorpyrifos. Nicotine or chlorpyrifos individually increased the expression of serotonin receptors, Because the unique effect is superimposed on the direct
with greater effects on males than on females and with distinct temporal and regional patterns actions of each toxicant, the smaller upregulation of 5HT
indicative of adaptive synaptic changes rather than simply an extension of initial injury. This receptors could be mistaken for partial neuroprotection by
interpretation was confirmed by our finding an increase in serotonin turnover, connoting presynaptic nicotine, but evaluation of presynaptic activity and of the
serotonergic hyperactivity. Animals receiving the combined treatment showed a reduction in these temporal, regional and sex-dependent patterns indicates
adaptive effects on receptor binding and turnover relative to the individual agents, or even an effect otherwise. This mirrors our conclusions for the effects of
in the opposite direction; further, normal sex differences in serotonin receptor concentrations were nicotine on the developmental neurotoxicity of chlorpyrifos
dissipated or reversed, an effect that was confirmed by behavioural evaluations in the Novel directed toward cholinergic pathways (Slotkin and Seidler,
Objection Recognition Test. In addition to the known liabilities associated with maternal smoking 2015) and resembles our findings for prenatal
during pregnancy, our results point to additional costs in the form of heightened vulnerability to dexamethasone treatment, which likewise interfered with
APPENDIX 2 101

Summary
Study and published abstract Study evaluation
No.

neurotoxic chemicals encountered later in life. adaptive responses to chlorpyrifos, thus augmenting the
deterioration of synaptic function and behavioural
performance (Levin et al, 2014; Slotkin et al, 2013, 2014). In
addition to the known liabilities associated with maternal
smoking during pregnancy, our results point to additional
costs in the form of heightened vulnerability to neurotoxic
chemicals encountered later in life.’
43 Toxicology. 2015 Dec 2; 338:8–16. doi: 10.1016/j.tox.2015.09.005. Epub 2015 Sep 28. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Prenatal drug exposures sensitize noradrenergic circuits to subsequent disruption by chlorpyrifos. additional useful data on low dose chlorpyrifos neurological
Slotkin TA, Skavicus S, Seidler FJ. effects.

We examined whether nicotine or dexamethasone, common prenatal drug exposures, sensitize the However, the study does illustrate the potentially important
developing brain to chlorpyrifos. We gave nicotine to pregnant rats throughout gestation at a dose interactions between nicotine or dexamethasone (common
(3mg/kg/day) producing plasma levels typical of smokers; offspring were then given chlorpyrifos on prenatal drug exposures) and chlorpyrifos and
postnatal days 1–4, at a dose (1mg/kg) that produces minimally-detectable inhibition of brain neurodevelopment.
cholinesterase activity. In a parallel study, we administered dexamethasone to pregnant rats on
gestational days 17–19 at a standard therapeutic dose (0.2mg/kg) used in the management of
preterm labor, followed by postnatal chlorpyrifos. We evaluated cerebellar noradrenergic
projections, a known target for each agent, and contrasted the effects with those in the cerebral
cortex. Either drug augmented the effect of chlorpyrifos, evidenced by deficits in cerebellar β-
adrenergic receptors; the receptor effects were not due to increased systemic toxicity or
cholinesterase inhibition, nor to altered chlorpyrifos pharmacokinetics. Further, the deficits were not
secondary adaptations to presynaptic hyperinnervation/hyperactivity, as there were significant
deficits in presynaptic norepinephrine levels that would serve to augment the functional
consequence of receptor deficits. The pretreatments also altered development of cerebrocortical
noradrenergic circuits, but with a different overall pattern, reflecting the dissimilar developmental
stages of the regions at the time of exposure. However, in each case the net effects represented a
change in the developmental trajectory of noradrenergic circuits, rather than simply a continuation of
an initial injury. Our results point to the ability of prenatal drug exposure to create a subpopulation
with heightened vulnerability to environmental neurotoxicants.
44 Biomed Pharmacother. 2017 Apr; 88:844–51. doi: 10.1016/j.biopha.2017.01.105. Epub 2017 Feb 4. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Immunoprotective activity and antioxidant properties of cactus (Opuntia ficus indica) extract against additional useful data on low dose chlorpyrifos neurological
chlorpyrifos toxicity in rats.
102 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

Smida A, Ncibi S, Taleb J, Ben Saad A, Ncib S, Zourgui L. effects.


Opuntia ficus indica (family Cactaceae) is a typical Mediterranean plant, mainly used in food and
tracceptable daily intaketional folk medicine. The present study was designed to evaluate the
protective effect of Opuntia ficus indica extract against chlorpyrifos (CPF)-induced immunotoxicity in
rats. The experimental animals consisted of four groups of Wistar rats (5-6 weeks old) of eight each:
a control group, a group treated with CPF (10mg/kg), a group treated with Opuntia ficus indica
extract (100mg/kg), and a group treated with cactus extract then treated with CPF. These
components were daily administered by gavage for 30 days. After treatment, immunotoxicity was
estimated by a count of thymocytes, splenocytes, stem cells in the bone marrow, relative weights of
thymus and spleen, DNA aspects, and oxidative stress status in these organs. Results showed that
CPF could induce thymus atrophy, splenomegaly, and a decrease in the cell number in the bone
marrow. It also increased the oxidative stress markers resulting in elevated levels of the lipid
peroxidation with a concomitant decrease in the levels of enzymatic antioxidants (SOD, CAT, GPx)
in both spleen and thymus, and also degradation of thymocyte and splenocyte DNA. Consistent
histological changes were found in the spleen and thymus under CPF treatment. However,
administration of Opuntia ficus indica extract was found to alleviate this CPF-induced damage.
45 Pol J Vet Sci. 2016; 19(2):371–8. doi: 10.1515/pjvs-2016-0046. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Concentration of hepatic vitamins A and E in rats exposed to chlorpyrifos and/or enrofloxacin. additional useful data on low dose chlorpyrifos neurological
Spodniewska A, Barski D. effects.

The aim of the study was to determine the level of antioxidant vitamins A and E in the liver of rats
exposed to chlorpyrifos and/or enrofloxacin. Chlorpyrifos (Group I) was administered at a dose of
0.04 LD50 (6 mg/kg b.w.) for 28 days, and enrofloxacin (Group II) at a dose of 5 mg/kg b.w. for 5
consecutive days. The animals of group III were given both of the mentioned above compounds at
the same manner as groups I and II, but enrofloxacin was applied to rats for the last 5 days of
chlorpyrifos exposure (ie on day 24, 25, 26, 27 and 28). Chlorpyrifos and enrofloxacin were
administered to rats intragastrically via a gastric tube. The quantitative determination of vitamins
was made by the HPLC method. The results of this study indicated a reduction in the hepatic
concentrations of vitamins A and E, compared to the control, which sustained for the entire period of
the experiment. The four-week administration of chlorpyrifos to rats resulted in a significant
decrease of vitamins in the initial period of the experiment, ie up to 24 hours after exposure. For
vitamin A the maximum drop was observed after 24 hours (19.24%) and for vitamin E after 6  hours
(23.19%). Enrofloxacin caused a slight (3–9%) reduction in the level of the analysed vitamins. In the
APPENDIX 2 103

Summary
Study and published abstract Study evaluation
No.

chlorpyrifos-enrofloxacin co-exposure group reduced vitamins A and E levels were also noted, but
changes in this group were less pronounced in comparison to the animals intoxicated with
chlorpyrifos only. The decrease in the antioxidant vitamin levels, particularly noticeable in the
chlorpyrifos—and the chlorpyrifos combined with enrofloxacin-treated groups, may result not only
from the increase in the concentration of free racceptable daily intakecals, but also from the
intensification of the secondary stages of lipid peroxidation.
46 J Biochem Mol Toxicol. 2018 May; 32(5):e22050. doi: 10.1002/jbt.22050. Epub 2018 Apr 18. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Ameliorative effect of nanoencapsulated flavonoid against chlorpyrifos-induced hepatic oxidative additional useful data on low dose chlorpyrifos neurological
damage and immunotoxicity in Wistar rats. effects.
Suke SG, Sherekar P, Kahale V, Patil S, Mundhada D, Nanoti VM.
The theme of the present work is to evaluate the protective effect of nanoencapsulated quercetin
(NEQ) against chlorpyrifos (CPF)-induced hepatic damage and immune alterations in animals.
Nanoparticles (NP) drug encapsulation was prepared. 40 male Wistar rats were divided into eight
groups. Two groups served as control and CPF (13.5 mg/kg) treatment for 28 days. Other three
groups were free quercetin (QC), NP and NEQ treated with 3 mg/kg respectively for 15 days;
whereas remaining three groups received treatment of CPF and QC, NP, NEQ, respectively, for 15
days. The results show that significantly altered oxidative stress in the liver tissue and liver enzyme
parameters in blood and immune responses in CPF-treated rats compared to controls.
Administration of NEQ attenuated biochemical and immunological parameters. The liver
histopathological analysis confirmed pathological improvement. Hence, use of NEQ appeared to be
beneficial to a great extent in attenuating and restoring hepatic oxidative damage and immune
alteration sustained by pesticide exposure.
47 Hum Exp Toxicol. 2016 Sep; 35(9):991–1004. doi: 10.1177/0960327115614384. Epub 2015 Oct 30. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
A model of chlorpyrifos distribution and its biochemical effects on the liver and kidneys of rats. additional useful data on low dose chlorpyrifos neurological
Tanvir EM, Afroz R, Chowdhury M, Gan SH, Karim N, Islam MN, Khalil MI. effects.

