23
Potential Developmental Neurotoxicity of Pesticides used in Europe (Article begins on next page) The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters. Citation Bjørling-Poulsen, Marina, Helle Raun Andersen, and Philippe Grandjean. 2008. Potential developmental neurotoxicity of pesticides used in Europe. Environmental Health 7:50. Published Version doi:10.1186/1476-069X-7-50 Accessed February 19, 2015 12:24:50 AM EST Citable Link http://nrs.harvard.edu/urn-3:HUL.InstRepos:4584624 Terms of Use This article was downloaded from Harvard University's DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http://nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of- use#LAA

Potential Developmental Neurotoxicity of Pesticides used in ...Laboratory experimental studies using model compounds suggest that many pesticides currently used in Europe – including

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Potential Developmental Neurotoxicity of Pesticides used in Europe

    (Article begins on next page)

    The Harvard community has made this article openly available.Please share how this access benefits you. Your story matters.

    Citation Bjørling-Poulsen, Marina, Helle Raun Andersen, and PhilippeGrandjean. 2008. Potential developmental neurotoxicity ofpesticides used in Europe. Environmental Health 7:50.

    Published Version doi:10.1186/1476-069X-7-50

    Accessed February 19, 2015 12:24:50 AM EST

    Citable Link http://nrs.harvard.edu/urn-3:HUL.InstRepos:4584624

    Terms of Use This article was downloaded from Harvard University's DASHrepository, and is made available under the terms and conditionsapplicable to Other Posted Material, as set forth athttp://nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of-use#LAA

    http://osc.hul.harvard.edu/dash/open-access-feedback?handle=1/4584624&title=Potential+Developmental+Neurotoxicity+of+Pesticides+used+in+Europehttp://dx.doi.org/10.1186/1476-069X-7-50http://nrs.harvard.edu/urn-3:HUL.InstRepos:4584624http://nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of-use#LAAhttp://nrs.harvard.edu/urn-3:HUL.InstRepos:dash.current.terms-of-use#LAA

  • BioMed CentralEnvironmental Health

    ss

    Open AcceReviewPotential developmental neurotoxicity of pesticides used in EuropeMarina Bjørling-Poulsen*1, Helle Raun Andersen1 and Philippe Grandjean1,2

    Address: 1Department of Environmental Medicine, University of Southern Denmark, Winslowparken 17, 5000 Odense, Denmark and 2Department of Environmental Health, Harvard School of Public Health, Landmark Building 3E-110, 401 Park Drive, Boston, MA 02215, USA

    Email: Marina Bjørling-Poulsen* - [email protected]; Helle Raun Andersen - [email protected]; Philippe Grandjean - [email protected]

    * Corresponding author

    AbstractPesticides used in agriculture are designed to protect crops against unwanted species, such asweeds, insects, and fungus. Many compounds target the nervous system of insect pests. Because ofthe similarity in brain biochemistry, such pesticides may also be neurotoxic to humans. Concernshave been raised that the developing brain may be particularly vulnerable to adverse effects ofneurotoxic pesticides. Current requirements for safety testing do not include developmentalneurotoxicity. We therefore undertook a systematic evaluation of published evidence onneurotoxicity of pesticides in current use, with specific emphasis on risks during early development.Epidemiologic studies show associations with neurodevelopmental deficits, but mainly deal withmixed exposures to pesticides. Laboratory experimental studies using model compounds suggestthat many pesticides currently used in Europe – including organophosphates, carbamates,pyrethroids, ethylenebisdithiocarbamates, and chlorophenoxy herbicides – can causeneurodevelopmental toxicity. Adverse effects on brain development can be severe and irreversible.Prevention should therefore be a public health priority. The occurrence of residues in food andother types of human exposures should be prevented with regard to the pesticide groups that areknown to be neurotoxic. For other substances, given their widespread use and the uniquevulnerability of the developing brain, the general lack of data on developmental neurotoxicity callsfor investment in targeted research. While awaiting more definite evidence, existing uncertaintiesshould be considered in light of the need for precautionary action to protect brain development.

    IntroductionPesticides are used widely in agriculture to maintain andincrease crop yields, and they are also applied in homesand gardens. The annual application of synthetic pesti-cides to food crops in the EU exceeds 140,000 tonnes [1],an amount that corresponds to 280 grams per EU citizenper year. Despite European policies to reduce pesticideuse, EU statistics data for 1992–2003 show that theannual pesticide consumption has not decreased [1]. Afew hundred different compounds are authorised for use

    in all EU member states, but a similar number of pesti-cides is in current use in different EU countries and arebeing evaluated for possible authorisation in all of EU.Approximately 300 different pesticides have beenreported as contaminants of food products of Europeanorigin [2]. Up to 50 percent of fruits, vegetables and cere-als grown in the European Union are known to containpesticide residues [2], but only a small fraction of pesti-cides in current use are included in the monitoring pro-grammes. Nonetheless, one out of twenty food items is

    Published: 22 October 2008

    Environmental Health 2008, 7:50 doi:10.1186/1476-069X-7-50

    Received: 26 August 2008Accepted: 22 October 2008

    This article is available from: http://www.ehjournal.net/content/7/1/50

    © 2008 Bjørling-Poulsen et al; licensee BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

    Page 1 of 22(page number not for citation purposes)

    http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?cmd=Retrieve&db=PubMed&dopt=Abstract&list_uids=18945337http://www.ehjournal.net/content/7/1/50http://creativecommons.org/licenses/by/2.0http://www.biomedcentral.com/http://www.biomedcentral.com/info/about/charter/

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    known to exceed a current EU legal limit for an individualpesticide [2]. Further, over 25% of fruits, vegetables, andcereals contain detectable residues of at least two pesti-cides [2]. Processed food and baby food are also com-monly contaminated. In addition, other sources, such ascontaminated drinking water, dusts and spray drift con-tribute to human exposures.

    The total level of population exposures to pesticides inEurope is unknown, but data from US population studiesshow that the majority of the population has detectableconcentrations of methyl phosphate, ethyl phosphate,and other pesticide metabolites in the urine [3].

    Many pesticides target the nervous system of insect pests.Because of the similarity of neurochemical processes,these compounds are also likely to be neurotoxic tohumans. This concern is of particular relevance to thedeveloping human brain, which is inherently much morevulnerable to injury caused by toxic agents than the brainof adults [4]. During prenatal life, the human brain mustdevelop from the ectodermal cells of the embryo into acomplex organ consisting of billions of precisely located,highly interconnected, and specialised cells. For optimumbrain development neurons must move along precisepathways from their points of origin to their assignedlocations, they must establish connections with othercells, and they must learn to communicate with other cellsvia these connections [4-6]. All of these processes have totake place within a tightly controlled time frame, and eachdevelopmental stage has to be reached on schedule and inthe correct sequence. If a developmental process in thebrain is halted or inhibited, there is little potential for laterrepair, and the consequences may therefore be permanent[4,6].

    Concerns in regard to developmental neurotoxicity due topesticides have been fuelled by recent epidemiologicobservations that children exposed prenatally or duringearly postnatal life suffer from various neurological defi-cits [7-12]. Urinary pesticide metabolite concentrationsassociated with adverse effects overlap with the rangesthat occur in the general population [3]. Although theidentity of the parent pesticides and the exact magnitudeof causative exposures are unclear, these observations sug-gest that developmental neurotoxicity from pesticideexposure is a public health concern.

    Despite the increasing recognition of the need to evaluatedevelopmental neurotoxicity in safety assessment [13-15], only very few of the commercial chemicals in currentuse have been examined with respect to neurodevelop-mental effects [16]. Validated rodent models exist, butthey are considered expensive and are only infrequentlyused. According to the current EU Plant Protection Direc-

    tive (91-414-EEC), a neurotoxicity test in hens is requiredonly for organophosphates and some carbamates to assessthe possible risk of delayed peripheral neurotoxicity fol-lowing acute exposure.

    From a public health viewpoint, the prevention of neu-rodevelopmental disorders is a priority; these disordersinclude learning disabilities, attention deficit hyperactiv-ity disorder (ADHD), autism spectrum disorders, devel-opmental delays, and emotional and behaviouralproblems. The causes of these disorders are unclear, andinteracting genetic, environmental, and social factors arelikely determinants of abnormal brain development [17].Medical statistics data are difficult to compare betweencountries, but one report suggests that 17% of US childrenunder18 years of age suffer from a developmental disabil-ity, in most cases affecting the nervous system [18]. In cal-culations of environmental burdens of disease inchildren, lead neurotoxicity to the developing brain is amajor contributor [19]. Pesticide effects could well be ofthe same magnitude, or larger, depending on the exposurelevels.

    A recent review [16] listed 201 chemicals known to beneurotoxic in humans; only 5 of these substances havebeen firmly documented as causes of developmental neu-rotoxicity. Identification of human neurotoxicity wasbased on available evidence, including poisoning inci-dents described in the scientific literature, as identifiedfrom the Hazardous Substances Data Bank of the U.S.National Library of Medicine. Although published clinicalinformation may not be representative for the relativeneurotoxicity risks due to industrial chemicals, it is note-worthy that a total of 90 (45%) of the neurotoxic sub-stances are pesticides. For these substances, onlyneurotoxicity in adults had been documented, therebydocumenting that access to the brain is possible and maycause toxic effects. Given the vulnerability of the develop-ing brain, it is likely that many of these substances willalso be capable of causing developmental neurotoxicity[16]. Indeed, studies in laboratory animals support thenotion that a wide range of industrial chemicals can causedevelopmental neurotoxicity even at low doses that arenot harmful to mature animals [14,20].

    Given the likely importance of pesticides in regard todevelopmental neurotoxicity in humans, this reviewfocuses on pesticides approved for current use in Europe,i.e. either authorised or being evaluated for authorisationwithin the European Union (Table 1). Our literaturesearch was conducted by similar means as the previousreview mentioned above [16], but included relevant datafrom laboratory experiments. The pesticides are groupedin accordance with the likely mechanism of action orchemical similarity. We focus on substances with a pri-

    Page 2 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    mary application as pesticides and therefore exclude sub-stances like nicotine, warfarin, and ethanol with otherprimary uses.

