THE NEMATODE CAENORHABDITIS ELEGANS AS A MODEL OF

ORGANOPHOSPHATE INDUCED MAMMALIAN NEUROTOXICITY

by

RUSSELL DAVID COLE

(Under the Direction of Phillip L. Williams)

ABSTRACT

Research has shown a number of basic similarities between the nervous systems

of the nematode Caenorhabditis elegans and higher animals including mammals. The purpose of this study was to investigate the toxicity of a chemical class known to be neurotoxic in C. elegans and compare this toxicity to that seen in mammals. The behavioral response of C. elegans to 15 (OP) pesticides was characterized using computer automated tracking. Toxicity was ranked and compared to the ranked toxicity of the same compounds in rats and mice. Toxic mechanism was also examined through cholinesterase activity assays performed on worms exposed to 8 of the

15 OPs. A significant correlation was found between rank order of toxicity in C. elegans and rats and mice. Cholinesterase inhibition by some OPs was also confirmed in C. elegans. Specific comparisons and implications concerning C. elegans’ potential as a

mammalian neurotoxicity model are discussed.

INDEX WORDS: Caenorhabditis elegans, Organophosphate, Neurotoxicity, Rats, Mice, Behavior, Cholinesterase, Biological Model

THE NEMATODE CAENORHABDITIS ELEGANS AS A MODEL OF

ORGANOPHOSPHATE INDUCED MAMMALIAN NEUROTOXICITY

by

RUSSELL DAVID COLE

B.S., The University of Georgia, 1998

A Thesis Submitted to the Graduate Faculty of The University of Georgia in Partial

Fulfillment of the Requirements for the Degree

MASTER OF SCIENCE

ATHENS, GEORGIA

2003

© 2003

Russell David Cole

All Rights Reserved

THE NEMATODE CAENORHABDITIS ELEGANS AS A MODEL OF

ORGANOPHOSPHATE INDUCED MAMMALIAN NEUROTOXICITY

by

RUSSELL DAVID COLE

Major Professor: Phillip L. Williams

Committee: Cham E. Dallas Jeffrey W. Fisher

Electronic Version Approved:

Maureen Grasso Dean of the Graduate School The University of Georgia August 2003

iv

DEDICATION

I have heard that J.S. Bach dedicated every piece of music he wrote to the glory of

God. I do not wish to draw comparisons, but only to follow the wisdom of his actions. I dedicate this thesis to the glory of my God, Jesus Christ. Over the last two years, as in

the rest of my life, his love has been unwavering, his grace overwhelming, his mercy

sustaining, his enablement sufficient, and his provision complete. It is he who has

brought me this far, and it is he who will take me on. v

ACKNOWLEDGEMENTS

I would like to thank Dr. Phillip Williams for serving as my advisor for this thesis

work, and to Dr. Jeffery Fisher and Dr. Cham Dallas for serving on my thesis committee.

I would also like to thank Dr. Gary Anderson, Dr. Windy Boyd, and Meagan Kane for their help in the lab. Thanks is also due to Deanna Conners for her help with the cholinesterase assay protocol, and to Dr. Marsha Black for the use of her lab during this portion of my investigation. Thanks to Sandra McPeake, Ella Willingham, and Regina

Davis for all of their help in various administrative and accounting endeavors. I would also like to offer a general thanks to the faculty, staff, and students of the Environmental

Health Science Department for an enjoyable environment in which to work over the past two years.

I would like to thank my wife, Sandy, for her love, steady encouragement,

unwavering support, and patience in sharing me with nematodes while I have pursued

this degree. I thank my family for their support and willingness over the last two years to

frequently sacrifice my presence in deference to this work. I would also like to thank my

church family at the Athens Vineyard for their steadfast prayers and support. I owe a

particular debt to Dr. Steve Holloway for his listening ear, ready encouragement, sound

advice, and overall friendship over these last years.