This study investigated the main target sites of chlorpyrifos (CPF), its effect on biochemical indices,
and the pathological changes observed in rat liver and kidney function using gas
chromatography/mass spectrometry. Adult female Wistar rats (n = 12) were randomly assigned into
two groups (one control and one test group; n = 6 each). The test group received CPF via oral
gavage for 21 days at 5 mg/kg daily. The distribution of CPF was determined in various organs
(liver, brain, heart, lung, kidney, ovary, acceptable daily intakepose tissue, and skeletal muscle),
104 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

urine and stool samples using GCMS. Approximately 6.18% of CPF was distributed in the body
tissues, and the highest CPF concentration (3.80%) was found in acceptable daily intakepose
tissue. CPF also accumulated in the liver (0.29%), brain (0.22%), kidney (0.10%), and ovary
(0.03%). Approximately 83.60% of CPF was detected in the urine. CPF exposure resulted in a
significant increase in plasma transaminases, alkaline phosphatase, and total bilirubin levels, a
significant reduction in total protein levels and an altered lipid profile. Oxidative stress due to CPF
administration was also evidenced by a significant increase in liver malondialdehyde levels. The
detrimental effects of CPF on kidney function consisted of a significant increase in plasma urea and
creatinine levels. Liver and kidney histology confirmed the observed biochemical changes. In
conclusion, CPF bioaccumulates over time and exerts toxic effects on animals.
48 Toxicology. 2015 Dec 2; 338:95–103. doi: 10.1016/j.tox.2015.10.010. Epub 2015 Oct 30. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
The effect of HMGB1 on sub-toxic chlorpyrifos exposure-induced neuroinflammation in amygdala of additional useful data on low dose chlorpyrifos neurological
neonatal rats. effects.
Tian J, Dai H, Deng Y, Zhang J, Li Y, Zhou J, Zhao M, Zhao M, Zhang C, Zhang Y, Wang P, Bing G,
Zhao L.
Chlorpyrifos (CPF), one of organophosphorus pesticides (OPs), is associated with developmental
neurotoxicity. Inflammatory response is closely related with CPF-induced neurotoxicity. The present
study aimed at exploring whether sub-toxic CPF exposure on neonatal rats results in
neuroinflammation that mediated by HMGB1/TLR4/NF-κB signaling pathway in the amygdala. The
neonatal rats were subcutaneously injected with 5mg/kg CPF for 4 consecutive days (postnatal day
11–14) with or without HMGB1 inhibitor, glycyrrhizin. We assessed the levels of pro-inflammatory
cytokines at 12, 24, and 72 h after CPF exposure. The role of HMGB1 on neuroinflammation in sub-
toxic exposure during brain development was studied. CPF-treated neonatal rats exhibited a
significant increase in the expression of pro-inflammatory cytokines, such as IL-6, TNF-α and
HMGB1, and a significant increase in the activation of NF-κB in the amygdala after CPF exposure.
Inhibited HMGB1 reduced the release of IL-6 and TNF-α, and inhibited activation of NF-κB. Our
findings indicate that CPF exposure on developmental brain might induce the activation of
neuroinflammation mediated by HMGB1/TLR4/NF-κB pathway in the amygdala.
49 Toxicol Ind Health. 2017 Feb; 33(2):159–70. doi: 10.1177/0748233715616553. Epub 2016 Jul 10. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
The protective role of alpha-lipoic acid on long-term exposure of rats to the combination of additional useful data on low dose chlorpyrifos neurological
chlorpyrifos and deltamethrin pesticides. effects.
APPENDIX 2 105

Summary
Study and published abstract Study evaluation
No.

Uchendu C, Ambali SF, Ayo JO, Esievo KA.


The study was aimed at evaluating the protective role of α-lipoic acid (ALA) on long-term exposure
of rats to the combination of chlorpyrifos (CPF) and deltamethrin (DLT). Forty-two (42) male Wistar
rats were divided into 6 exposure groups with 7 animals in each group: (I) soya oil (2 ml kg-1), (II)
ALA (60 mg kg-1), (III) DLT (6.25 mg kg-1), (IV) CPF (4.75 mg kg-1), (V) (CPF + DLT) DLT (6.25 mg
kg-1) and CPF (4.75 mg kg-1; 1/20th of the previously determined median lethal dose) and (VI) (ALA
+ CPF + DLT) pretreated with ALA (60 mg kg-1) and then co-exposed to CPF and DLT, 45 min later.
The regimens were administered by gavage once daily for a period of 16 weeks. Sera obtained from
blood collected at the end of the experimental period were used for the evaluation of serum glucose,
total protein, albumin, urea, creatinine and the activities of alanine aminotransferase, aspartate
aminotransferase, alkaline phosphatase, lactate dehydrogenase and acetylcholinesterase. The liver
homogenate was used to assay for the activities of superoxide dismutase and glutathione
peroxidase and the concentrations of malondialdehyde, cytokine and tumour necrotic factor α. The
result showed that the combination of CPF and DLT resulted in marked alterations of these
biochemical parameters in most cases compared to either of the pesticides singly, supplementation
with ALA ameliorated these alterations.
50 J Steroid Biochem Mol Biol. 2016 Feb; 156:1–9. doi: 10.1016/j.jsbmb.2015.10.010. Epub 2015 Oct Dosing: 0, 0.01, 1 mg/kg bw/day for 100 days. An unusual
27. vehicle (castor oil) was used. Chlorpyrifos exposure at dose
0.01 mg/kg bw/day were associated with a small, non-dose
Pesticide chlorpyrifos acts as an endocrine disruptor in adult rats causing changes in mammary related, significant (p < 0.05) increase in the number of
gland and hormonal balance. mammary ducts per histological section without any change
Ventura C, Nieto MR, Bourguignon N, Lux-Lantos V, Rodriguez H, Cao G, Randi A, Cocca C, Núñez in the number of lobar buds (lobules). Animals dosed at 1
M. mg/kg/day had an increased incidence of hyperplastic
mammary ducts cf. controls (45.6 ± 7.1% vs. 35.1 ± 8.5%,
Endocrine disruptors (EDs) are compounds that interfere with hormone regulation and influence respectively, p < 0.05). A significantly higher incidence of
mammary carcinogenesis. We have previously demonstrated that the pesticide chlorpyrifos (CPF) florid duct hyperplasias were present in animals dosed at ≥
acts as an ED in vitro, since it induces human breast cancer cells proliferation through estrogen 0.01 mg/kg/day cf. the control group. Animals dosed at 0.01
receptor alpha (ERα) pathway. In this work, we studied the effects of CPF at environmental doses mg/kg bw/day had a significantly (p < 0.05) increased
(0.01 and 1mg/kg/day) on mammary gland, steroid hormone receptors expression and serum steroid incidence of lobular adenosis cf. controls; however no such
hormone levels. It was carried out using female Sprague-Dawley 40-days-old rats exposed to the changes occurred in the high dose group (ie the finding was
pesticide during 100 days. We observed a proliferating ductal network with a higher number of ducts not dose related). Animals dosed at 1 mg/kg bw/day had
and alveolar structures. We also found an increased number of benign breast diseases, such as lower serum estracceptable daily intakeol, progesterone and
hyperplasia and adenosis. CPF enhanced progesterone receptor (PgR) along with the proliferating testosterone levels cf. controls (p < 0.05). Serum LH levels
cell nuclear antigen (PCNA) in epithelial ductal cells. On the other hand, the pesticide reduced the were significantly (p < 0.05) reduced cf. controls following
106 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

expression of co-repressors of estrogen receptor activity REA and SMRT and it decreased serum dosing at ≥ 0.01 mg/kg bw/day. No effects on serum FSH
estracceptable daily intakeol (E2), progesterone (Pg) and luteinizing hormone (LH) levels. Finally, levels were detected. Dosing at 0.01 mg/kg bw/day (but not
we found a persistent decrease in LH levels among ovariectomized rats exposed to CPF. Therefore, at higher doses ie a dose response was not detected)
CPF alters the endocrine balance acting as an ED in vivo. These findings warn about the harmful resulted in reduced nuclear receptor co-repressor 2 protein
effects that CPF exerts on mammary gland, suggesting that this compound may act as a risk factor expression in mammary gland.
for breast cancer.
Overall the study provides some limited evidence of possible
endocrine effects in rats. However, consistent dose
responses and a biologically coherent pattern of events was
not fully established by this study. No historical control data
was provided for any of the measurements performed.
Furthermore the statistical power of the study was very
limited (n =6). Thus the results of this study are not
consistent with the available high quality regulatory data.
Further confirmatory data is required in order for this study to
be of regulatory value.
51 J Steroid Biochem Mol Biol. 2018 Oct 2. pii: S0960-0760(18)30187-0. doi: The statistical power of this study (n = 6) is considered to be
10.1016/j.jsbmb.2018.09.021. [Epub ahead of print] inadequate. The minimum n for a study of this type is 50
females/dose at the commencement of the study with at least
Effects of the pesticide chlorpyrifos on breast cancer disease. Implication of epigenetic 20 animals survival to study completion. The study is not
mechanisms. regarded as being of regulatory value.
Ventura C, Zappia CD, Lasagna M, Pavicic W, Richard S, Bolzan AD, Monczor F, Núñez M, Cocca
C.
Chlorpyrifos (CPF) is an organophosphorus pesticide used for agricultural pest control all over the
world. We have previously demonstrated that environmental concentrations of this pesticide alter
mammary gland histological structure and hormonal balance in rats chronically exposed. In this
work, we analyzed the effects of CPF on mammary tumors development. Our results demonstrated
that CPF increases tumor incidence and reduces latency of NMU-induced mammary tumors.
Although no changes were observed in tumor growth rate, we found a reduced steroid hormone
receptor expression in the tumors of animals exposed to the pesticide. Moreover, we analyzed the
role of epigenetic mechanisms in CPF effects. Our results indicated that CPF alters HDAC1 mRNA
expression in mammary gland, although no changes were observed in DNA methylation. In
summary, we demonstrate that the exposure to CPF promotes mammary tumors development with a
reduced steroid receptors expression. It has also been found that CPF affects HDAC1 mRNA levels
APPENDIX 2 107