    Search strategy and selectionWe first identified pesticides that have caused neurotoxiceffects in humans from the Hazardous Substances DataBank (HSDB) of the U.S. National Library of Medicine[16]. We searched for the terms "pesticide" and "neuro*".From the list of substances obtained in this way, we iden-tified the pesticides, for which neurotoxic effects inhumans had been reported. In addition, we searched theU.S. National Institute of Occupational Safety and Health(NIOSH) – Pocket Guide to Chemical Hazards http://www.cdc.gov/Niosh/npg/npgsyn-p.html, using the search

    terms "pesticide", "insecticide", "herbicide", "fungicide","fumigant", and "rodenticide" in combination with "cen-tral nervous system". The list of neurotoxic pesticidesidentified in this way was then compared to the currentAnnex 1 list (as of August, 2008) of pesticides authorisedin the European Union according to Plant ProtectionDirective 91-414-EEC (an Excel data sheet with the statusof active substances under EU review can be downloadedfrom http://ec.europa.eu/food/plant/protection/pesticides/index_en.print.htm). We chose pesticides with anAnnex 1 status "in" or "pending" for consideration (Table1).

    For each neurotoxic pesticide in current use, we searchedPubMed to identify published data on developmental

    Table 1: Neurotoxic pesticides, which are authorised or pending evaluation for authorisation in the EU

    Pesticide Annex 1 status

    Organophosphate insecticidesChlorpyrifos InDimethoate InEthoprophos InPhosmet InFenamiphos (nematicide) In

    CarbamatesPirimicarb InMethomyl Application resubmitted

    Pyrethroid insecticidesCypermethrin (type II) InDeltamethrin (type II) InPyrethrum/pyrethrin (natural pyrethrin) Pending

    Other insecticidesNicotine Pending

    Dithiocarbamate fungicidesManeb InThiram In

    Chlorophenoxy herbicides2,4-D In

    Bipyridyl herbicidesDiquat dibromide In

    RodenticidesWarfarin In

    FumigantsPhosphides (zinc, magnesium, and aluminum phosphides) PendingSulfuryl fluoride Pending

    The list includes pesticicides, which are registered as "in" or "pending" on the current EU Annex 1 list (as of August 2008), and for which neurotoxicity in humans has been reported in The Hazardous Substances Data Bank and/or the NIOSH Pocket Guide to Chemical Hazards. The full Annex 1 list with the status of active substances under EU review can be downloaded as an Excel sheet at http://ec.europa.eu/food/plant/protection/pesticides/index_en.print.htm.

    Page 3 of 22(page number not for citation purposes)

    http://www.cdc.gov/Niosh/npg/npgsyn-p.htmlhttp://www.cdc.gov/Niosh/npg/npgsyn-p.htmlhttp://ec.europa.eu/food/plant/protection/pesticides/index_en.print.htmhttp://ec.europa.eu/food/plant/protection/pesticides/index_en.print.htmhttp://ec.europa.eu/food/plant/protection/pesticides/index_en.print.htmhttp://ec.europa.eu/food/plant/protection/pesticides/index_en.print.htm

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    neurotoxicity. We used pesticide synonyms, commercialnames and the CAS number, in combination with each ofthe terms "neurotoxic", "neurotoxicity", "neurologic","neurological" and "nervous system", and additionalsearches included the terms "prenatal", "pregnancy","fetus", "fetal", "maternal", "developmental" and "child".

    Organophosphate insecticidesToxic mechanismsThe primary target of organophosphate (OP) insecticidesis the enzyme acetylcholinesterase (AChE), which hydro-lyses the neurotransmitter acetylcholine in both theperipheral and the central nervous system. OPs contain-ing a P = O moiety are effective inhibitors of AChE,whereas OPs with a P = S moiety require bioactivation toform an "oxon" or oxygen analogue of the parent com-pound. Inhibition of AChE by OPs is obtained by the P =O moiety forming a covalent bond with the active site ofthe enzyme. The enzyme-inhibitor complex can become"aged" by a non-enzymatic hydrolysis of one of the tworadical groups in the OP, and once the complex has aged,inhibition of AChE is irreversible (reviewed in [21]). Inhi-bition of AChE causes accumulation of acetylcholine atcholinergic synapses, leading to over-stimulation of mus-carinic and nicotinic receptors. In addition, acetylcholinehas important functions during brain development [22].

    In severe cases of OP poisoning in adults (AChE inhibi-tion exceeding 70%) [23], a "cholinergic syndrome" iselicited, including various central nervous system (CNS)effects such as headache, drowsiness, dizziness, confu-sion, blurred vision, slurred speech, ataxia, coma, convul-sions and block of respiratory centre [24]. Some OPs canalso induce a delayed neuropathy which does not involveinhibition of AChE but rather the neuropathy target este-rase (NTE) [25,26]. The physiological functions of NTEare still unknown, and it is obscure how phosphorylationand aging of NTE leads to axonal degeneration [27].

    The syndromes described above are observed only follow-ing high dose, acute exposures to OPs. Survivors recoverfrom these syndromes, but it is likely that the exposurealso produces long-term adverse health effects. In rats, asingle high exposure to an OP can cause long lastingbehavioural effects [28,29], and the same has beenreported from several human studies (e.g. [30,31]).

    The concern is growing that also chronic, low exposures toOPs may produce neurological effects, although the evi-dence remains somewhat equivocal (reviewed in [32-34]). Most studies have found an association of OP expo-sure with increased neurological symptom prevalence. Asan example, Hispanic immigrant farm workers in the UShave a poorer neurobehavioural performance than non-agricultural Hispanic immigrants. Within the group of

    agricultural workers there was a positive correlationbetween urinary OP metabolite levels and poorer per-formance on some neurobehavioural tests [35]. A cross-sectional study of pesticide applicators reported that neu-rological symptoms were associated with cumulativeexposure to moderate levels of organophosphate andorganochlorine insecticides, regardless of recent exposurehistory [36].

    Acetylcholine and other neurotransmitters play uniquetrophic roles in the development of the CNS [37,38], andinhibition of AChE by OPs and the resulting accumula-tion of acetylcholine may then conceivably disturb thisdevelopment. Still, developing rats recover faster fromAChE inhibition than adults, largely due to the fact thatdeveloping organisms have a rapid synthesis of new AChEmolecules [39-41]. It therefore seems that either develop-mental toxicity may be unrelated to AChE inhibition, orthat even a brief period of AChE inhibition is sufficient todisrupt development [42].

    Chlorpyrifos is the most extensively studied OP withrespect to developmental neurotoxicity in laboratorymodels. Prenatal or neonatal exposure has been shown tocause a variety of behavioural abnormalities in both miceand rats, including changes in locomotor skills and cogni-tive performance [43-46]. At concentrations comparableto those found in human meconium [47], experiments onrat embryo cultures showed mitotic abnormalities andevidence of apoptosis during the neural tube develop-ment stage, and significant effects even at concentrationsmore than an order of magnitude below those present inhuman meconium [48]. However, exposure of rat foe-tuses to chlorpyrifos by maternal administration did notinduce large immediate effects on brain development[49], but chlorpyrifos treatment during gestation, didcause deficits in brain cell numbers, neuritic projections,and synaptic communication, which emerged in adoles-cence and continued into adulthood. This finding indi-cates that chlorpyrifos exposure during gestation results inaltered programming of synaptic development [50,51].

    The window of vulnerability to chlorpyrifos extends intorelatively late stages of brain development, and chlorpyri-fos can induce neurobehavioural abnormalities duringthe second and third postnatal weeks in rat [43,52,53],corresponding to the neonatal stage in humans [54]. Thisperiod is outside the major phase of neurogenesis in mostbrain regions, but it is the period of peak gliogenesis andsynaptogenesis; developing glia have been found to beeven more sensitive to chlorpyrifos than neurons [55-57].

    Deficits elicited by prenatal exposure to chlorpyrifos areevident even at exposures below the threshold for detect-able AChE inhibition, i.e. far below the 70% inhibition of

    Page 4 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    AChE required for systemic toxicity in adults [43-46,51].These findings suggest that mechanisms other than inhi-bition of AChE activity may, at least in part, be responsi-ble for the developmental neurotoxicity of chlorpyrifos.

    The non-cholinergic mechanisms of chlorpyrifos are notclear, but a possible target may be the signalling cascadesinvolved in neuronal and hormonal inputs, including thecyclic-AMP – protein kinase A cascade, receptor signallingthrough protein kinase C, and direct effects on the expres-sion and function of nuclear transcription factors mediat-ing the switch from proliferation to differentiation,including c-fos, p53, AP-1, Sp1 and CREB (Ca2+/cAMPresponse element binding protein) (reviewed in [42]).

    The notion that chlorpyrifos may exert developmentalneurotoxicity through mechanisms other than inhibitionof AChE opens the possibility that OPs may have com-pound specific effects that may be unrelated to the com-mon AChE inhibitory effect. For example, microarrayanalysis has shown that the two OPs, chlorpyrifos anddiazinon, have many similar effects on gene expression inthe neonatal rat brain, but also notable disparities. All ofthe changes in gene expression induced by the two OPswere observed with doses, which did not induce biologi-cally significant AChE inhibition [58,59]. In neonatal rats,diazinon and chlorpyrifos elicit each their unique patternof damage/repair and altered synaptic function, eventhough OPs as a class target neural cell development andACh systems [60].

    Thus, findings of OP induced developmental neurotoxic-ity through individual mechanism other than the com-mon AChE inhibition complicate extrapolation of effectsfrom one OP to another. The existence of clear effects ofOPs at doses below the threshold for AChE inhibitionclearly demonstrate that it is inadequate to use AChEmeasurements alone as a biomarker for defining safeexposure limits for developmental neurotoxicity of OPs[60].

    Epidemiologic evidenceWith respect to developmental neurotoxicity of OPs inhumans, knowledge is still relatively sparse, and moststudies reflect exposures to more than one pesticide.

    In California, USA, an association was found betweenreflex abnormalities in neonates and increased concentra-tions of OP metabolites measured in the mother's urineduring pregnancy [7]. In a follow-up of the same cohort,urinary dialkyl phosphate metabolite levels during preg-nancy, particularly from dimethyl phosphate pesticides,were negatively associated with mental development inthe children at 24 months of age. No associations were

    observed between neurodevelopment and metabolitesspecific to malathion and chlorpyrifos [8].