Lastly, I would like to thank Dr. Michael Petelle who taught me 10th grade

Biology with such excellence and enthusiasm as to set the standard for my studies in the

biological sciences. vi

TABLE OF CONTENTS

Page

ACKNOWLEDGEMENTS...... v

LIST OF TABLES...... viii

LIST OF FIGURES ...... ix

CHAPTER

1 INTRODUCTION ...... 1

References ...... 4

2 LITERATURE REVIEW...... 5

References ...... 28

3 THE NEMATODE CAENORHABDITIS ELEGANS AS A

MODEL OF ORGANOPHOSPHATE INDUCED MAMMALIAN

NEUROTOXICITY...... 49

Abstract ...... 50

Introduction ...... 51

Materials and Methods ...... 54

Results ...... 60

Discussion ...... 62

Acknowledgements ...... 72

References ...... 73

4 CONCLUSIONS...... 86 vii

LIST OF TABLES

Page

Table 2.1. Basic Description of Chemicals Tested...... 41

Table 3.1. Toxicity values and results for Spearman’s Rank Order Correlation comparison………………………..……………….79 viii

LIST OF FIGURES

Page

Figure 3.1. Comparison of toxicity values for C. elegans (behavioral EC50s) to rats and mice (LD50s) ...... 80

Figure 3.2. Graphical comparison of ChE activity levels in nmols activity/mg protein/minute...... 81

Figure 3.3. Response of movement behavior to decreasing exposure solution pH ...... 82

Figure 3.4. Graphical comparison of ChE activity levels expressed as % control activity………………………………………...... 83

Figure 3.5. Oxidative desulfuration of the ...... 84

Figure 3.6. Organophosphate induced change in movement versus exposure solution pH...... 85

1

CHAPTER 1

INTRODUCTION

In the decades since the so called “Green Revolution,” our ability to synthesize organic chemicals has grown at a rate beyond our ability to assess the benefits and dangers of newly discovered compounds. As our understanding of biology and chemistry has grown, we have also been faced with the need to return to previously studied chemicals to better understand their actions and fates in the body and the environment. In this setting there is an ever present need for testing the toxicity of chemicals to humans.

Because of the unethical nature of conducting most toxicity tests in humans, the majority of testing is carried out using animals, primarily mammals, or their tissues. This solution also creates problems. Animal testing is expensive, time consuming, and is in some cases deemed unacceptable in its toll on animal life. In an effort to limit the number of animal tests carried out, three main types of alternative testing are often undertaken as preliminary toxicity screening tools. Structure-activity relationships seek to predict toxicity based upon the toxicity of well characterized compounds with similar chemical structure. This does not directly require the use of animals, but can also be unreliable. In vitro testing provides for a more efficient use of animals and is selective for the particular tissues in question. In vitro tests, however, completely eliminate many pharmacokinetic factors important to toxicity such as absorption, distribution, and biotransformation.

Toxicity testing using alternative organisms includes aspects of pharmacokinetics and 2

pharmacodynamics into the toxicity model, but care must be taken to ensure that these

factors in the test organism accurately model the chemical behavior, target organ, and site

of action in humans.

This purpose of this thesis is to continue in the exploration of the nematode

Caenorhabditis elegans as an alternative testing organism for some forms of

neurotoxicity in mammals. This question is addressed through toxicity testing of the

known neurotoxic organophosphate pesticides in C. elegans, and comparison of the

results of this testing to published mammalian toxicity values. The second chapter of this

thesis is a review of the literature pertinent to this modeling question. It includes

discussions of C. elegans and its neurological similarities to mammals as well as a

discussion of organophosphate toxicity in mammals and C. elegans. Chapter 3 describes

a series of experiments undertaken to compare acute organophosphate behavioral toxicity

in C. elegans to lethality in mammals.