Summary
Study and published abstract Study evaluation
No.

in mammary tissue pointing that CPF may act as a breast cancer risk factor.
52 Toxicology. 2018 Dec 1; 410:106–15. doi: 10.1016/j.tox.2018.09.008. Epub 2018 Sep 17. CPF was dissolved in DMSO to provide consistent absorption
and was injected subcutaneously in a volume of 1 mL/kg
Evaluation of the effects of chlorpyrifos combined with lipopolysaccharide stress on body weight at a dose of 2.5 mg/kg given daily at PND 11–14.
neuroinflammation and spatial memory in neonatal rats. The route of exposure is of limited regulatory relevance. The
Wang P, Dai H, Zhang C, Tian J, Deng Y, Zhao M, Zhao M, Bing G, Zhao L. dose and likely systemic exposure far exceeds the NOEL for
inhibition of blood cholinesterases in rats. The study provides
Chlorpyrifos (CPF) may weaken the immune defenses of children, making them vulnerable to no additional useful data on low dose chlorpyrifos effects.
opportunistic bacterial infection. CPF combined with bacterial infection is a potential problem for
children during their childhood development. However, there is a lack of studies on the joint effects
of these two factors on children. Here, we assessed the effects of CPF combined with
lipopolysaccharide (LPS) on the inflammation and development of the nervous system. In this study,
the cell toxicity of CPF plus LPS in cultured astrocytes, and the pathogenic effects of CPF plus LPS
in neonatal rat models were observed. The hydrogen (H2)-inhalation was used for treatment to
explore its therapeutic potential. We found that CPF plus LPS activated the astrocyte, which
increased the expressions of HMGB1, TLR4, and p-NF-κB p65, while H2-inhalation reduced the
expressions (p < 0.05). We also found that CPF plus LPS induced long-lasting spatial memory
deficits throughout brain maturation. However, H2-inhalation improved rat performance in these
behavioural experiments (p < 0.05). In conclusion, the sub-toxic concentration of CPF did not cause
a significant damage in short term, but induced a severe long-term damage to the brain when
combined with LPS. H2-inhalation reduced the neuronal damage and behavioural abnormalities
caused by CPF and LPS exposure.
53 Environ Toxicol. 2017 Jul; 32(7):1927–36. doi: 10.1002/tox.22415. Epub 2017 Mar 15. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Cadmium and chlorpyrifos inhibit cellular immune response in spleen of rats. additional useful data on low dose chlorpyrifos effects.
Wang P, Wang J, Sun YJ, Yang L, Wu YJ.
Cadmium (Cd) and chlorpyrifos (CPF) are common pollutants coexisting in the environment, and
both of them have been reported to have immunotoxicity to organisms. However, the joint effects of
these two chemicals on the immune system are still unknown. In this study, we used CdCl2 and CPF
to study their combined effects on immune functions in the spleen of rats. In in vivo experiments, SD
rats were exposed to different doses of CdCl2 (0.7 and 6 mg  kg-1 body weight/day) and CPF (1.7
and 15 mg kg-1 body weight/day) or their combinations for consecutive 28 days. The proliferation
and cytokine production ability of the splenocytes isolated from the treated animals were assessed.
In in vitro experiments, we used different concentrations of CdCl2 and CPF to treat concanavalin A
108 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

(Con A)-induced splenocytes isolated from untreated rats. We found that the combination of CPF
and high dose of CdCl2 had a synergistic inhibitory effect on production of IFN-γ by spleen cells
induced by Con A. The in vitro results showed that two chemicals had different effects on the cell
proliferation and cytokine production depending on the exposure doses and time. This result
suggests that exposure to both CdCl2 and CPF at the environmentally-relevant low dose may be
potentially more hazardous than exposure to each individual toxicant.
54 Chem Res Toxicol. 2015 Jun 15; 28(6):1216–23. doi: 10.1021/acs.chemrestox.5b00054. Epub 2015 Chlorpyrifos dose used exceeds the in vivo NOAELs for
Apr 22. cholinesterase inhibition in rats. The study provides no
additional useful data on low dose chlorpyrifos effects.
Metabolomics analysis and biomarker identification for brains of rats exposed subchronically to the
mixtures of low-dose cadmium and chlorpyrifos.
Xu MY, Sun YJ, Wang P, Xu HY, Chen LP, Zhu L, Wu YJ.
Cadmium (Cd) and chlorpyrifos (CPF) are widespread harmful environmental pollutants with
neurotoxicity to mammals. Although the exposure to Cd and CPF at the same time may pose a
significant risk to human health, the subchronic combined neurotoxicity of these two chemicals at
low levels in the brain is poorly understood. In this study, we treated rats with three doses (low,
middle, and high) of Cd, CPF, or their mixture for 90 days. No obvious symptom was observed in the
treated animals except those treated with high-dose CPF. Histological results showed that middle
and high doses of the chemicals caused neuronal cell damage in brains. GC-MS-based
metabolomics analysis revealed that energy and amino acid metabolism were disturbed in the
brains of rats exposed to the two chemicals and their combinations even at low doses. We further
identified the unique brain metabolite biomarkers for rats treated with Cd, CPF, or both. Two amino
acids, tyrosine and l-leucine, were identified as the biomarkers for Cd and CPF treatment,
respectively. In addition, a set of five unique biomarkers (1,2-propanediol-1-phosphate, d-gluconic
acid, 9H-purine, serine, and 2-ketoisovaleric acid) was identified for the mixtures of Cd and CPF.
Therefore, the metabolomics analysis is more sensitive than regular clinical observation and
pathological examination for detecting the neurotoxicity of the individual and combined Cd and CPF
at low levels. Overall, these results identified the unique biomarkers for Cd and CPF exposure,
which provide new insights into the mechanism of their joint toxicity.
55 Toxicology. 2017 Jun 1; 384:50–58. doi: 10.1016/j.tox.2017.04.008. Epub 2017 Apr 19. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Metabolomic analysis for combined hepatotoxicity of chlorpyrifos and cadmium in rats. additional useful data on low dose chlorpyrifos effects.
Xu MY, Wang P, Sun YJ, Wu YJ.
APPENDIX 2 109

Summary
Study and published abstract Study evaluation
No.

Pesticides and heavy metals are widespread environmental pollutants. Although the acute toxicity of
organophosphorus pesticide chlorpyrifos (CPF) and toxic heavy metal cadmium (Cd) is well
characterized, the combined toxicity of CPF and Cd, especially the hepatotoxicity of the two
chemicals with long-term exposure at a low dose, remained unclear. In this study, we investigated
the toxicity in the liver of rats upon subchronic exposure to CPF and Cd at environmentally relevant
doses. Rats were given three different doses (1/135 LD50, 1/45 LD50 and 1/15 LD50) of CPF and
Cd as well as their mixtures by oral gavage for 90days. After treatment, the liver tissues were
subjected to histopathological examination and biochemical analysis. Gas chromatography-mass
spectrometry (GC-MS) was used to analyze the metabolomic changes in the rat liver upon CPF, Cd
and their mixtures treatment. The results showed that CPF and Cd-induced oxidative damage and
disrupted energy, amino acid, and fatty acid metabolism in the liver. Eleven biomarkers in liver were
identified for CPF-, Cd-, and their mixture-treated rats. Three metabolites, ie butanedioic acid, myo-
inositol, and urea, were identified as unique biomarkers for the mixture-treated rats. Moreover, we
found that Cd could accelerate the metabolism of CPF in the liver when given together to the rats,
which may lead to the potential antagonistic interaction between CPF and Cd. In conclusion, our
results indicated that even at environmentally relevant doses, CPF and Cd could disrupt the liver
metabolism. In addition, the accelerated metabolism of CPF by Cd may lead to their potential
antagonistic interaction.
56 Toxicology. 2017 Jun 1; 384:50–58. doi: 10.1016/j.tox.2017.04.008. Epub 2017 Apr 19. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Metabolomic analysis for combined hepatotoxicity of chlorpyrifos and cadmium in rats. additional useful data on low dose chlorpyrifos effects.
Xu MY, Wang P, Sun YJ, Wu YJ.
Pesticides and heavy metals are widespread environmental pollutants. Although the acute toxicity of
organophosphorus pesticide chlorpyrifos (CPF) and toxic heavy metal cadmium (Cd) is well
characterized, the combined toxicity of CPF and Cd, especially the hepatotoxicity of the two
chemicals with long-term exposure at a low dose, remained unclear. In this study, we investigated
the toxicity in the liver of rats upon subchronic exposure to CPF and Cd at environmentally relevant
doses. Rats were given three different doses (1/135 LD50, 1/45 LD50 and 1/15 LD50) of CPF and
Cd as well as their mixtures by oral gavage for 90days. After treatment, the liver tissues were
subjected to histopathological examination and biochemical analysis. Gas chromatography-mass
spectrometry (GC-MS) was used to analyze the metabolomic changes in the rat liver upon CPF, Cd
and their mixtures treatment. The results showed that CPF and Cd-induced oxidative damage and
disrupted energy, amino acid, and fatty acid metabolism in the liver. Eleven biomarkers in liver were
identified for CPF-, Cd-, and their mixture-treated rats. Three metabolites, ie, butanedioic acid, myo-
110 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