    In a cohort study of mothers and infants in New York City,USA, maternal levels of chlorpyrifos above the limit ofdetection, coupled with low maternal levels of paraoxo-nase activity (an enzyme which hydrolyses certain OPs,including chlorpyrifos oxon), were associated withreduced head circumference in the infants [61]. In thesame cohort, prenatal levels of OP metabolites in themother's urine were associated with anomalies of primi-tive reflexes in the infants [9].

    In another New York City cohort, prenatal chlorpyrifosexposures were found to be inversely associated with birthweight and length [62]. In a follow up of this study, thechildren's cognitive and motor development was exam-ined at 1, 2 and 3 years of age. The adjusted mean 3-yearPsychomotor Development Index and Mental Develop-ment Index scores of the highly exposed children differedby 7.1 and 3.0 points, respectively, from the scores of chil-dren with low prenatal exposure to chlorpyrifos. The pro-portion of delayed children in the high-exposure group,compared with the low-exposure group, was five timesgreater for the Psychomotor Development Index and 2.4times greater for the Mental Development Index [10].

    Ecuadorian schoolchildren, whose mothers had beenexposed to OPs and other pesticides during pregnancy byworking in greenhouses, showed visuospatial deficitscompared to children, whose mothers had not beenexposed to pesticides during pregnancy. Furthermore, cur-rent exposure of the children, measured as the excretion ofOP metabolites in urine, was found to be associated withincreased reaction time [11].

    In two US states, Ohio and Mississippi, children wereacutely exposed to the OP, methyl parathion, and whenanalysed for neurobehavioural development, the exposedchildren were found to suffer from persistent problemswith short-term memory and attention [12].

    Although the epidemiological evidence for developmen-tal neurotoxicity of OPs in humans is relatively sparse,there are clear indices of adverse effects. Urinary pesticidemetabolite levels in the above studies were similar tothose that have been recorded from the US general popu-lation [3,63] and in EU countries [64-66].

    Carbamate insecticidesCarbamate insecticides, like the OP insecticides, inhibitAChE and elicit cholinergic hyperstimulation. However,carbamates cause only reversible inhibition of AChE [67].Thus, AChE inhibition by carbamates lasts only minutesor hours, whereas the effects of OPs with respect to AChE

    Page 5 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    can last for 3–4 months (reviewed in [32]). Because of thetransient inhibition of AChE, acute intoxication by car-bamates generally resolves within a few hours [67].

    When comparing the clinical course of carbamate poison-ing (by aldicarb or methomyl) in young children (1–8years old) and adults (17–41 years old), it was found thatthe predominant symptoms in children were CNS depres-sion and hypotonia, and the most common muscariniceffect was diarrhoea. In adults the main symptoms weremiosis and fasciculations, whereas CNS depression, hypo-tonia, and diarrhoea were uncommon [68]. Symptoms inchildren poisoned with OPs were found to be similar tosymptoms for carbamate poisoning [69]. Thus symptomsof carbamate poisoning do not differ markedly fromsymptoms of OP poisoning in children, but rather thesymptoms in children, differ from symptoms in adults.

    It is possible that some carbamates may also be involvedin oxidative stress [70,71]. The carbamate, carbofuran, hasbeen observed to accentuate oxidative stress in rat brainby inducing lipid peroxidation and diminishing the anti-oxidant defence [70].

    As for the OPs, it is likely that poisoning with carbamatesmay result in long term neurological effects [72]. Twopatients showed cognitive deficit in attention, memory,perceptual, and motor domains 12 months after a poison-ing incident [72]. With respect to long term, low levelexposures to carbamates, reports concerning chronic tox-icity are almost non-existent.

    No epidemiological studies of developmental neurotoxic-ity of carbamates in humans could be found, and datafrom animal experiments are very sparse as well.

    Assuming that some of the neurotoxic effects observed inassociation with prenatal exposure to OPs, such as chlo-rpyrifos, are due to inhibition of AChE, it is possible thatcarbamates may have similar developmental effects, eventhough the inhibition of AChE by carbamates is only tran-sient. Induction of oxidative stress by some carbamatesmight also cause developmental neurotoxicity. It shouldalso be noted that the carbamate physostigmine inhibitsDNA synthesis in undifferentiated neuronotypic PC12cells (a standard in vitro model for neuronal develop-ment). When differentiation was induced, adverse effectson DNA synthesis were intensified, and effects on cellnumber after prolonged exposure were also worsened bydifferentiation. Furthermore, differentiating cells dis-played signs of oxidative stress, as measured by lipid per-oxidation. Finally, the transmitter fate of the cells wasshifted away from cholinergic phaenotype toward the cat-echolaminergic phaenotype. Similar findings were made

    when incubating the cells with the OPs chlorpyrifos,diazinon and parathion [73].

    Pyrethroid insecticidesThe pyrethroids are a class of insecticides derived fromnaturally occurring pyrethrins from the Chrysanthemumgenus of plants [74]. Pyrethroids contain several commonfeatures: an acid moiety, a central ester, and an alcoholmoiety. Several stereoisomers exist of each pyrethroidcompound, and their effects are stereospecific, indicatingpresence of specific binding sites (reviewed in [75]).

    The acute toxicity of pyrethroids is mainly mediated byprolongation of the kinetics of voltage-gated sodiumchannels, which are responsible for generation of theinward sodium current that produces the action potentialin excitable cells. Specific interaction of pyrethroids withthe sodium channel slows down both the activation andinactivation properties of the channel, leading to a hyper-excitable state. Although activation is slowed at the singlechannel level, the density of sodium channels in excitablecells is so high that there are always sufficient unmodifiedchannels to ensure that the activation phase of the actionpotential is not delayed. However, in the falling phase ofthe action potential, even a low proportion of modifiedchannels can generate enough extra current to delay inac-tivation. This delay causes prolonged depolarisation,which, if the current is large enough and lasts long enoughfor neighbouring unmodified channels to recover excita-bility, can trigger a second action potential (reviewed in[76]).

    Two types of pyrethroid structures exist. The type II pyre-throids contain a cyano-group in the α-position, whereastype I pyrethroids do not contain a cyano-group [77]. Thetwo types differ with respect to the toxic signs they pro-duce in rats, and with respect to the prolongation time ofthe sodium current they induce. Type I compounds pro-long channel opening just long enough to induce repeti-tive firing of action potentials (time constants less than 10msec), whereas type II compounds (time constants ofmore than 10 msec) hold the channels open for so longthat the membrane potential ultimately becomes depolar-ised to the point at which generation of action potentialsis no longer possible (reviewed in [75]).

    Human pyrethroid poisoning is rare, and almost entirelyinvolves type II pyrethroids. The main adverse effect ofdermal exposure to type II pyrethroids is paresthesias, pre-sumably due to hyperactivity of cutaneous sensory nervefibres. Dizziness, headache and fatigue are commonsymptoms following ingestion of type II pyrethroids. Insevere cases coma and convulsions are the principal life-threatening features [77].

    Page 6 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    The effects of pyrethroids on the CNS are complex andmay also involve antagonism of γ-aminobutyric acid(GABA), modulation of nicotinic cholinergic transmis-sion, enhancement of noradrenalin release, and directactions on calcium or chloride ion channels. Still, becauseneurotransmitter-specific pharmacological agents do notprotect very well against pyrethroid poisoning, it isunlikely that any one of these effects represents a primarytoxic mechanism of action of pyrethroids. More likely,they are secondary to the effects on sodium channels,since most neurotransmitters are released secondary toincreased sodium entry (reviewed in [76]).

    In the few existing accounts of poisonings of adults withpyrethroids, successful recovery after the acute phase ofpoisoning has been described [78,79]. However, nodetailed neuropsychological testing was applied to thesepatients, and also no post mortem examinations have beenreported, and therefore it is unknown if such poisoningsmay have lasting effects. Likewise, no information is avail-able on long term effects of low level chronic exposure inhumans.

    Neonatal rats are 4–17 times more vulnerable to the acutetoxicity of pyrethroids (including permethrin (type I), del-tamethrin (type II), cypermethrin (type II)) than adult rats[80,81]. The higher toxicity in neonates is affected by thelower capacity for metabolic detoxification, sinceneonates and adults have similar brain concentrations atdifferent, but equitoxic, doses [80]. However, anotherstudy did not observe any age-dependency of the toxicityof the two type I pyrethroids, cismethrin and permethrin[82]. It has therefore been argued that age-dependent sen-sitivity to pyrethroids is only apparent at high acute doses,not at doses below those causing overt toxicity [82].

    In addition to the possibility that young animals are morevulnerable to pyrethroids due to lower metabolic detoxi-fication, there is also a possibility that increased vulnera-bility in young animals may be due to more specificeffects of early life exposures. For example, several studieshave found that embryonically expressed forms of volt-age-gated sodium channels are replaced by adult forms asneurodevelopment proceeds (reviewed in[75]), and thisdifference in expression profile may affect the sensitivitytowards pyrethroids. In mutation and knockout modelsof the voltage-gated sodium channels, perturbation ofchannel function during development impairs nervoussystem structure and function, underlining the impor-tance of these channels in neurodevelopment. (reviewedin [75]).

    Also in humans, perturbations of nervous system develop-ment have been associated with altered structure andfunction of voltage-gated sodium channels. Mutations in

    genes encoding sodium channel subunits have been iden-tified, which result in neuronal hyperexcitability due tosubtle changes in channel gating and inactivation [83].Since pyrethroids also alter the activation and inactivationof sodium channels, and thereby the neuronal excitabil-ity, it is possible that these may have effects similar tomutations in the sodium channels. However, the mecha-nisms and magnitude of mutational versus pyrethroideffects are different, and also the duration of effect will dif-fer (pyrethroids have a relatively short half-life, whereasmutations are permanent) [75].

    Another possible indication that pyrethroid effects onsodium channels may be relevant to neurodevelopment isthe observation that developmental exposure to pheny-toin, an anticonvulsant that blocks sodium channels andother ion channels, disrupts nervous system structure andfunction [84]. The use of anticonvulsants during preg-nancy has been associated with adverse effects, includingmicrocephaly and intellectual impairment (reviewed in[75]). Although differences in doses and in pathogenesismay occur, this evidence would support a concern aboutthe effect of pyrethroids on ion channels.