The hypothesis of this work was that sublethal behavioral toxicity in C. elegans

exposed to should bear enough similarity to mammalian toxicity for C.

elegans to be useful as a model in OP toxicity in mammals. To address this question, 15

organophosphates of varying mammalian toxicity were tested for their affect on C.

elegans movement behavior following 4-hour aquatic exposures. Triplicate measures of movement by worms exposed to five OP concentrations and a blank control were used to characterize the concentration-response of C. elegans to each compound. Assessment of movement behavior was carried out with a computer automated tracking program interfaced with a video camera continuously collecting images of nematodes under dark field illumination. Concentration-response data were used to estimate the chemical 3

concentration effectively reducing C. elegans movement by 50% (EC50). These EC50 values were compared to rat and mouse LD50 values obtained from the RTECS database

using Spearman’s Rank Order Correlation Coefficient.

Organophosphate insecticides exert toxicity primarily by the inhibition of acetylcholinesterase (AChE) in the nervous system. Although the literature includes tests with AChE mutants which seem to indicate that AChE inhibition also occurs in C.

elegans, this experiment measured directly cholinesterase activity in worms. Worms

were exposed to eight of the fifteen OPs tested at concentrations approximating the EC50

and high concentration used in behavioral toxicity tests. Cholinesterase activity was

quantified in control and exposed worms following 4-hour exposures using a

modification of Ellman’s procedure (1961). Exposed groups were compared to controls

using analysis of variance. This chapter also addresses the affect of different

organophosphate compounds on exposure solution pH and the effect of pH on worm

movement. Similarities and differences in toxicity are discussed.

Chapter 4 summarizes the conclusions that can be drawn from the presented

study, and suggests possible ideas for future experiments. 4

Reference

Ellman, G.L., Courtney, K.D., Andres, Jr. V., Featherstone, R.M. 1961. A new and

rapid colorimetric determination of acetylcholinesterase activity. Biochem.

Pharmacol. 7, 88-95.

5

CHAPTER 2

LITERATURE REVIEW

Organophosphorus Esters

Organophosphorus esters (organophosphates or OPs) are a broad and important class of pesticidal chemicals which serve as the active ingredient in many insecticides, acaricides, nematocides, helminthicides, herbicides, and fungicides (Chambers, 1992).

Virtually unknown until the late 1930s, the German chemist Gerhard Schrader was the first to explore OP synthesis and utility. Schrader’s work lead to the discovery of a number of insecticides as well as two extremely toxic OPs ( and ) that have since been employed as chemical warfare agents (Chambers, 1992). Primarily due to their high efficacy and low environmental persistence, organophosphates became the insecticide of choice in the 1970s (Pope, 1999). The utility and risk of OPs comes from their potential to poison the nervous system of invertebrate and vertebrate animals by inhibiting the essential enzymatic activity of acetylcholinesterase. This has resulted not only in the control of pestilent insect species, but also in the accidental poisoning of nontarget animal species, most notably humans. Movement in recent years to greater use of the more selective pyrethroid and carbamate insecticides has reduced the dependence on OP usage, but these chemicals still account for a large and important part of pest control efforts in the world today. 6

The term organophosphate is used to describe a group of chemicals in which there is a pentavalent phosphorus sharing a double bond with either oxygen or sulfur, and some combination of carbon containing substituents, traditionally denoted X, Y, and Z.

Though the X and Y substituents can display almost endless variety of alkyl or aryl identities, their attachment to the central phosphorus atom is typically through a direct bond with an oxygen, nitrogen or sulfur (Ecobichon, 1982). The Z designation is typically reserved for that substituent which is most easily replaced through chemical substitution. Enormous variation in the Z substituent is also possible, but in the case of organophosphate pesticides, this group is most commonly a halogen, aryl group, phosphate, cyanide, thiocyanate, carboxylate, phenoxy, or thiophenoxy group

(Ecobichon, 1982). The nomenclature associated with organophosphates (OPs) is complicated and historically inconsistent. An in-depth discussion of this nomenclature can be found in O’Brien (1960).