inositol, and urea, were identified as unique biomarkers for the mixture-treated rats. Moreover, we
found that Cd could accelerate the metabolism of CPF in the liver when given together to the rats,
which may lead to the potential antagonistic interaction between CPF and Cd. In conclusion, our
results indicated that even at environmentally relevant doses, CPF and Cd could disrupt the liver
metabolism. In addition, the accelerated metabolism of CPF by Cd may lead to their potential
antagonistic interaction.
57 Food Chem Toxicol. 2017 May; 103:246–52. doi: 10.1016/j.fct.2017.03.013. Epub 2017 Mar 9. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Joint toxicity of chlorpyrifos and cadmium on the oxidative stress and mitochondrial damage in additional useful data on low dose chlorpyrifos effects.
neuronal cells.
Xu MY, Wang P, Sun YJ, Yang L, Wu YJ.
Pesticides and heavy metals can be easily biomagnified in food chains and bioaccumulated in
individuals, thus pose significant threat to human health. However, their joint toxicity for long-term
exposure at low dose has not been thoroughly investigated. In the present study, we investigated
the oxidative damages in brain of rats exposed subchronically to organophosphorus pesticide
chlorpyrifos (CPF) and heavy metal cadmium (Cd), and their mixtures at the environmentally
relevant doses. Rats were given different doses of CPF and Cd by oral gavage for three months.
After treatment, brain tissues were subjected for biochemical analysis. Mitochondrial damage and
reactive oxidative species were also measured in neuroblastoma SH-SY5Y cells treated with CPF,
Cd and their mixtures. The results showed that CPF and Cd generated protein and lipid
peroxidation, disturbed the total antioxidant capability, and altered mitochondria ultrastructure in the
brain. Lipids and proteins were sensitive to the oxidative damage induced by CPF and Cd. CPF and
Cd decreased mitochondrial potential and induced reactive oxygen species in SH-SY5Y cells.
However, the mixture did not display higher toxicity than the sum of that of the individual treatments.
Thus, CPF and Cd could have a potential antagonistic interaction on the induction of oxidative
stress.
58 Environ Sci Pollut Res Int. 2018 Jun; 25(16):15616–29. doi: 10.1007/s11356-018-1776-x. Epub 2018 Chlorpyrifos dose used exceeds the in vivo NOAELs for
Mar 23. cholinesterase inhibition in rats. Furthermore the study was
conducted at a dose that produced overt anticholinesterase-
Assessment of neurohepatic DNA damage in male Sprague-Dawley rats exposed to induced acute neurotoxicity that likely included respiratory
organophosphates and pyrethroid insecticides. compromise. The study provides no additional useful data on
Yahia D, Ali MF. low dose chlorpyrifos effects.

The current work was undertaken to test the genotoxic potential of chlorpyrifos (CPF), dimethoate,
APPENDIX 2 111

Summary
Study and published abstract Study evaluation
No.

and lambda cyhalothrin (LCT) insecticides in rat brain and liver using the single cell gel
electrophoresis (comet assay). Three groups of adult male Sprague-Dawley rats were exposed
orally to one third LD50of CPF, dimethoate, or LCT for 24 and 48 h while the control group received
corn oil. Serum samples were collected for estimation of malondialdehyde (MDA) and glutathione
peroxidase (GPx); the brain and liver samples were used for comet assay and for histopathological
examination. Results showed that signs of neurotoxicity appeared clinically as backward stretching
of hind limb and splayed gait in dimethoate and LCT groups, respectively. CPF, LCT, and dimethoate
induced oxidative stress indicated by increased MDA and decreased GPx levels. CPF and LCT
caused severe DNA damage in the brain and liver at 24 and 48 h indicated by increased percentage
of DNA in tail, tail length, tail moment, and olive tail moment. Dimethoate induced mild DNA damage
in the brain and liver at 48 h. Histopathological changes were observed in the cerebrum, cerebellum,
and liver of exposed rats. The results concluded that CPF, LCT, and dimethoate insecticides induced
oxidative stress and DNA damage associated with histological changes in the brain and liver of
exposed rats.
59 Environ Toxicol. 2017 Sep; 32(9):2191–202. doi: 10.1002/tox.22432. Epub 2017 Jun 1. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Protective effects of Ziziphora tenuior extract against chlorpyrifos induced liver and lung toxicity in additional useful data on low dose chlorpyrifos effects.
rat: Mechanistic approaches in subchronic study.
Yazdinezhad A, Abbasian M, Hojjat Hosseini S, Naserzadeh P, Agh-Atabay AH, Hosseini MJ.
Chlorpyrifos (CPF) is one of the most widely used organophosphorus, which has spurred renewed
interest. This study was conducted to investigate the protective effect of ziziphora tenuior extract
against CPF-induced liver and lung toxicity. This study conducted 8-week rat sub-chronic toxicity
study and then the effect of ziziphora tenuior extract in 3 different doses (40, 80, 160 mg/kg) was
determined. We administrated maximum tolerated dose of CPF (6.75 mg/kg) by gavage for 8 weeks
(5 times in week) to male rats. Rats were sacrificed 24 h after last dose and the biochemical
analysis, which confirms involvement of oxidative stress in the pathogenesis of CPF toxicity in liver
including increased in lipid peroxidation, protein carbonyl content, and ROS formation, glutathione
depletion, decreased of antioxidant effect via frap oxidation and cytochrome c expulsion. In addition,
pathological lesions confirm the dysfunction of the organs (liver and lung). In addition, using of
ziziphora extract as an antioxidant is resulted in amelioration of oxidative stress marker in liver and
lung damage. In conclusion, the current study revealed that CPF toxicity is related to oxidative
stress and induction of cell death signaling and cotreatment with ziziphora extract is recommended
in the routine therapy for the protection against CPF induced liver and lung tissue damage.
112 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

60 Toxicology. 2015 Oct 2; 336:17–25. doi: 10.1016/j.tox.2015.07.014. Epub 2015 Jul 26. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Neonatal chlorpyrifos exposure induces loss of dopaminergic neurons in young adult rats. additional useful data on low dose chlorpyrifos effects.
Zhang J, Dai H, Deng Y, Tian J, Zhang C, Hu Z, Bing G, Zhao L.
Increasing epidemiological and toxicological evidence suggests that pesticides and other
environmental exposures may be associated with the development of Parkinson's disease (PD).
Chlorpyrifos (CPF) is a widely used organophosphorous pesticide with developmental neurotoxicity.
Its neurotoxicity, notably on the monoamine system, suggests that exposure of CPF may induce
dopaminergic neuronal injury. We investigated whether neonatal exposure to CPF contributes to
initiation and progression of dopaminergic neurotoxicity and explored the possible underlying
mechanisms. The newborn rats were administrated 5 mg/kg CPF subcutaneously from postnatal day
(PND) 11 to PND 14 daily. The effect of CPF on dopaminergic neurons, microglia, astrocyte, nuclear
factor-κB (NF-κB) p. 65 and p. 38 mitogen-activated protein kinase (MAPK) signaling pathways was
analyzed in the substantia nigra of rats at 12h, 24h, 72 h, 16d and 46 d after exposure. CPF-treated
rats exhibited significant reduction of dopaminergic neurons at 16d and 46 d after exposure, and a
significant increase in the expression of microglia and astrocytes in the substantia nigra after CPF
exposure. Intense activation of NF-κB p. 65 and p. 38 MAPK inflammatory signaling pathways was
observed. Our findings indicate that neonatal exposure to CPF may induce long-term dopaminergic
neuronal damage in the substantia nigra mediated by the activation of inflammatory response via
NF-κB p. 65 and p. 38 MAPK pathways in the nigrostriatal system.
61 Chemosphere. 2016 Jun; 153:287–93. doi: 10.1016/j.chemosphere.2016.03.055. Epub 2016 Mar 26. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. The study provides no
Effects of chlorpyrifos on the gut microbiome and urine metabolome in mouse (Mus musculus). additional useful data on low dose chlorpyrifos effects. Given
Zhao Y, Zhang Y, Wang G, Han R, Xie X. chlorpyrifos’s mode of action and the importance of
cholinergic signalling in the control of gut motility and other
In this study, the toxic effects of clorpyrifos (CPF) on the gut microbiome and related urine gut functions, it is hardly surprising that neurotoxic doses of
metabolome in mouse (Mus musculus) were investigated. Mice were exposed to a daily dose of 1 chlorpyrifos result in changes in gut microflora.
mg kg(-1) bodyweight of CPF for 30 d. As a result, CPF significantly altered the gut microbiota
composition in terms of the relative abundance of key microbes. Meanwhile, CPF exposure induced
the alterations of urine metabolites related to the metabolism of amino acids, energy, short-chain
fatty acids (SCFAs), phenyl derivatives and bile acids. High correlations were observed between
perturbed gut microbiome and altered metabolic profiles. These perturbations finally resulted in
intestinal inflammation and abnormal intestinal permeability, which were also confirm by the
histologic changes in colon and remarkable increase of lipopolysaccharide (LPS) and diamine
APPENDIX 2 113

Summary
Study and published abstract Study evaluation
No.