    All existing studies of developmental neurotoxicity ofpyrethroids were conducted with rodents as test animals,and although several of them have reported persistentchanges in behaviour and/or neurochemistry in the ani-mals, results appear somewhat inconsistent (reviewed in[75]). Several studies performed by Eriksson's group [85-87] have shown that mice exposed to pyrethroids on post-natal day 10–16 exhibit increased motor activity and alack of habituation. These mice exhibit changes in densityof muscarinic acetylcholine receptor (mAChR) bindingfor as long as 5 months after cessation of exposure [88].Others have reported persistent changes in behaviourand/or biochemistry, including learning [89], motoractivity [90], sexual behaviour [91], mAChR expression[92,93], and blood-brain barrier permeability [94]. Arecent study in rats showed that neonatal exposure to per-methrin and cypermethrin caused lasting behaviouraleffects, changes in monoamine concentrations in thestriatum as well as increased oxidative stress [95]. In onestudy, both male and female mice were exposed to thetype I pyrethroid, permethrin, before mating, and the fol-lowing functions were affected in the offspring (withparental exposure to 9.8 mg/kg/day or more for 4 weeksbefore mating): development of reflexes, swimming abil-ity and open field activity [96].

    The potential developmental neurotoxicity of pyrethroidshas also been investigated in vitro using cell lines. Forexample non-toxic concentrations (10-6 M) of bifenthrininhibited neurite outgrowth in PC12 cells, indicating thatbifenthrin may have deleterious effects on the developing

    Page 7 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    nervous system at concentrations lower than those capa-ble of causing toxicity in the adult brain [97].

    Existing data indicate that human exposures to pyre-throids occur and result in detectable concentrations inbody fluids [98-100], but there is insufficient informationavailable to adequately evaluate the range of internaldoses in humans, and the consequences of these expo-sures are so far unknown.

    Dithiocarbamate fungicidesDithiocarbamates are non-cholinesterase inhibiting, sul-fur-containing carbamates, which are primarily used asfungicides and herbicides. Four major classes of dithiocar-bamates exist; the methyldithiocarbamates, the dimethyl-dithiocarbamates, the diethyldithiocarbamates (DEDC),and the ethylenebisdithiocarbamates (EBDCs) (reviewedin [101]). The dithiocarbamates used as fungicidesinclude metam sodium (methyldithiocarbamate), thiram(dimethyldithiocarbamate/tetramethyldithiocarbamate),and several EBDCs (mancozeb, maneb, metiram, zineband nabam).

    Dithiocarbamates form lipophilic complexes with di- andtrivalent metallic cations, bonding through the sulfuratoms [102]. They are non-specific in action, and it is dif-ficult to identify a single mechanism for their neurotoxiceffects. Because of their metal-chelating capacity and theiraffinity for sulfhydryl groups, they are biologically highlyactive [103,104]. DEDCs are particularly known to mod-ify the cellular redox state by inducing a copper-depend-ent oxidative stress [105,106], and inhibition of cytosolicCu/Zn superoxidedismutase (SOD1), a key enzyme in theantioxidant response, has been observed in mice treatedwith DEDC [107]. The EBDCs can uncouple the mito-chondrial electron transport chain [108,109]. Mitochon-drial dysfunction is often associated with generation ofreactive oxygen species (ROS), and ROS production wasalso found to play a role in mancozeb induced neuronaltoxicity in mesencephalic cells, likely via redox cyclingwith extracellular and intracellular oxidases [110]. Fur-ther, ethylenethiourea (ETU), which is a degradationproduct of EBDCs, has been shown to inhibit thyroid per-oxidase (TPX), the enzyme that catalyses synthesis of thethyroid hormones [111,112]. In addition, interference ofdithiocarbamates with the vesicular transport of gluta-mate may play a role in their neurotoxicity [113]. Due tothe differences in biochemical effects, these compoundsseem to exhibit a range of different potencies in regard todevelopmental neurotoxicity.

    Dithiocarbamates are reported to display low acute toxic-ity in humans and experimental animals [114]. Both inhumans and laboratory animals, prolonged exposure todithiocarbamates may cause neurotoxicity. Notably,

    peripheral neuropathy and extrapyramidal symptomsresembling parkinsonism have been associated withchronic exposure to dithiocarbamate pesticides [115].

    As mentioned, chronic exposure of humans to EBDCs hasbeen associated with neurocognitive impairment and par-kinsonism [116]. In particular, exposure to maneb, whichcontains manganese, has been linked to development ofparkinsonian-like symptoms in agricultural workers[117,118]. This finding may be related to the inhibition ofcomplex III of the mitochondrial electron transport chain[108], disruption of the glutathione antioxidant system indopaminergic cells [119], inhibition of proteasomal func-tion and induction of α-synuclein aggregates in dopamin-ergic cells [120], induction of catechol autooxidation[121], and potentiation of the neurotoxicity of 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) in mice[122-124]. All of these observations support the notionthat maneb may cause parkinsonian-like symptoms.DEDCs, though not methyldithiocarbamate, can alsoenhance MPTP- induced striatal dopamine depletion inmice [124].

    Both thiram and ziram (dimethyldithiocarbamates) caninduce apoptotic cell death in PC12 cells, in a dose- andtime-dependent manner [125]. Both compounds inducedrapid and sustained increases of intracellular Ca2+ in thecells, which were almost completely blocked byflufenamic acid, an inhibitor of non-selective cation chan-nels. Also, BAPTA-AM, which is an intracellular Ca2+ che-lator, inhibited the thiram and ziram induced apoptoticcell death, indicating that thiram and ziram induce apop-totic neuronal cell death by Ca2+ influx through non-selec-tive cation channels [125].

    The EBDCs maneb, mancozeb and metiram can inducemalformations in rat foetuses, apparently mediatedthrough formation of the ETU metabolite. The malforma-tions predominantly affect the nervous system and thehead, and they correspond to those expected as the resultof thyroid insufficiency. They occur only at doses in excessof those that produce significant thyroid inhibition inadult rats, and they have been prevented, at least in part,by co-administration of thyroxin (reviewed in [126]). Akey concern with thyroid inhibitors is that impaired thy-roid function may alter hormone-mediated events duringdevelopment, thereby possibly leading to permanentalterations in brain morphology and function [127,128].Functional deficits are likely to occur during brain devel-opment even at mild degrees of hypothyroidism [129].Even withing the normal range, a relatively slight reduc-tion of the concentration of maternal thyroid hormonesduring pregnancy can lead to intelligence deficits in thechildren [130]. In addition to EBDCs/ETU, many otherenvironmental contaminants have been found to interfere

    Page 8 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    with thyroid function, for example the chlorophenoxyherbicide, 2,4-D (se below). Some of the mechanisms ofaction with respect to thyroid inhibition are shared bymancozeb/ETU and 2,4-D (including interference withuptake of iodide by the thyroid gland and interferencewith serum protein-bound iodide level) [131], and expo-sure to both EBDCs and chlorophenoxy herbicides maytherefore result in additive effects.

    Evidence that developmental exposure to maneb may beinvolved in development of Parkinson's disease (PD) laterin life includes the finding that postnatal exposure of miceto maneb in combination with paraquat (a classic bipyri-dyl herbicide, which is no longer authorised in EU) led toa permanent and selective loss of dopaminergic neuronsin the substantia nigra pars compacta [132]. The postnatalexposure to these pesticides enhanced the effect of thesame pesticides administered during adulthood, relativeto exposures during development only or adulthood only.Furthermore, exposure to maneb alone during gestationresulted in a dramatic response to paraquat in adulthood,including notable reductions in levels of dopamine and aloss of nigral dopamine neurons. Thus, these results sup-port the notion that a silent neurotoxicity produced bydevelopmental insults can be unmasked by insults later inlife [132].

    For specific dithiocarbamates, especially the EBDCsmaneb and mancozeb, substantial evidence supports thepossibility of developmental neurotoxicity. In addition,the likely mechanisms of toxicity for thiram and ziramindicate that these compounds too may be capable ofcausing developmental neurotoxicity in small doses.

    Chlorophenoxy herbicidesThe chlorophenoxy herbicides are widely used for thecontrol of broad-leaved weeds. Structurally, they consistof a simple aliphatic carboxylic acid moiety, which isattached to a chlorine- (or methyl-) substituted aromaticring by an ether bond. In vivo the salts and esters are rap-idly dissociated or hydrolysed, and therefore the toxicityof each chlorophenoxy compound depends principallyon the acid form of the pesticide [133]. The chlorophe-noxy herbicides bind strongly to albumin [134], andbinding is favoured by longer acid chains and by moregreatly substituted aromatic rings. Therefore the bioavail-ability and toxicity of the herbicides vary for different her-bicides [135]. The mechanisms of neurotoxicity of thechlorophenoxy herbicides are incompletely known, butthey seem to primarily involve cell membrane damage(reviewed in [135]).

    2,4-Dichlorophenoxyacetic acid (2,4-D) is the mostwidely used chlorophenoxy herbicide and also the mostwidely studied. With respect to membrane damage, it

    does not cause significant penetration of lipid monolayersin vitro at concentrations below 0,1 μM [134], but athigher concentrations (10–100 μM) it increases bilayerwidth and causes deep structural perturbations of thehydrophobic region of model membrane systems. At thehigher concentrations it also damages human erythrocytecell membranes [136]. This dose- dependent effect onplasma membranes may in part explain the dose-depend-ent CNS toxicity caused by chlorophenoxy herbicides. Inexperimental animals (rats, mice and rabbits), only smallamounts of herbicide were found in the brain followingadministration of 100 mg/kg or less [137-139], likelybecause low concentrations of herbicide have little effecton the plasma membranes comprising the blood-brainbarrier. When exposing rats to high doses (250–500 mg/kg) of the herbicide, a reversible selective damage to theblood-brain barrier occurred, and as a result serum albu-min and IgG could be detected in the brain along with theherbicide itself [140]. The severity of the herbicide-induced cerebrovascular damage in rats has been reportedto increase in the order 2,4,5-T (2,4,5-trichlorophenoxy-acetic acid) < MCPA (4-chloro-2-methylphenoxyaceticacid) < 2,4-D [141].