oxidase (DAO) in the serum of CPF-treated mice. Our findings will provide a new perspective to
reveal the mechanism of CPF toxicity.
62 Neurotox Res. 2018 Feb; 33(2):247–58. doi: 10.1007/s12640-017-9823-9. Epub 2017 Oct 3. None of these studies are regarded as being of regulatory
quality due to inadequacies of the statistical methods used,
Developmental Exposure to Pesticides Alters Motor Activity and Coordination in Rats: Sex lack of blinding of the observer and/or other techniques to
Differences and Underlying Mechanisms. reduce observer bias, lack of assay validation by
Gómez-Giménez B, Felipo V, Cabrera-Pastor A, Agustí A, Hernández-Rabaza V, Llansola M. incorporation of appropriate control groups and lack of
historical control data. Because of the unreliability of the
It has been proposed that developmental exposure to pesticides contributes to increasing study a NOAEL was not established.
prevalence of neurodevelopmental disorders in children, such as attention deficit with hyperactivity
(ADHD) and to alterations in coordination skills. However, the mechanisms involved in these
alterations remain unclear. We analyzed the effects on spontaneous motor activity and motor
coordination of developmental exposure to a representative pesticide of each one of the four main
chemical families: organophosphates (chlorpyrifos), carbamates (carbaryl), organochlorines
(endosulfan), and pyrethroids (cypermethrin). Pesticides were administered once a day orally, in a
sweet jelly, from gestational day 7 to post-natal day 21. Spontaneous motor activity was assessed
by an actimeter and motor coordination using the rotarod, when rats were adults. The effects were
analyzed separately in males and females. Extracellular GABA in cerebellum and NMDA receptor
subunits in hippocampus were assessed as possible underlying mechanisms of motor alterations.
Motor coordination was impaired by developmental exposure to endosulfan, cypermethrin, and
chlorpyrifos in females but not in males. The effect of endosulfan and cypermethrin would be due to
increased extracellular GABA in cerebellum, which remains unaltered in male rats. Chlorpyrifos
increased motor activity in males and females. Cypermethrin decreased motor activity mainly in
males. In male rats, but not in females, expression of the NR2B subunit of NMDA receptor in
hippocampus correlated with motor activity. These results show sex-specific effects of different
pesticides on motor activity and coordination, associated with neurotransmission alterations. These
data contribute to better understand the relationship between developmental exposure to the main
pesticide families and motor disorders in children.
63 Food Chem Toxicol. 2017 Jan; 99:135–48. doi: 10.1016/j.fct.2016.11.028. Epub 2016 Nov 29. None of these studies are regarded as being of regulatory
quality due to inadequacies of the statistical methods used,
Sex-dependent effects of developmental exposure to different pesticides on spatial learning. The lack of blinding of the observer and/or other techniques to
role of induced neuroinflammation in the hippocampus. reduce observer bias, lack of assay validation by
Gómez-Giménez B, Llansola M, Hernández-Rabaza V, Cabrera-Pastor A, Malaguarnera M, Agusti A, incorporation of appropriate control groups and lack of
Felipo V. historical control data. Because of the unreliability of the
114 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

The use of pesticides has been associated with impaired neurodevelopment in children. The aims of study a NOAEL was not established.
this work were to assess: 1) the effects on spatial learning of developmental exposure to pesticides
2) if the effects are sex-dependent and 3) if hippocampal neuroinflammation is associated with the
impairment of spatial learning. We analyzed the effects of developmental exposure to four
pesticides: chlorpyrifos, carbaryl, endosulfan and cypermethrin. Exposure was from gestational day
7 to post-natal day 21 and spatial learning and memory was assessed when the rats were young
adults. The effects of pesticides on spatial learning were pesticide and gender-dependent. Carbaryl
did not affect spatial learning in males or females. Endosulfan and chlorpyrifos impaired learning in
males but not in females. Cypermethrin improved spatial learning in the Morris water maze both in
males and females while impaired learning in the racceptable daily intakeal maze only in males.
Spatial learning ability was lower in control female rats than in males. All pesticides induced
neuroinflammation, increasing IL-1b content in the hippocampus and there is a negative correlation
between IL-1b levels in the hippocampus and spatial learning. Neuroinflammation would contribute
to the effects of pesticides on spatial learning.
APPENDIX 2 115

Summary
Study and published abstract Study evaluation
No.

64 Environ Health. 2018 Nov 16; 17(1):77. doi: 10.1186/s12940-018-0421-y.


Safety of Safety Evaluation of Pesticides: developmental neurotoxicity of chlorpyrifos and
chlorpyrifos-methyl.
Mie A, Rudén C, Grandjean P.
Authorization of pesticides for market release requires toxicity testing on animals, typically
performed by test laboratories on contract with the pesticide producer. The latter provides the
results and summary to the regulatory authorities. For the commonly used pesticide chlorpyrifos, an
industry-funded toxicity study concludes that no selective effects on neurodevelopment occur even
at high exposures. In contrast, the evidence from independent studies points to adverse effects of
current exposures on cognitive development in children. We reviewed the industry-funded
developmental neurotoxicity test data on chlorpyrifos and the related substance chlorpyrifos-methyl.
We noted treatment-related changes in a brain dimension measure for chlorpyrifos at all dose levels
tested, although not been reported in the original test summary. We further found issues which
inappropriately decrease the ability of the studies to reveal true effects, including a dosage regimen
that resulted in too low exposure of the nursing pups for chlorpyrifos and possibly for chlorpyrifos-
methyl, and a failure to detect any neurobehavioural effects of lead nitrate used as positive control
in the chlorpyrifos study. Our observations thus suggest that conclusions in test reports submitted by
the producer may be misleacceptable daily intakeng. This discrepancy affects the ability of
regulatory authorities to perform a valid and safe evaluation of these pesticides. The difference
between raw data and conclusions in the test reports indicates a potential existence of bias that
would require regulatory attention and possible resolution.

The claimed critical effect identified in this re-evaluation of


one part of the chlorpyrifos developmental neurotoxicity
regulatory study was an effect of cerebellum height in rat
pups at post-natal day 11 of age following maternal dosing at
≥ 0.3 mg/kg bw/day. Notably, this claimed effect occurred at a
dose that exceeds the NOEL for inhibition of blood
cholinesterases in rats.
116 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

Guinea Pig
65 Braz J Otorhinolaryngol. 2016 Jan-Feb; 82(1):11–6. doi: 10.1016/j.bjorl.2015.10.001. Epub 2015 Oct Chlorpyrifos dose used exceeds the in vivo NOAELs for
24. cholinesterase inhibition in rats. Since specific data on the
NOAELs for cholinesterase inhibition in guinea pigs is limited,
Morphological analysis of the vestibular system of guinea pigs poisoned by organophosphate. it is assumed on the basis of read across that the relevant
Cogo LA, Santos Filha VA, Murashima Ade A, Hyppolito MA, Silveira AF. thresholds for guinea pigs are similar to those in rats.

INTRODUCTION: The vestibular system is responsible for body balance. There are substances that
damage it, causing dizziness; these are termed vestibulotoxic substances. Agrochemicals have
been investigated for ototoxicity because of studies that identified dizziness as a recurrent symptom
among rural workers' complaints. OBJECTIVE: To histopathologically evaluate the vestibular system
in guinea pigs exposed to an organophosphate, and to identify the drug's effects on this system.
METHODS: Experimental clinical study. Eighteen guinea pigs were used; six of them poisoned with
the organophosphate chlorpyrifos at doses of 0.5mg/kg/day and seven of them at 1mg/kg/day; and a
control group of five guinea pigs was exposed to distilled water, all for 10 consecutive days. Later,
ciliary tufts of saccule and utricle maculae were counted by scanning electron microscopy.
RESULTS: Comparing the groups, a one-way ANOVA test for the variable ‘saccule’ (p=0.0569) and a
Kruskal-Wallis test for the variable "utricle" (p=0.8958) were performed, revealing no difference
among groups in both variables. CONCLUSION: The histopathologic analysis of the vestibular
system of guinea pigs exposed to an organophosphate showed no difference in the amount of ciliary
tufts of saccule and utricle maculae at the doses tested, although the result for the variable
"saccule" was considered borderline, showing a trend for significance.
66 Neurotoxicology. 2016 Sep; 56:17–28. doi: 10.1016/j.neuro.2016.06.008. Epub 2016 Jun 11. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. Since specific data on the
Spatial learning impairment in prepubertal guinea pigs prenatally exposed to the organophosphorus NOAELs for cholinesterase inhibition in guinea pigs is limited,
pesticide chlorpyrifos: Toxicological implications. it is assumed on the basis of read across that the relevant
Mamczarz J, Pescrille JD, Gavrushenko L, Burke RD, Fawcett WP, DeTolla LJ Jr, Chen H, Pereira thresholds for guinea pigs are similar to those in rats.
EFR, Albuquerque EX.
Exposure of the developing brain to chlorpyrifos (CPF), an organophosphorus (OP) pesticide used
extensively in agriculture worldwide, has been associated with increased prevalence of cognitive
deficits in children, particularly boys. The present study was designed to test the hypothesis that
cognitive deficits induced by prenatal exposure to sub-acute doses of CPF can be reproduced in
precocial small species. To address this hypothesis, pregnant guinea pigs were injected daily with
CPF (25mg/kg,s.c.) or vehicle (peanut oil) for 10days starting on presumed gestation day (GD) 53-
APPENDIX 2 117

Summary
Study and published abstract Study evaluation
No.

55. Offspring were born around GD 65, weaned on postnatal day (PND) 20, and subjected to
behavioural tests starting around PND 30. On the day of birth, butyrylcholinesterase (BuChE), an
OP bioscavenger used as a biomarker of OP exposures, and acetylcholinesterase (AChE), a major
molecular target of OP compounds, were significantly inhibited in the blood of CPF-exposed
offspring. In their brains, BuChE, but not AChE, was significantly inhibited. Prenatal CPF exposure
had no significant effect on locomotor activity or on locomotor habituation, a form of non-associative
memory assessed in open fields. Spatial navigation in the Morris water maze (MWM) was found to
be sexually dimorphic among guinea pigs, with males outperforming females. Prenatal CPF
exposure impaired spatial learning more significantly among male than female guinea pigs and,
consequently, reduced the sexual dimorphism of the task. The results presented here, which
strongly support the test hypothesis, reveal that the guinea pig is a valuable animal model for
preclinical assessment of the developmental neurotoxicity of OP pesticides. These findings are far
reaching as they lay the groundwork for future studies aimed at identifying therapeutic interventions
to treat and/or prevent the neurotoxic effects of CPF in the developing brain.
67 Neurotoxicology. 2015 May; 48:9–20. doi: 10.1016/j.neuro.2015.02.002. Epub 2015 Feb 19. Chlorpyrifos dose used exceeds the in vivo NOAELs for
cholinesterase inhibition in rats. Since specific data on the
Prenatal exposure of guinea pigs to the organophosphorus pesticide chlorpyrifos disrupts the NOAELs for cholinesterase inhibition in guinea pigs is limited,
structural and functional integrity of the brain. it is assumed on the basis of read across that the relevant
Mullins RJ, Xu S, Pereira EF, Pescrille JD, Todd SW, Mamczarz J, Albuquerque EX, Gullapalli RP. thresholds for guinea pigs are similar to those in rats.

This study was designed to test the hypothesis that prenatal exposure of guinea pigs to the
organophosphorus (OP) pesticide chlorpyrifos (CPF) disrupts the structural and functional integrity
of the brain. Pregnant guinea pigs were injected with chlorpyrifos (25 mg/kg, s.c.) or vehicle (peanut
oil) once per day for 10 consecutive days, starting approximately on the 50th day of gestation.
Cognitive behaviour of female offspring was examined starting at 40–45 post-natal days (PND)
using the Morris water maze (MWM), and brain structural integrity was analyzed at PND 70 using
magnetic resonance imaging (MRI) methods, including T2-weighted anatomical scans and diffusion
kurtosis imaging (DKI). The offspring of exposed mothers had significantly decreased body weight
and brain volume, particularly in the frontal regions of the brain including the striatum. Furthermore,
the offspring demonstrated significant spatial learning deficits in MWM recall compared to the
vehicle group. Diffusion measures revealed reduced white matter integrity within the striatum and
amygdala that correlated with spatial learning performance. These findings reveal the lasting effect
of prenatal exposure to CPF as well as the danger of mother to child transmission of CPF in the
environment.
118 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

Human
68 Ann Agric Environ Med. 2015; 22(2):275–80. doi: 10.5604/12321966.1152080. The dosimetry method used was dependent on back
calculating estimated absorbed daily dose based upon the
A new method for setting guidelines to protect human health from agricultural exposure by using creatinine corrected (based on the individual daily creatinine
chlorpyrifos as an example. excretion rate (g/d) calculated ‘typical’ population values
Phung DT, Connell D, Chu C. based on age, gender, height and weight of the exposed
individual) of concentration of the chlorpyrifos metabolite
INTRODUCTION AND OBJECTIVES: Guidelines set by various agencies for the control and TCP y in urine. The limitation of this approach is that while
management of chlorpyrifos cover a wide range of values reflecting difficulties in the procedures for TCP y excretion in urine dose increase with increasing
their development. To overcome these difficulties a new method to set guidelines would be exposure (ie the two variables are statistically significantly
developed. Published data derived from epidemiological investigations on human populations would correlated p < 0.05), the correlation coefficient (r; Pearson
be used to develop a dose-response relationship for chlorpyrifos allowing the calculation of product-moment correlation coefficient) between the two
threshold values which can be used as guidelines. MATERIALS AND METHOD: Data from the variables is fairly low (typically ≈ 0.6). 3 Furthermore, the
scientific literature on human populations were collected to evaluate the adverse response doses for validity of the metric is based on the reliability of using
a range of health effects. The Cumulative Frequency Distribution (CFD) for the minimum levels of ‘typical’ population means for daily creatinine excretion.
adverse effects measured in terms of the Lifetime Average Daily Dose (LADD(D)) and the Absorbed There are many factors that influence such typical population
Daily Dose for neurological (ADD(DN)) and non-neurological effects were used. RESULTS: Linear values and no attempt was made to account for the variance
regression equations were fitted to the CFD plots giving R 2 values of 0.93 and 0.86 indicating a of such measures across the human population(s) of interest.
normal distribution of the data. Using these CFD plots, the chronic and acute threshold values were
calculated at the 5% cumulative frequency level for chlorpyrifos exposure giving values at Thus while this approach provides useful supporting
0.5 µg/kg/d and 3 µg/kg/d respectively. information for the derivation of human health based
guidance values, it is only sufficiently accurate for derivation
CONCLUSIONS: Guidelines set using this technique at the values at 0.5 µg/kg/d and 3 µg/kg/d for of semi-quantitative exposure estimates (eg distinguishing
chronic and acute exposure respectively provide an alternative to the currently used biological between low, medium and high exposures).
endpoint and safety factor method.
However, this approach does provide some degree of support
for the proposed updated APVMA acceptable daily intake and
ArfD values for chlorpyrifos that are presented in this report.
69 & 70 Cortex. 2016 Jan; 74:383–95. doi: 10.1016/j.cortex.2015.09.011. Epub 2015 Oct 20. The studies evaluated male adolescent (12–21 years old)
chlorpyrifos applicators. The level of chlorpyrifos exposure in
A 10-month prospective study of organophosphorus pesticide exposure and neurobehavioural the high exposure cohort was very high over the 25 day
performance among adolescents in Egypt. chlorpyrifos application period with above the low exposure

3
Burns CJ, Garabrant D, Albers JW, Berent S, Giordani B, Haidar S, Garrison R, Richardson RJ. Chlorpyrifos exposure and biological monitoring among manufacturing workers.
Occup Environ Med. 2006 Mar; 63(3):218-20. PubMed PMID: 16497866; PubMed Central PMCID: PMC2078143.
APPENDIX 2 119

Summary
Study and published abstract Study evaluation
No.

Rohlman DS, Ismail AA, Rasoul GA, Bonner MR, Hendy O, Mara K, Wang K, Olson JR. cohort exposure for > 200 days. Applicators in the high
exposure group reported working for the Ministry of
Chlorpyrifos is an organophosphorus (OP) pesticide widely used around the world for agricultural Agriculture an average of 22 h per week.
operations. Although studies have examined exposure in children, there is limited information on
adolescents who are occupationally exposed. Furthermore, there is limited research addressing the The majority of neurobehavioural summary measures, 18 out
change in exposure patterns and outcomes across the application season. The goal of the current of 22, indicated worse performance for the high exposure
study was to examine the impact of chlorpyrifos exposure on neurobehavioural performance in group compared to the low exposure group.
adolescents before, during and after the application season. The longitudinal study was conducted
in Egypt from April 2010 to January 2011, quantifying exposure and neurobehavioural performance Many of the workers evaluated in these studies had levels of
with repeated measures prior to, during, and following the application period. At each test session, chlorpyrifos exposure associated with inhibition of blood
participants completed a neurobehavioural test battery and urine was collected for analysis of the cholinesterases. Accordingly the studies do not provide
chlorpyrifos metabolite 3,5,6-trichloro-2 pyridinol (TCP y ) (biomarker of exposure). Cumulative useful information on possible neurological effects at
urinary TCP y over the study period was used to classify participants into low (<median) and high (≥ exposures below those that inhibit blood cholinesterases.
median) exposure groups. The urinary TCP y concentrations increased for both groups during the
application season and decreased following the end of application. TCP y levels were significantly
elevated in the high exposure group compared to the low exposure groups at all time intervals
except baseline. Deficits in cumulative neurobehavioural performance were found among the high
exposure group compared with the low exposure group. Additionally, changes in neurobehavioural
performance across the application season indicate a pattern of impaired performance in the high
exposure group compared to the low exposure group. Deficits increased during the application
season and remained even months after application ceased. This study is the first to examine the
impact of changes in pesticide exposure and neurobehavioural performance not only before and
after the application season, but also within the application season. Furthermore, this study
examines the impact of pesticide exposure on an adolescent population who may be at greater risk
than adult populations.
Int J Hyg Environ Health. 2015 Mar; 218(2):203–11. doi: 10.1016/j.ijheh.2014.10.005. Epub 2014
Nov 7.
Longitudinal assessment of occupational exposures to the organophosphorous insecticides
chlorpyrifos and profenofos in Egyptian cotton field workers.
Singleton ST, Lein PJ, Dadson OA, McGarrigle BP, Farahat FM, Farahat T, Bonner MR, Fenske RA,
Galvin K, Lasarev MR, Anger WK, Rohlman DS, Olson JR.
Chlorpyrifos (CPF) and profenofos (PFF) are organophosphorus (OP) insecticides that are applied
seasonally in Egypt to cotton fields. Urinary trichloro-2-pyridinol (TCP y ), a specific CPF metabolite,
120 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

and 4-bromo-2-chlorophenol (BCP), a specific PFF metabolite, are biomarkers of exposure, while
inhibition of blood butyrylcholinesterase (BChE) and acetylcholinesterase (AChE) activities are
effect biomarkers that may be associated with neurotoxicity. Urinary TCP y and BCP and blood BChE
and AChE activities were measured in 37 adult Egyptian Ministry of Agriculture workers during and
after 9–17 consecutive days of CPF application followed by an application of PFF (9–11 days), and
a second CPF application (5 days) in 2008. During the OP applications, urinary TCP y and BCP
levels were significantly higher than baseline levels, remained elevated following the application
periods, and were associated with an exposure related inhibition of blood BChE and Ache. Analysis
of blood AChE levels before and after the PFF application period suggests that individual workers
with peak BCP levels greater than 1000 μg/g creatinine exhibited further inhibition of blood AChE
with PFF application, demonstrating that PFF exposure had a negative impact on AChE activity in
this highly exposed worker population. While large interindividual differences in exposure were
observed throughout this longitudinal study (peak urinary BCP and peak TCP y levels for individuals
ranging from 13.4 to 8052 and 16.4 to 30,107 μg/g creatinine, respectively), these urinary
biomarkers were highly correlated within workers (r=0.75, p<0.001). This suggests that the relative
exposures to CPF and PFF were highly correlated for a given worker. The variable exposures
between job classification and work site suggest that job title and work location should not be used
as the sole basis for categorizing OP exposures when assessing neurobehavioural and other health
outcomes in Egyptian cotton field workers. Together, these findings will be important in educating
the Egyptian insecticide application workers in order to encourage the development and
implementation of work practices and personal protective equipment to reduce their exposure to
CPF and PFF.
71 Environ Health Perspect. 2017 Sep 8; 125(9):097007. doi: 10.1289/EHP604. Bystander pesticide exposure assessment: Exposure was
based on California pesticide use reports (PURs). PUR data
Combined Prenatal Pesticide Exposure and Folic Acid Intake in Relation to Autism Spectrum measures pesticide uses down to 2.59 square km (1 square
Disorder. mile) units (MTRS). This data was evaluated using a spatial
Schmidt RJ, Kogan V, Shelton JF, Delwiche L, Hansen RL, Ozonoff S, Ma CC, McCanlies EC, model, for each day, circular buffers were drawn around each
Bennett DH, Hertz-Picciotto I, Tancredi DJ, Volk HE. home with racceptable daily intakei of 1,250, 1,500, and 1750
m. If the buffer intersected the centroid (centermost point) of
BACKGROUND: Maternal folic acid (FA) protects against developmental toxicity from certain an MTRS where pesticides had been applied, the type and
environmental chemicals. OBJECTIVE: We examined combined exposures to maternal FA and amount were linked to the home as a proximal exposure. This
pesticides in relation to autism spectrum disorder (ASD). METHODS: Participants were California model generated an exposure profile by day of pregnancy. All
children born from 2000–07 who were enrolled in the Childhood Autism Risks from Genetics and the records with no exposure identified were assigned to zero
Environment (CHARGE) case-control study at age 2–5 y, were clinically confirmed to have ASD pounds applied. The daily exposure profile was then
(n=296) or typical development (n=220), and had information on maternal supplemental FA and aggregated into time periods of interest for analysis, such as
APPENDIX 2 121