    Chlorophenoxy herbicides can also disrupt cell mem-brane transport mechanisms. They competitively inhibitand ultimately saturate the organic anion transport sys-tem in the choroid plexus, which facilitates the removal ofpotentially toxic anions (including endogenous neuro-transmitter metabolites and exogenous organic acids)[139,142,143]. Homovanillic acid and 5-hydroxy-3-indoleacetic acid, i.e. metabolites of dopamine and serot-onin, respectively, accumulate in the CNS of rats follow-ing 2,4-D administration [144].

    It has also been reported that 2,4-D induced neurotoxicitymay be partly due to generation of free radicals. Whenincubating rat cerebellar granule cells with 2,4-D in vitro,glutathione (GSH) levels and catalase activity were signif-icantly reduced, whereas generation of reactive oxygenspecies (ROS) and activity of selenium-glutathione perox-idase (Se-GPx) were augmented [145].

    Furthermore, chlorophenoxy acids are structurally relatedto acetic acid and are able to form analogues of acetyl-CoA(e.g. 2,4-D-CoA) in vitro. Formation of such analogues hasthe potential of disrupting several pathways involvingacetyl-CoA, including the synthesis of acetylcholine. Pos-sible formation of choline esters (e.g. 2,4-D-Ach) may actas false cholinergic messengers (reviewed in [135]).

    In cerebellar granule cells, 2,4-D produced a striking anddose-dependent inhibition of neurite extension, and invitro 2,4-D inhibited polymerisation of purified tubulin.Thus, it was suggested that at least one mechanism of 2,4-

    Page 9 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    D neurotoxicity involves inhibition of microtubuleassembly [146]. Yet another study with cerebellar granulecells showed that 2,4-D induced apoptosis when cellswere exposed to millimolar concentrations of the com-pound [147].

    Chlorophenoxy herbicide poisoning in humans isuncommon, but it may produce severe sequelae. In areview of 66 cases of chlorophenoxy herbicide poisoning[135], the majority of cases involved ingestion of 2,4-D,either alone or in combination with other chlorophenoxyherbicides. Neurotoxic effects included coma, hypertonia,hyperreflexia, ataxia, nystagmus, miosis, hallucinations,convulsions, fasciculations, and paralysis. Some degree ofperipheral neuromuscular involvement occurred inapproximately one third of the cases reviewed. Still, otherconstituents, such as surfactants or solvents, in the formu-lations of the herbicides could possibly have contributedto some of the effects observed [135].

    The information with respect to possible neurologicaleffects of chronic exposures to low doses of chlorophe-noxy herbicides is sparse, and in a review from 2002, itwas concluded that it is unlikely that 2,4-D has any neu-rotoxic potential at doses below those required to inducesystemic toxicity [148]. However, a cohort study suggestedan increased risk of amyotrophic lateral sclerosis (ALS)among workers chronically exposed to 2,4-D, comparedto non-exposed employees at the same company,although this conclusion was based on only three deaths[149].

    Although neurotoxicity in adults from low, chronic expo-sures to chlorophenoxy herbicides has not been reported,developmental exposure to low levels of these herbicidesmay still pose a threat. One case of cephalic malforma-tions and severe mental retardation has been observed inan infant whose parents received prolonged exposure to2,4-D via the dermal route from forest spraying [150].

    Evidence of developmental neurotoxicity of chlorophe-noxy herbicides, in particular 2,4-D, has also beenobtained from experimental animals. For example, exter-nal treatment of fertilised hens' eggs with 2,4-dichloroph-enoxyacetic butyl ester produced hypomyelination in thechicks, and reductions in "myelin markers" (includingsulfatides, cerebrosides and 2'3'-cyclic nucleotide 3'-phos-phohydrolase activity) were seen in chick embryos evenbefore the period of active myelination [151]. A deficit inmyelin lipid deposition was also detected in neonatal ratsexposed to 2,4-D through lactation [152]. Other findingsin response to neonatal exposure of rats to 2,4-D throughlactation include a delay in CNS development [153], anincrease in size and density of serotonin immunoreactiveneuronal somata as well as an increase in fibre length in

    the dorsal and medial raphe nuclei [154]; and oxidativestress in specific brain areas, including midbrain, stria-tum, and prefrontal cortex [155].

    Behavioural effects in the offspring have also beenreported following prenatal and continued exposure to2,4-D [156]. Also following prenatal and continued expo-sure of rats to 2,4-D, even beyond lactation, the dopamineD2-type receptor was increased about 40% in the striatum.Increased levels of the receptor were also found in the pre-frontal cortex and cerebellum. However, when discontin-uing exposure after weaning, no differences in dopamineD2-type receptors could be detected compared to controlrats, suggesting that the effects of 2,4-D on these receptorsmay be reversible [157].

    Thus, even though the evidence is sparse, some chloroph-enoxy herbicides, in particular 2,4-D, have neurotoxicpotentials and may cause developmental neurotoxicity.

    Bipyridyl herbicidesThe bipyridyl herbicides share common toxic mecha-nisms [158,159]; paraquat has been used as a model sub-stance, but is no longer allowed in the EU. Intracellularly,both paraquat and diquat undergo redox cycling, leadingto the generation of superoxide anions. These anions mayreact to form hydrogen peroxide and subsequently thehighly reactive hydroxyl radical, which may then causelipid peroxidation and cell death [159,160]. Another fac-tor contributing to toxicity is the depletion of nicotina-mide adenine dinucleotide phosphate with a boundhydrogen ion (NADPH), as both herbicide redox cyclingand hydrogen peroxide detoxification via glutathione areNADPH dependent [159,161]. In addition to redox-cycling, there is some evidence that paraquat may be ableto interact with enzymatic targets in the CNS, such asAChE and butylcholinesterase [162].

    The initial phase of moderate to severe intoxication withparaquat and diquat is characterised by renal and liverfailure, but the subsequent clinical course differs betweenthe two, with intestinal paralysis and fluid loss as promi-nent features of diquat intoxication [160,163-165]. Insevere and usually fatal cases of diquat poisoning, comahas also been reported [160]. Severe neurological andneuropsychiatric complications due to brain stem infarc-tion and/or intracranial haemorrhage have also beendescribed [161,163,166].

    In regard to long-term consequences of exposure to bipy-ridyl herbicides, paraquat is a prime suspect with respectto induction of PD. It causes selective degeneration oftyrosine hydroxylase immunopositive (TH+) neurons inthe substantia nigra pars compacta, and long-term expo-sure has been found to increase the risk of PD in a Taiwan

    Page 10 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    population that sprays paraquat on rice fields [167-169].A case report has described PD following diquat exposure[170], but because of a long induction period and the dif-ficulties in retrospective exposure assessment, the hypoth-esis of delayed appearance of degenerative nervous systemdisease is difficult to verify. Since both paraquat anddiquat can generate the formation of ROS, these com-pounds may well be involved in neurodegenerative dis-eases other than PD, such as Alzheimer's disease, but littleevidence is available to evaluate this potential.

    Even though it is rather clear that the cytotoxicity ofparaquat involves oxidative stress [171], the sensitivity ofdopaminergic neurons is difficult to explain [172]. Possi-bly, the dopaminergic neurons may be particularly sensi-tive to the reactive oxygen species (ROS) from paraquat,since dopamine metabolism also creates ROS [173]. Inmice treated with paraquat once a week for 3 weeks, theeffect on catecholaminergic neurons was reminiscent ofthat in PD, with a preferential loss of dopaminergic neu-rons in the substantia nigra pars compacta. This is consist-ent with the results from several similar studies[168,169,171].

    PD has also been explored as a relevant outcome withrespect to developmental neurotoxicity. When neonatalmice were exposed to paraquat, a marked hypoactive con-dition was apparent at 60 days of age and became evenmore pronounced at 120 days of age [174]. Furthermore,paraquat reduced the striatal content of dopamine andmetabolites without affecting serotonin [174]. As alreadymentioned above under dithiocarbamates, other evidencesuggests that maneb and paraquat may jointly and indi-vidually induce loss of dopaminergic neurons in mice.Administration of these pesticides postnatally enhancedthe effect of the same pesticides administered duringadulthood. Furthermore, exposure to maneb alone duringgestation resulted in a dramatically increased response toparaquat in adulthood, including notable reductions inlevels of dopamine and a loss of nigral dopamine neurons[132]. Similarly, the greatest effect on locomotor activityin mice occurred in males after exposure to maneb prena-tally and to paraquat in adulthood [175]. This finding wassupported by decreased levels of striatal dopamine,increased striatal dopamine turnover, and selective reduc-tion in tyrosine hydroxylase-immunoreactive neurons ofthe substantia nigra pars compacta.

    These observations are in agreement with the notion thatan initially silent toxicity was later unmasked, and wasaffected by the specific order-of-presentation of the pesti-cides in regard to the developmental stage (not just aneffect of the combination of pesticides). Thus, it seemsthat prenatal exposure to maneb, rather than paraquat,may sensitise/predispose mice to development of PD (or

    lead to a state of increased vulnerability), whereasparaquat exposure later in life may unmask the silent toxiceffect of the earlier maneb exposure and then lead to clin-ical symptoms of the disease. Therefore it is possible thatin the case of PD, developmental exposure to paraquatmay not be as damaging as later exposure, particularly ifthis later exposure follows developmental exposure tomaneb.

    FumigantsThe mechanisms of toxicity employed by various types offumigants are poorly known. A common mechanism ofaction is not expected, and the fumigants are thereforereviewed one by one.

    Among metal phosphide fumigants, aluminium phos-phide is one of the most extensively used. The phosphidesare very toxic, because of their ability to liberate phos-phine under moist conditions (reviewed in [176]). Phos-phine is a reductant and predictably reacts with metal ionssuch as the iron in haem and the divalent metals of metaldependent enzymes [177]. Cytochrome c oxidase, of themitochondrial electron transport chain, has been sug-gested as the primary site of action for phosphine[176,178,179]. A 50% inhibition of this enzyme wasfound to be sufficient for generation of superoxide anions,and it was suggested that the toxicity of phosphine wasdue to damage by free radicals [178]. In agreement withthis hypothesis, aluminium phosphide has been found toincrease lipid peroxidation in rat brain [180].