Summary
Study and published abstract Study evaluation
No.

pesticide exposures. Maternal supplemental FA and household pesticide product use were months and trimesters; for this study, we used the six-mo
retrospectively collected in telephone interviews from 2003–11. High vs. low daily FA intake was period beginning three mo before conception through the end
dichotomized at 800μg (median). Mothers' addresses were linked to a statewide database of of the third month of pregnancy (end of the first trimester).
commercial applications to estimate agricultural pesticide exposure. RESULTS: High FA intake Exposure over the entire pregnancy period was also
(≥800μg) during the first pregnancy month and no known pesticide exposure was the reference assessed.
group for all analyses. Compared with this group, ASD was increased in association with <800μg  FA
and any indoor pesticide exposure {adjusted odds ratio [OR]=2.5 [95% confidence interval (CI): 1.3, Occupational pesticide exposure: exposure assessment was
4.7]} compared with low FA [OR=1.2 (95% CI: 0.7, 2.2)] or indoor pesticides [OR=1.7 (95% CI: 1.1, qualitative only (code of 0 (none), 1 (exposure above
2.8)] alone. ORs for the combination of low FA and regular pregnancy exposure (≥6  mo) to pet background levels; no more than a few days per year), 2
pesticides or to outdoor sprays and foggers were 3.9 (95% CI: 1.4, 11.5) and 4.1 (95% CI: 1.7, (most likely exposed; exposure was unlikely to be daily), or 3
10.1), respectively. ORs for low maternal FA and agricultural pesticide exposure 3 mo before or after (definitely exposed; frequent or routine exposure) was
conception were 2.2 (95% CI: 0.7, 6.5) for chlorpyrifos, 2.3 (95% CI: 0.98, 5.3) for entered to estimate both the frequency and intensity for each
organophosphates, 2.1 (95% CI: 0.9, 4.8) for pyrethroids, and 1.5 (95% CI: 0.5, 4.8) for carbamates. of the agents of interest).
Except for carbamates, these ORs were approximately two times greater than those for either While the study results are noted, the exposure assessment
exposure alone or for the expected ORs for combined exposures under multiplicative or additive was insufficiently quantitative to be of use for setting human
models. CONCLUSIONS: In this study population, associations between pesticide exposures and health based guidance values for chlorpyrifos.
ASD were attenuated among those with high versus low FA intake during the first month of
pregnancy. Confirmatory and mechanistic studies are needed. https://doi.org/10.1289/EHP604.
122 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.
APPENDIX 2 123

Summary
Study and published abstract Study evaluation
No.

72 & 73 Environ Int. 2017 Sep; 106:248–56. doi: 10.1016/j.envint.2017.05.015. Epub 2017 Jun 8. The concentration of chlorpyrifos in placental cord blood
plasma samples that was detected in this study was very high
Prenatal naled and chlorpyrifos exposure is associated with deficits in infant motor function in a (shown as ng/mL = µg/L):
cohort of Chinese infants.
Silver MK, Shao J, Zhu B, Chen M, Xia Y, Kaciroti N, Lozoff B, Meeker JD.
Comment in: Environ Int. 2018 May; 114:50–51.
BACKGROUND: Organophosphate insecticides (OPs) are used worldwide, yet despite nearly
ubiquitous exposure in the general population, few have been studied outside the laboratory. Fetal
brains undergo rapid growth and development, leaving them susceptible to long-term effects of Notably, the 90 th percentile chord blood plasma concentration
neurotoxic OPs. The objective here was to investigate the extent to which prenatal exposure to OPs for chlorpyrifos (1.92 ng/mL) is several orders of magnitude
affects infant motor development. METHODS: 30 OPs were measured in umbilical cord blood using higher than the maximums reported in the U.S. (0.002–0.065
gas chromatography tandem mass spectrometry in a cohort of Chinese infants. Motor function was ng/mL). For comparison, the Columbia Childrens
assessed at 6-weeks and 9-months using Peabody Developmental Motor Scales 2 nd edition (PDMS- Environmental Health Study identified chlorpyrifos at a
2) (n=199). Outcomes included subtest scores: reflexes, stationary, locomotion, grasping, visual- concentration of 9 pg/g in pooled blood samples.
motor integration (V-M), composite scores: gross (GM), fine (FM), total motor (TM), and Furthermore, In a human volunteer study involving a single
standardized motor quotients: gross (GMQ), fine (FMQ), total motor (TMQ). RESULTS: Naled, oral dose of 0.5 mg/kg bw, peak blood chlorpyrifos levels
methamidophos, trichlorfon, chlorpyrifos, and phorate were detected in ≥10% of samples. Prenatal were less than 30 ng/ml, and represented only a fraction of
naled and chlorpyrifos were associated with decreased 9-month motor function. Scores were 0.55, the levels of TCP y .4 Detectable inhibition of plasma
0.85, and 0.90 points lower per 1ng/mL increase in log-naled, for V-M (p=0.04), FM (p=0.04), and cholinesterase occurred at this does level.
FMQ (p=0.08), respectively. For chlorpyrifos, scores were 0.50, 1.98, 0.80, 1.91, 3.49, 2.71, 6.29,
2.56, 2.04, and 2.59 points lower for exposed versus unexposed infants, for reflexes (p=0.04), Assuming that the study’s analytical work is correct (see
locomotion (p=0.02), grasping (p=0.05), V-M (p<0.001), GM (p=0.007), FM (p=0.002), TM (p<0.001), comments below), the data implies very high levels of
GMQ (p=0.01), FMQ (p=0.07), and TMQ (p=0.008), respectively. Girls appeared to be more exposure to chlorpyrifos in the studied individuals ie
sensitive to the negative effects of OPs on 9-month motor function than boys. CONCLUSIONS: We exposure levels potentially well above those that inhibit blood
found deficits in 9-month motor function in infants with prenatal exposure to naled and chlorpyrifos. cholinesterases.
Naled is being aerially sprayed to combat mosquitoes carrying Zika virus, yet this is the first non- However there are a number of valid methodological
occupational human study of its health effects. Delays in early-motor skill acquisition may be concerns with the analytical work carried out in this study: (a)
detrimental for downstream development and cognition. the validation of the assay was performed using serum not
Int J Hyg Environ Health. 2018 Apr; 221(3):469–78. doi: 10.1016/j.ijheh.2018.01.010. plasma samples; (b) the assay quality control samples were
generated using serum samples not plasma samples; (c)
Prenatal organophosphate insecticide exposure and infant sensory function. there was prolonged storage of the samples without the

4
Nolan, R.J., Rick, D.L., Freshour, N.L., and Saunders, J.H. (1984). Chlorpyrifos: Pharmacokinetics in human volunteers. Toxicol. Appl. Pharmacol. 73:8–15.
124 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

Silver MK, Shao J, Ji C, Zhu B, Xu L, Li M, Chen M, Xia Y, Kaciroti N, Lozoff B, Meeker JD. appropriate stability and validation data that is required for
such a procedure; (d) there is a lack of appropriate recovery
BACKGROUND: Occupational studies suggest that exposure to organophosphate insecticides (OPs) data with the analysis (for comparison, the Columbia
can lead to vision or hearing loss. Yet the effects of early-life exposure on visual and auditory Children’s Environmental Health study provided excellent
function are unknown. Here we examined associations between prenatal OP exposure and grating quality control data that ensured confidence in the analytical
visual acuity (VA) and auditory brainstem response (ABR) during infancy. METHODS: 30 OPs were data even though extraction efficiency of chlorpyrifos from the
measured in umbilical cord blood using gas chromatography tandem mass spectrometry in a cohort blood was relatively low,∼18%, the sensitivity of the analysis
of Chinese infants. Grating visual acuity (VA) (n = 179–200) and auditory brainstem response (ABR) plus the incorporation of deuterated internal standards
(n = 139–183) were assessed at 6 weeks, 9 months, and 18 months. Outcomes included VA score, allowed for reliable estimates).
ABR wave V latency and central conduction time, and head circumference (HC). Associations
between sensory outcomes during infancy and cord OPs were examined using linear mixed models. There are also other methodological concerns, most notable
RESULTS: Prenatal chlorpyrifos exposure was associated with lower 9-month grating VA scores; of which are: (a) that the primary purpose of the study was as
scores were 0.64 (95% CI: -1.22, -0.06) points lower for exposed versus unexposed infants a longitudinal study of iron deficiency and infant
(p = 0.03). The OPs examined were not associated with infant ABR latencies, but chlorpyrifos and neurodevelopment; (b) only 237 out of 359 participants had
phorate were both significantly inversely associated with HC at 9 months; HCs were 0.41 (95% CI: sufficient cord blood for pesticide analysis; and (c) of the 237
0.75, 0.6) cm and 0.44 (95% CI: 0.88, 0.1) cm smaller for chlorpyrifos (p = 0.02) and phorate participants with sufficient sample size 19.3% had low cord
(p = 0.04), respectively. CONCLUSIONS: We found deficits in grating VA and HC in 9-month-old ferritin.
infants with prenatal exposure to chlorpyrifos. The clinical significance of these small but statistically
significant deficits is unclear. However, the disruption of visual or auditory pathway maturation in Importantly, given that the initial elimination rate of
infancy could potentially negatively affect downstream cognitive development. chlorpyrifos from plasma is very rapid (plasma half-life of
chlorpyrifos is very short on the order of 1 h or less), it is
likely that the levels of chlorpyrifos detected in maternal and
cord blood at the time of delivery are reflective of a ‘steady-
state’ value that is greatly influenced by an equilibrium
between acceptable daily intakepose tissue and blood lipids.
However, little is known about the terminal half-life of
chlorpyrifos after steady state is achieved following repeated,
relatively low-level exposure. If the terminal half-life was
measured in months, then the values collected at the time of
delivery could be reasonable indicators of the levels of
exposure throughout pregnancy. Conversely, if the terminal
half-life was measured in days, then the values obtained at
the time of delivery might have little relationship to exposure
levels that were present during most the pregnancy. Thus, it
is unclear whether this single measurement of chlorpyrifos in
plasma, collected at the time of delivery, provides a relevant
APPENDIX 2 125

Summary
Study and published abstract Study evaluation
No.

measure of chlorpyrifos exposure during the period of


prenatal central nervous system (CNS) development (last 3–4
months of pregnancy).
Because of its multiple limitations, the results of these
studies (particularly the analytical results) are not regarded
as being reliable. Furthermore, if the blood analytical results
are correct, the level of exposure likely resulted in inhibition
of blood cholinesterases.
74 Int J Hyg Environ Health. 2015 Mar; 2 18(2):203–11. doi: 10.1016/j.ijheh.2014.10.005. Epub 2014 The results of this study indicate that worker exposure
Nov 7. resulted in inhibition of blood cholinesterases ie the exposure
likely exceeded the Australian health based guidance values.
Longitudinal assessment of occupational exposures to the organophosphorous insecticides
chlorpyrifos and profenofos in Egyptian cotton field workers.
Singleton ST, Lein PJ, Dadson OA, McGarrigle BP, Farahat FM, Farahat T, Bonner MR, Fenske RA,
Galvin K, Lasarev MR, Anger WK, Rohlman DS, Olson JR.
Chlorpyrifos (CPF) and profenofos (PFF) are organophosphorus (OP) insecticides that are applied
seasonally in Egypt to cotton fields. Urinary trichloro-2-pyridinol (TCP y ), a specific CPF metabolite,
and 4-bromo-2-chlorophenol (BCP), a specific PFF metabolite, are biomarkers of exposure, while
inhibition of blood butyrylcholinesterase (BChE) and acetylcholinesterase (AChE) activities are
effect biomarkers that may be associated with neurotoxicity. Urinary TCP y and BCP and blood BChE
and AChE activities were measured in 37 adult Egyptian Ministry of Agriculture workers during and
after 9–17 consecutive days of CPF application followed by an application of PFF (9–11 days), and
a second CPF application (5 days) in 2008. During the OP applications, urinary TCP y and BCP
levels were significantly higher than baseline levels, remained elevated following the application
periods, and were associated with an exposure related inhibition of blood BChE and Ache. Analysis
of blood AChE levels before and after the PFF application period suggests that individual workers
with peak BCP levels greater than 1000 μg/g creatinine exhibited further inhibition of blood AChE
with PFF application, demonstrating that PFF exposure had a negative impact on AChE activity in
this highly exposed worker population. While large interindividual differences in exposure were
observed throughout this longitudinal study (peak urinary BCP and peak TCP y levels for individuals
ranging from 13.4 to 8052 and 16.4 to 30,107 μg/g creatinine, respectively), these urinary
biomarkers were highly correlated within workers (r=0.75, p<0.001). This suggests that the relative
exposures to CPF and PFF were highly correlated for a given worker. The variable exposures
126 RECONSIDERATION OF CHLORPYRIFOS: 2019 TOXICOLOGY UPDATE

Summary
Study and published abstract Study evaluation
No.

between job classification and work site suggest that job title and work location should not be used
as the sole basis for categorizing OP exposures when assessing neurobehavioural and other health
outcomes in Egyptian cotton field workers. Together, these findings will be important in educating
the Egyptian insecticide application workers in order to encourage the development and
implementation of work practices and personal protective equipment to reduce their exposure to
CPF and PFF.
75 Environ Res. 2016 Aug; 149:164–70. doi: 10.1016/j.envres.2016.05.011. Epub 2016 May 18. The results of this study indicate that worker exposure very
likely exceeded the Australian health based guidance values.
Chlorpyrifos exposure in farmers and urban adults: Metabolic characteristic, exposure estimation, Importantly, the difficulties and inaccuracies in using urinary
and potential effect of oxidative damage. TCP y values to back calculate likely chlorpyrifos exposure
Wang L, Liu Z, Zhang J, Wu Y, Sun H. levels are discussed above. Such techniques are likely semi-
quantitative at best, thus this study may tend to overestimate
Chlorpyrifos is a widely used organophosphorus pesticide that efficiently protects crops against the accuracy of this approach.
pests. However, recent studies suggest that severe exposure to chlorpyrifos may present adverse
health effects in human. To analyze the exposure level and metabolic characteristics of chlorpyrifos
pesticide in urban adults and farmers with/without occupation pesticide contact, the occurrence of
urinary chlorpyrifos and methyl chlorpyrifos (CP-me), as well as their metabolite, 3,5,6-trichloro-2-
pyridinol (TCP y), was determined in farmers of an agricultural village in China, and in urban adults of
a nearby town. The geometric mean (GM) concentrations of TCP y, which is the major marker of
chlorpyrifos exposure, were 4.29 and 7.57μg/g-creatinine in urban adults and farmers before
pesticide application, respectively. Chlorpyrifos spraying significantly increased the concentrations
of urinary TCP y. In the first day after spraying, a GM concentration of 43.7μg/g-creatinine was
detected in the urine specimens from farmers, which decreased to 38.1 and 22.8μg/g-creatinine in
the second and third day after chlorpyrifos spraying. The ratio of TCP y and its parent compounds, ie
chlorpyrifos and CP-me, was positively associated with the sum concentration of urinary
chlorpyrifos, CP-me, and TCP y, suggesting the increasing metabolic efficiency of chlorpyrifos to
TCP y at higher chlorpyrifos exposure levels. To estimate the farmers' occupational exposure to
chlorpyrifos pesticide, a new model based on the fitted first-order elimination kinetics of TCP y was
established. Occupational chlorpyrifos exposure in a farmer was estimated to be 3.70μg/kg-bw/day
(GM), which is an exposure level that is higher than the recommended guideline levels. Significant
increase of urinary 8-hydroxydeoxyguanosine (8-OHdG) was observed on the first day after
chlorpyrifos spraying, which indicates a potential oxidative damage in farmers. However, urinary 8-
OHdG returned to its baseline level within two days.
76 Regul Toxicol Pharmacol. 2015 Mar; 71(2):235–43. doi: 10.1016/j.yrtph.2014.12.013. Epub 2014 While the study provides useful background data regarding
APPENDIX 2 127

Summary
Study and published abstract Study evaluation
No.

Dec 25. the interpretation of biomonitoring data based on inhibition of


erythrocyte cholinesterase, the surrogate key effect in
Derivation of human Biomonitoring Guidance Values for chlorpyrifos using a physiologically based humans in Australia is regarded as being plasma
pharmacokinetic and pharmacodynamic model of cholinesterase inhibition. cholinesterase inhibition.
Arnold SM, Morriss A, Velovitch J, Juberg D, Burns CJ, Bartels M, Aggarwal M, Poet T, Hays S,
Price P.
A number of biomonitoring surveys have been performed for chlorpyrifos (CPF) and its metabolite
(3,5,6-trichloro-2-pyridinol, TCP y ); however, there is no available guidance on how to interpret these
data in a health risk assessment context. To address this gap, Biomonitoring Guidance Values
(BGVs) are developed using a physiologically based pharmacokinetic and pharmacodynamic
(PBPK/PD) model. The PBPK/PD model is used to predict the impact of age and human variability
on the relationship between an early marker of cholinesterase (ChE) inhibition in the peripheral and
central nervous systems [10% red blood cell (RBC) ChE inhibition] and levels of systemic
biomarkers. Since the PBPK/PD model characterizes variation of sensitivity to CPF in humans,
interspecies and intraspecies uncertainty factors are not needed. Derived BGVs represent the
concentration of blood CPF and urinary TCP y associated with 95% of the population having less
than or equal to 10% RBC ChE inhibition. Blood BGV values for CPF in adults and infants are 6100
ng/L and 4200 ng/L, respectively. Urinary TCP y BGVs for adults and infants are 2100 μg/L and 520
μg/L, respectively. The reported biomonitoring data are more than 150-fold lower than the BGVs
suggesting that current US population exposures to CPF are well below levels associated with any
adverse health effect.

You might also like