    Further, in 45 phosphine poisoning patients, increasedlevels of superoxide dismutase (SOD) and malondialde-hyde (MDA) were detected in non-survivors, while cata-lase was inhibited [181]. Oxidation of phosphine can leadto formation of reactive phosphorylating species [182],thus suggesting that effects on cholinesterase may be pos-sible [183]. Studies of grain fumigant applicators [184]and in vitro studies of human red blood cells [185] haveshown that significant phosphine-induced inhibition ofred blood cell cholinesterase occurs at concentrations ofphosphine exceeding 10 μg/ml.

    Neurological changes like ataxia, stupor, tremors and con-vulsions have been observed following aluminium phos-phide poisoning. Acute hypoxic encephalopathy has alsobeen observed following aluminium phosphide poison-ing, which may lead to death as a result of completedepression of the central nervous system and paralysis ofthe respiratory centres of the brain (reviewed in [176]).

    With respect to consequences of chronic phosphide expo-sure knowledge is sparse, but one descriptive studyreported that most of a group of workers exposed to zincphosphide had one or more neuropsychiatric symptoms

    Page 11 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    including anxiety, impotence and easy fatigue. About halfof the workers showed hyperreflexia, polyneuropathy,lumbar radiculopathy, and cervical myelopathy, as well asanxious mood, impaired attention, and psychomotorstimulation. EEG recordings showed abnormal findingsin 17.4% of the subjects, mainly those with longer expo-sure [186]. These preliminary findings should invite fur-ther studies in this area.

    For the fumigant sulfuryl fluoride, very little is knownconcerning the mechanism of toxicity. The fluoride ionmay play a role, since many of the observations in rodentsoverexposed to sulfuryl fluoride are typical of acute fluo-ride poisoning [187]. In humans, short-term inhalationexposure to high concentrations of sulfuryl fluoride hasbeen reported to cause central nervous system effects[188]. A case report describes an elderly couple, whoreturned to their home 5–8 hours after fumigation withsulfuryl fluoride. The wife experienced weakness, nausea,and repeated vomiting, while the husband complained ofdyspnea and restlessness. Within 48 hours the husbandhad a generalised seizure followed by cardiopulmonaryarrest. The wife died within 7 days due to ventricularfibrillation. The serum fluoride concentration of the wifesix days after the fumigation was reported to be as high as0.5 mg/L [189].

    Workers with a chronic, low level exposure to sulfuryl flu-oride showed non-significantly reduced performance onall applied neurobehavioural tests compared to the con-trol group in one study [190]. Education levels, ethnicityand drug use differed between the workers and the controlgroup in this study. In a later study of structural fumiga-tion workers [191], sulfuryl fluoride exposure during theyear preceding the examination was associated with sig-nificantly reduced performance on the Pattern MemoryTest (a test of cognitive and visual memory) and an olfac-tory test. No pattern of cognitive deficits was detected.

    None of these fumigants has been examined in detail forpossible developmental neurotoxicity. Pregnant rats andrabbits exposed to sulfuryl fluoride were reported to showno evidence of embryotoxicity, foetotoxicity, or tera-togenicity at concentrations of sulfuryl fluoride as high as225 ppm, although body weights of rabbit foetuses as wellas the dams at the highest exposure were lower than in thecontrol group [192].

    In regard to phosphine, a large epidemiological studyfound that adverse neurological and neurobehaviouraldevelopmental effects clustered among children fatheredby applicators of phosphine (odds ratio = 2,48; 95% con-fidence interval: 1.2, 5.1) [193]. Other than this study, noinformation regarding developmental neurotoxicity ofphosphine was identified.

    Other pesticidesThe present review on neurotoxicity has focused on asmall number of substances out of the total numberapproved for use as pesticides in the EU. Quite likely,much evidence exists on neurotoxicity, but has not beenpublished in biomedical journals. Nicotine, warfarin andethanol are additional well documented neurotoxicants,but their primary use is not as pesticides. The same appliesto other substances listed, such as sodium hypochloriteand aluminium sulfate, which may potentially add toneurotoxic hazards.

    Public health implicationsSome of the substances belonging to the groups of pesti-cides reviewed here (including OPs, carbamates, pyre-throids, ethylenebisdithiocarbamates, chlorophenoxyherbicides, and bipyridyl herbicides) appear to share com-mon mechanisms of action with respect to induction ofneurotoxicity. Thus, members of these chemical groups ofpesticides other than those identified as neurotoxic in thepresent review, would then be highly likely also to causeneurotoxicity. For other groups of pesticides without aplausible common mechanism of action (e.g. the fumi-gants), it is not possible to predict whether group mem-bers might share neurotoxicity potentials.

    Further refinement of this prediction is difficult. As antic-ipated, the literature on developmental neurotoxicity issparse for most of the pesticides. However, some evidencedoes exist to suggest that several of the neurotoxic pesti-cides in current use in the EU may cause developmentalneurotoxicity in small doses. Table 2 summarises theexisting evidence of developmental neurotoxicity forgroups of pesticides with common mechanisms of action.

    Most evidence is available for the OPs, especially chlorpy-rifos. The evidence strongly supports the notion thatdevelopmental neurotoxicity may be induced by very lowexposure levels, i.e. much below those causing any neuro-toxicity in adults. Most evidence still comes from studiesin laboratory animals, but some epidemiological data arehighly suggestive of neurotoxic effects caused by develop-mental exposure of humans to OPs (including chlorpyri-fos). In the case of OPs, which share inhibition of AChEas a common mechanism of action in high doses, chlorpy-rifos may employ other mechanisms of action at lowerdoses associated with developmental neurotoxicity. Infact developmental neurotoxicity in mice and rats can beinduced at doses, which cause no detectable inhibitionAChE [43-46,51]. Thus, even though a group of pesticidesshares a common mechanism of action at larger doses, itcannot be excluded that compound specific mechanismsmay also exist at lower doses. This fact unfortunately com-plicates the extrapolation of developmental neurotoxicityfrom one member of a group of pesticides to another.

    Page 12 of 22(page number not for citation purposes)

  • Page

    13

    of 2

    2(p

    age

    num

    ber n

    ot fo

    r cita

    tion

    purp

    oses

    )

    eurotoxicity

    tal neurotoxicity animals

    References

    amming of synaptic in rats (Chlorpyrifos)

    [50,51]

    bnormalities including omotor skills and ormance in rats and rifos)

    [43-46]

    ere found

    tor activity, lack of hanges in mAChR e

    [85-88]

    ges in rats [89]

    otor activity in rats [90]

    xual behaviour and of the dopaminergic

    [91]

    AChR expression in [92,93]

    ood-brain permeability [94]

    Env

    ironm

    enta

    l Hea

    lth 2

    008,

    7:5

    0ht

    tp://

    ww

    w.e

    hjou

    rnal

    .net

    /con

    tent

    /7/1

    /50

    Table 2: Evidence of developmental neurotoxicity caused by pesticides belonging to groups with likely common mechanisms of n

    Group of pesticides (n)* Mechanism of neurotoxicity Developmental neurotoxicity reported in humans

    References Developmenreported in

    Organophosphates (8) Inhibition of AChE (+ interference with signaling cascades at low doses)

    Reflex abnormalities in neonates + affected mental development

    [7,8] Altered progrdevelopment

    Reduced head circumference in infants + anomalies in primitive reflexes (Chlorpyrifos)

    [61,9] Behavioural achanges in loccognitive perfmice (Chlorpy

    Reduced birth weight and length + developmental delay at 3 years of age (Chlorpyrifos)

    [62,10]

    Visuospatial deficits (prenatal exposure) + increased reaction time (current exposure in children)

    [11]

    Reduced short term memory and attention (Methyl parathion)

    [12]

    Carbamates (5) Inhibition of AChE (+ oxidative stress)

    No reports were found No reports w

    Pyrethroids (7) Prolongation of kinetics of voltage-gated sodium channels

    Increased mohabituation, cdensity in mic

    Learning chan

    Changes in m

    Changes in sehigher activitysystem in rats

    Changes in mrats

    Changes in blin rats

  • Page

    14

    of 2

    2(p

    age

    num

    ber n

    ot fo

    r cita

    tion

    purp

    oses

    )

    lopment of reflexes, ity, open field activity in

    sure prior to mating)

    [96]

    bination with ces loss of

    neurons in substantia pacta in mice

    [132]

    e of EBDCs, ETU, rmations of the nervous sponding to thyroid in rats

    Reviewed in [126]

    ion in chicks [151]

    lin lipid deposition in [152]

    development in rats [153]

    and densitiy of ctive neuronal somata fiber length in dorsal phe nuclei in rats

    [154]

    ss in specific brain in, striatum, prefrontal s

    [155]

    ffects in rats including ng reflex, negative tor abnormalities,

    oming and vertical head yperactivity

    [156]

    eurotoxicity (Continued)

    Env

    ironm

    enta

    l Hea

    lth 2

    008,

    7:5

    0ht

    tp://

    ww

    w.e

    hjou

    rnal

    .net

    /con

    tent

    /7/1

    /50

    Affected deveswimming abilmice (parental expo

    Dithiocarbamates (EBDCs) (6)

    Generation of ROS (metal chelating capacity, uncoupling of mitochondrial electron transport chain)The EBDC metabolite, ETU, inhibits thyroid peroxidase (synthesis of thyroid hormones)

    Maneb (in comparaquat) indudopaminergicnigra pars com

    The metabolitinduces malfosystem (correinsufficiency)

    Chlorophenoxy herbicides (11)

    Not completely known: includes membrane damage, generation of free radicals, perhaps uncoupling of oxidative phosphorylation

    A case of cephalic malformations and severe mental retardation in infant whose parents were heavily exposed to 2,4-D

    [150] Hypomyelinat(2,4-D)

    Deficit in myerats(2,4-D)

    Delayed CNS(2,4-D)

    Increased sizeserotonin-reaand increasedand medial ra(2,4-D)

    Oxidative streareas (midbracortex) in rat(2,4-D)

    Behavioural edelay of rightigeotaxis + moexcessive gromovements, h(2,4-D)

    Table 2: Evidence of developmental neurotoxicity caused by pesticides belonging to groups with likely common mechanisms of n

  • Page

    15

    of 2

    2(p

    age

    num

    ber n

    ot fo

    r cita

    tion

    purp

    oses

    )

    f developmental araquat in later of PD like features in

    [174]

    ombination with es loss of dopaminergic bstantia nigra pars ice

    [132]

    major evidence on developmental neurotoxicity in

    eurotoxicity (Continued)

    Env

    ironm

    enta

    l Hea

    lth 2

    008,

    7:5

    0ht

    tp://

    ww

    w.e

    hjou

    rnal

    .net

    /con

    tent

    /7/1

    /50

    Bipyridyl herbicides (1) Induction of oxidative stress Involvement oexposure to pdevelopment mice

    Paraquat (in cmaneb) inducneurons in sucompacta in m

    *The number in parenthesis is the total number of pesticides from each group currently authorised for use in the EU as of August 2008. Only humans or in laboratory animals has been included.

    Table 2: Evidence of developmental neurotoxicity caused by pesticides belonging to groups with likely common mechanisms of n

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    However, the combined human evidence on develop-mental neurotoxicity associated with OP exposure cannotbe ascribed to chlorpyrifos alone.

    Other than for OPs, the evidence of developmental neu-rotoxicity in humans is sparse, but evidence on develop-mental neurotoxicity in laboratory animals exists forpyrethroids, ethylenebisdithiocarbamates, and chloroph-enoxy herbicides (mainly 2,4-D).

    In the case of dithiocarbamates, evidence from laboratoryanimals suggests that developmental exposure to, e.g.maneb, may predispose the individual to development ofPD later in life in response to another exposure, in partic-ular paraquat. Other experimental studies suggest thatprenatal exposure to paraquat can also predispose todevelopment of PD later in life. It seems that the greatesteffect of paraquat with respect to induction of PD isobtained from exposure later in life, following early prim-ing exposure to maneb [175]. Although PD is a degenera-tive disease associated with aging, these data suggest thatdevelopmental exposure to pesticides (e.g. maneb) mayconstitute an aetiological factor that sensitises the individ-ual to later insults (e.g. subsequent pesticide exposure,and aging).

    For the remaining pesticides that belong to groups with-out a common mechanism of toxicity, the lack of researchon developmental neurotoxicity complicates the evalua-tion of their safety. In a few cases (e.g. the fumigant sulfu-ryl fluoride), the existing evidence from animalexperiments indicates that developmental neurotoxicitymay be unlikely to occur at doses below those causingmaternal toxicity [192,194]. Still, in these experiments,possible later emerging effects or sensitisation caused bydevelopmental exposure has not been studied, so anyconclusion in this regard would be tentative.

    On the other hand, with respect to the metal phosphidefumigants, which release phosphine under moist condi-tions, some evidence of developmental neurotoxicity doesexist. An epidemiologic study has found adverse neuro-logical and neurodevelopmental effects among childrenfathered by applicators of phosphine [193]. For theremaining pesticides reviewed, no data from eitherhuman or animal studies could be located by our search.

    This review has focused on those pesticides, for whichhuman neurotoxicity has been reported in relation to spe-cific exposures to the particular pesticide. This means thatwe have excluded poisoning cases involving more thanone compound, where the contribution by each substancemay be unknown. Thus, our list of neurotoxic pesticides islikely a substantial underestimate of the true number ofneurotoxic pesticides. The fact that no poisoning incident

    with neurotoxic effects has been reported for a given pes-ticide is of course no guarantee that the pesticide is notneurotoxic, especially in regard to developmental expo-sure. A prudent evaluation of the evidence would there-fore suggest that, if individual members of a chemicalgrouping of pesticides have been documented as neuro-toxic, then all members of that group should be consid-ered to be neurotoxic as well.

    In addition to the problem of scarce – in many cases evennon-existing – scientific evidence on developmental neu-rotoxicity of the pesticides in current use, some discrepan-cies exist between results of animal studies. An importantfactor in regard to apparent discrepancies is that the tim-ing of exposure varies between studies. In some studies,animals are exposed prenatally, in other studies neona-tally (during the first weeks of life), and in some studiesboth prenatally and neonatally. The timing of exposuremay greatly influence the extent and type of neurotoxicityinduced. Most animal studies have been performed inrodents, where brain development is mainly neonatal andspans the first three to four weeks of postnatal life[14,195]. Thus, although neurotoxic effects may beinduced in rodents by only prenatal exposure, it is highlylikely that these studies underestimate the neurotoxiceffects, which may occur in response to prenatal exposureof humans, where the third trimester of pregnancy is a cru-cial period of brain development.

    A further concern is that humans are very likely to beexposed to a number of pesticides and other neurotoxiccompounds simultaneously. Because it is possible thatsome of these may have synergistic or additive effects,exposure to even very low doses during development maycause neurotoxic damage.

    In addition to "direct" neurotoxicity, there is also evidencethat several pesticides may indirectly cause neurotoxicity,e.g. by interference with thyroid function. Some 60% ofall herbicides, in particular 2,4-D, acetochlor, aminotria-zole, amitrole, bromoxynil, pendamethalin, and thiou-reas have been reported to interfere with thyroid function(reviewed in[196]). In addition, EBDC dithiocarbamates,organophosphates and synthetic pyrethroids are thoughtto interfere with thyroid function (reviewed in [197]). Akey concern with thyroid inhibitors is that impaired thy-roid function may alter hormone-mediated events duringdevelopment, leading to permanent alterations in brainmorphology and function [127,128]. Other types ofendocrine disruption can conceivably lead to neurobe-havioural deficits, but this evidence has not been includedhere.

    The current evidence can therefore be summed up as fol-lows. A substantial proportion of pesticides in current use

    Page 16 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    are known to be neurotoxic. However, neurotoxicitypotentials of pesticides have not necessarily been exam-ined, as legally mandated tests do not require specificassessment of neurotoxic potentials, apart from tests forperipheral neurotoxicity in hens required for OPs. A testbattery for developmental neurotoxicity has only recentlybeen completed by OECD, and very limited test data areavailable for pesticides. Because developmental neurotox-icity can occur at exposures much below those that causetoxicity to the adult brain, usage restrictions and legal lim-its for pesticide residues in food may not be sufficientlyprotective against developmental neurotoxicity. In addi-tion, experimental, clinical and epidemiologic evidencesupports the notion that neurotoxicity may be much moresevere and possibly irreversible when the exposure occursduring early development.

    Unless documentation exists for a particular pesticide tofalsify this notion, all neurotoxic pesticides should beconsidered likely of inducing developmental neurotoxic-ity at low doses. The public health significance of thisissue is illustrated by the epidemiologic observation ofneurodevelopmental deficits at exposure levels that seemto be commonly occurring in the general population.Although the exact identity of the causative substancesmay be uncertain, pesticide contamination of foods iscommon in the EU, it often exceeds previously identifiedlegal limits, and it involves substances that are known tobe neurotoxic. Given the substantial impact of neurode-velopmental abnormalities in society and the likelyimpact of environmental aetiologies, prevention of pesti-cide exposure appears to be an obvious public health pri-ority.

    ConclusionGiven the widespread use and exposure to pesticides, thegeneral lack of data on developmental neurotoxicity is aserious impediment. For certain pesticides, a requirementexists for neurotoxicity tests in adult animals, but develop-mental neurotoxicity is usually not considered whendetermining pesticide safety. Experimental, clinical, andepidemiologic evidence suggests that neurotoxic pesti-cides can also cause developmental neurotoxicity, andthat the effects are more severe and lasting, and that theyoccur at much lower exposure levels. Some of this evi-dence relates to model substances that have now beenbanned or restricted, but currently used substances withsimilar mechanisms of toxicity should be regarded toshare the same toxic potentials. Thus, many widely usedpesticides, such as organophosphates, carbamates, pyre-throids, ethylenebisdithiocarbamates, and chlorophe-noxy herbicides should be consideredneurodevelopmental toxicants, unless convincing evi-dence exists for individual substances that they deviatefrom the general group characteristics. Given the likely

    environmental aetiology of neurodevelopmental deficitsand their importance to families and to society, preven-tion of exposures to neurotoxic pesticides should be madea public health priority. Existing uncertainties should notbe used as an excuse for rejecting precautionary action.

    AbbreviationsACh: Acetylcholine; AChE: Acetylcholinesterase; ADHD:Attention Deficit Hyperactivity Disorder; AMP: Adenosinemonophosphate; ALS: Amyotrophic Lateral Sclerosis;CNS: Central Nervous System; CREB Ca2+/cAMPResponse Element Binding protein; CT: Computed Tom-ography; 2,4-D: 2,4-Dichlorophenoxyacetic acid; DEDC:Diethyldithiocarbamate; EBDC: Ethylenebisdithiocar-bamate; EEG: Electroencephalogram; ETU: Ethyl-enethiourea; EU: European Union; GABA: Gamma-aminobutyric acid; GSH: Glutathione; HSDB: HazardousSubstances Data Bank; IgG: Immunoglobulin G; mAChR:muscarinic acetylcholine receptor; MCPA: 4-chloro-2-methylphenoxyacetic acid; MDA: Malondialdehyde;MPTP: 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine;MRI: Magnetic Resonance Imaging; NADPH: Nicotina-mide Adenine Dinucleotide Phosphate with a boundHydrogen ion; NIOSH: National Institute of Occupa-tional Safety and Health; NTE: Neuropathy Target Este-rase; OECD: Organisation for Economic Co-operationand Development; OP: Organophosphate; OPIDP: Orga-noPhosphate-Induced Delayed Polyneuropathy; PC12cells: Cancer cell line from a pheochromocytoma of therat adrenal medulla; ROS: Reactive Oxygen Species; Se-GPx: Selenium-glutathione peroxidase; SOD: Superoxidedismutase; 2,4,5-T: 2,4,5-trichlorophenoxyacetic acid;TH+: Tyrosine Hydroxylase immunopositive; TPX: Thy-roid peroxidase.

    Competing interestsPG is an editor of Environmental Health but was notinvolved in the editorial handling of this manuscript. Theauthors declare that they have no competing interests.

    Authors' contributionsMBP, HRA and PG jointly conceived the review, MBP andHRA from mechanistic and toxicologic considerationsand PG from an epidemiologic viewpoint. MBP con-ducted the literature survey and wrote the first draft,which all authors revised and updated. The final manu-script was approved by all authors.

    AcknowledgementsThis study was supported by a European Commission grant to the HENVI-NET project (Contract No. 037019 coordinated by the Norwegian Insti-tute for Air Research). We thank Drs Gemma Calamandrei, Lilian Corra, Janna Koppe and Margaret Saunders for comments on an earlier draft of this manuscript.

    Page 17 of 22(page number not for citation purposes)

  • Environmental Health 2008, 7:50 http://www.ehjournal.net/content/7/1/50

    References1. The use of plant protection products in the European Union.

    Data 1992–2003. Eurostat statistical books [http://epp.eurostat.ec.europa.eu/cache/ITY_OFFPUB/KS-76-06-669/EN/KS-76-06-669-EN.PDF]

    2. Monitoring of Pesticide Residues in Products of Plant Originin the European Union, Norway, Iceland and Liechtenstein,2005. Commision staff working document [http://ec.europa.eu/food/fvo/specialreports/pesticide_residues/report_2005_en.pdf]

    3. Mage DT, Allen RH, Gondy G, Smith W, Barr DB, Needham LL: Esti-mating pesticide dose from urinary pesticide concentrationdata by creatinine correction in the Third National Healthand Nutrition Examination Survey (NHANES-III). J Expo AnalEnviron Epidemiol 2004, 14:457-465.

    4. Dobbing J: Vulnerable periods in developing brain. In AppliedNeurochemistry Edited by: Davison AN, Dobbing J. Philadelphia: Davis;1968:287-316.

    5. Rodier PM: Developing brain as a target of toxicity. EnvironHealth Perspect 1995, 103(Suppl 6):73-76.

    6. Rice D, Barone S Jr: Critical periods of vulnerability for thedeveloping nervous system: evidence from humans and ani-mal models. Environ Health Perspect 2000, 108(Suppl 3):511-533.

    7. Young JG, Eskenazi B, Gladstone EA, Bradman A, Pedersen L, JohnsonC, Barr DB, Furlong CE, Holland NT: Association between inutero organophosphate pesticide exposure and abnormalreflexes in neonates. Neurotoxicology 2005, 26:199-209.

    8. Eskenazi B, Marks AR, Bradman A, Harley K, Barr DB, Johnson C,Morga N, Jewell NP: Organophosphate pesticide exposure andneurodevelopment in young Mexican-American children.Environ Health Perspect 2007, 115:792-798.

    9. Engel SM, Berkowitz GS, Barr DB, Teitelbaum SL, Siskind J, Meisel SJ,Wetmur JG, Wolff MS: Prenatal organophosphate metaboliteand organochlorine levels and performance on the Brazel-ton Neonatal Behavioral Assessment Scale in a multiethnicpregnancy cohort. Am J Epidemiol 2007, 165:1397-1404.

    10. Rauh VA, Garfinkel R, Perera FP, Andrews HF, Hoepner L, Barr DB,Whitehead R, Tang D, Whyatt RW: Impact of prenatal chlorpy-rifos exposure on neurodevelopment in the first 3 years oflife among inner-city children. Pediatrics 2006, 118:e1845-e1859.

    11. Grandjean P, Harari R, Barr DB, Debes F: Pesticide exposure andstunting as independent predictors of neurobehavioral defi-cits in Ecuadorian school children. Pediatrics 2006,117:e546-e556.

    12. Ruckart PZ, Kakolewski K, Bove FJ, Kaye WE: Long-term neu-robehavioral health effects of methyl parathion exposure inchildren in Mississippi and Ohio. Environ Health Perspect 2004,112:46-51.

    13. Claudio L, Kwa WC, Russell AL, Wallinga D: Testing methods fordevelopmental neurotoxicity of environmental chemicals.Toxicol Appl Pharmacol 2000, 164:1-14.

    14. Eriksson P: Developmental neurotoxicity of environmentalagents in the neonate. Neurotoxicology 1997, 18:719-726.

    15. Tilson HA: The concern for developmental neurotoxicology:is it justified and what is being done about it? Environ Health Per-spect 1995, 103(Suppl 6):147-151.

    16. Grandjean P, Landrigan PJ: Developmental neurotoxicity ofindustrial chemicals. Lancet 2006, 368:2167-2178.

    17. Schettler T: Toxic threats to neurologic development of chil-dren. Environ Health Perspect 2001, 109(Suppl 6):813-816.

    18. Boyle CA, Decoufle P, Yeargin-Allsopp M: Prevalence and healthimpact of developmental disabilities in US children. Pediatrics1994, 93:399-403.

    19. Landrigan PJ, Schechter CB, Lipton JM, Fahs MC, Schwartz J: Envi-ronmental pollutants and disease in American children: esti-mates of morbidity, mortality, and costs for lead poisoning,asthma, cancer, and developmental disabilities. Environ HealthPerspect 2002, 110:721-728.

    20. Tilson HA: Neurotoxicology risk assessment guidelines:developmental neurotoxicology. Neurotoxicology 2000,21:189-194.

    21. Costa LG: Current issues in organophosphate toxicology. ClinChim Acta 2006, 366:1-13.

    22. Lauder JM, Schambra UB: Morphogenetic roles of acetylcholine.Environ Health Perspect 1999, 107(Suppl 1):65-69.

    23. Clegg DJ, van Gemert M: Determination of the reference dosefor chlorpyrifos: proceedings of an expert panel. J Toxicol Envi-ron Health B Crit Rev 1999, 2:211-255.

    24. Lotti M: Clinical Toxicology of Anticholinesterase Agents inHumans. In Handbook of Pesticide Toxicology Volume 2. 2nd edition.Edited by: Krieger RI. San Diego: Academic Press; 2001:1043-1085.

    25. Lotti M: The pathogenesis of organophosphate polyneuropa-thy. Crit Rev Toxicol 1991, 21:465-487.

    26. Johnson MK, Glynn P: Neuropathy target esterase (NTE) andorganophosphorus-induced delayed polyneuropathy(OPIDP): recent advances. Toxicol Lett 1995, 82–83:459-463.

    27. Lotti M, Moretto A: Organophosphate-induced delayedpolyneuropathy. Toxicol Rev 2005, 24:37-49.

    28. Pope CN, Chakraborti TK, Chapman ML, Farrar JD: Long-termneurochemical and behavioral effects induced by acute chlo-rpyrifos treatment. Pharmacol Biochem Behav 1992, 42:251-256.

    29. Sanchez-Santed F, Canadas F, Flores P, Lopez-Grancha M, Cardona D:Long-term functional neurotoxicity of paraoxon and chlo-rpyrifos: behavioural and pharmacological evidence. Neuro-toxicol Teratol 2004, 26:305-317.

    30. Rosenstock L, Keifer M, Daniell WE, McConnell R, Claypoole K:Chronic central nervous system effects of acute organophos-phate pesticide intoxication. The Pesticide Health EffectsStudy Group. Lancet 1991, 338:223-227.

    31. Steenland K: Chronic neurological effects of organophosphatepesticides. BMJ 1996, 312:1312-1313.

    32. Kamel F, Hoppin JA: Association of pesticide exposure withneurologic dysfunction and disease. Environ Health Perspect2004, 112:950-958.

    33. Jamal GA, Hansen S, Julu PO: Low level exposures to organo-phosphorus esters may cause neurotoxicity. Toxicology 2002,181–182:23-33.

    34. Colosio C, Tiramani M, Maroni M: Neurobehavioral effects ofpesticides: state of the art. Neurotoxicology 2003, 24:577-591.

    35. Rothlein J, Rohlman D, Lasarev M, Phillips J, Muniz J, McCauley L:Organophosphate pesticide exposure and neurobehavioralperformance in agricultural and non-agricultural Hispanicworkers. Environ Health Perspect 2006, 114:691-696.

    36. Kamel F, Engel LS, Gladen BC, Hoppin JA, Alavanja MC, Sandler DP:Neurologic symptoms in licensed pesticide applicators in theAgricultural Health Study. Hum Exp Toxicol 2007, 26:243-250.

    37. Lauder JM: Roles for neurotransmitters in development: pos-sible interaction with drugs during the fetal and neonatalperiods. Prog Clin Biol Res 1985, 163C:375-380.

    38. Whitaker-Azmitia PM: Role of serotonin and other neurotrans-mitter receptors in brain development: basis for develop-mental pharmacology. Pharmacol Rev 1991, 43:553-561.

    39. Pope CN, Chakraborti TK: Dose-related inhibition of brain andplasma cholinesterase in neonatal and adult rats followingsublethal organophosphate exposures. Toxicology 1992,73:35-43.

    40. Pope CN, Chakraborti TK, Chapman ML, Farrar JD, Arthun D: Com-parison of in vivo cholinesterase inhibition in neonatal andadult rats by three organophosphorothioate insecticides.Toxicology 1991, 68:51-61.

    41. Song X, Seidler FJ, Saleh JL, Zhang J, Padilla S, Slotkin TA: Cellularmechanisms for developmental toxicity of chlorpyrifos: tar-geting the adenylyl cyclase signaling cascade. Toxicol Appl Phar-macol 1997, 145:158-174.

    42. Slotkin TA: Cholinergic systems in brain development and dis-ruption by neurotoxicants: nicotine, environmental tobaccosmoke, organophosphates. Toxicol Appl Pharmacol 2004,198:132-151.

    43. Dam K, Seidler FJ, Slotkin TA: Chlorpyrifos exposure during acritical neonatal period elicits gender-selective deficits in thedevelopment of coordination skills and locomotor activity.Brain Res Dev Brain Res 2000, 121:179-187.

    44. Ricceri L, Markina N, Valanzano A, Fortuna S, Cometa MF, MeneguzA, Calamandrei G: Developmental exposure to chlorpyrifosalters reactivity to environmental and social cues in adoles-cent mice. Toxicol Appl Pharmacol 2003, 191:189-201.

    45. Icenogle LM, Christopher NC, Blackwelder WP, Caldwell DP, QiaoD, Seidler FJ, Slotkin TA, Levin ED: Behavioral alterations in ado-lescent and adult rats caused by a brief subtoxic exposure tochlorpyrifos during neurulation. Neurotoxicol Teratol 2004,26:95-101.

    Page 18 of 22(page number not for citation purposes)

    http://epp.eurostat.ec.europa.eu/cache/ITY_OFFPUB/KS-76-06-669