The physicochemical properties of a particular organophosphate ester are driven by the nature of its substituent groups. The vast array of possible substituent combinations creates an even greater variety of overall physicochemical characteristics.

The individual and combined qualities of these component groups are crucial in determining the rate of absorption, distribution, biotransformation and target site interactions which culminate in the potency and selectivity of each chemical (Ecobichon,

1982). Attempts have been made to predict how changing substituents and the resulting alterations in electronic, steric, and hydrophobic properties modify biological activity.

These efforts have met with some success, and are discussed by Hansch (1969).

7

Acute Toxicity

The acute toxicity of organophosphate compounds is classically attributed to

inhibition of the nervous system enzyme acetylcholinesterase (AChE). Under normal

circumstances, AChE serves to end the transmission of cholinergic nervous impulses by

degrading the neurotransmitter acetylcholine (ACh) at neural synapses and

neuromuscular junctions. The inactivation of ACh is accomplished via transesterification

whereby free choline is released, and the active serine of the enzyme is acetylated.

Subsequent hydrolysis rapidly restores the enzyme to the active state. In the presence of

an organophosphate, the active serine displaces the Z substituent and becomes

phosphorylated. Hydrolytic restoration of the phosphorylated enzyme to the active state

proceeds at a rates typically orders of magnitude slower than the acetylated case

effectively inactivating the enzyme. Instead of reactivation, the phosphorylated enzyme

can alternatively undergo a process termed aging. In aging, dealkylation occurs at the

phosphoester X or Y position stabilizing the phosphorylated enzyme to the point that

hydrolytic reactivation becomes virtually impossible (Ecobichon, 1982). Several factors

determine the degree of AChE recovery in the hours and days following inhibition.

Rapid and virtually complete recovery is characteristic with some exposures while

inhibition by other OPs results in the slow reestablishment of AChE activity through de

novo enzyme synthesis.

The degree to which AChE inhibition occurs in vivo depends upon factors

specific to the particular chemical and animal species in question. Variability in toxicity is frequently observed between taxonomic groups and among species within those groups

(Benke et al., 1974; Moss and Fahrney, 1978; Chambers and Carr, 1995). Hutson and 8

Millburn (1991) describe these differences as being rooted in the multifactorial nature of

OP absorption, bioavailability, metabolism and inhibitory potential. A number of

structural factors have been proposed by Wallace (1992) as modifiers of the pharmacodynamic efficiency of AChE inhibition. After comparing the hydrolytic rates

for several substrates by the AChE of rats, hens, and rainbow trout, Kemp and Wallace

(1990) concluded that steric occlusion and nucleophilic strength at the enzyme’s active

esteratic site play important roles in vulnerability to phosphorylation. Proximity of the

AChE active site to its choline-coordinating anionic region, and allosteric interaction

between the anionic and active sites may also be important to inhibitory dynamics within

and between species (Wallace, 1992).

Recent studies indicate that some OPs cause secondary effects which may modify the impact of AChE inhibition. Pope (1999) reviewed several studies reporting pre- and postsynaptic interactions of different OPs with both nicotinic and muscarinic acetylcholine receptors (nAChR and mAChR, respectively). Kahn et al. (2000) found acute exposures in rats not only caused increased binding of mAChRs and nAChRs, but also persistent stimulation of choline acetyltrasnferase, an enzyme involved in the biosynthesis of ACh. These effects combine into a complicated picture. The OP binding of postsynaptic nAChRs could enhance the effects of heightened ACh stimulation, but significant binding may only occur at OP concentrations well above what is necessary to cause AChE inhibition (Pope, 1999). Stimulation of choline acetyltransferase could increase ACh release, but presynaptic mAChRs regulate choline reabsorption which is the rate limiting step in ACh production. In contrast to these studies, Camara et al.

(1997) concluded that methamidophos exposure did not impact neurotransmitter release 9 or post-synaptic nAChRs. It seems, therefore, that the potential for OPs to affect ACh release and neurotransmitter receptors may be both chemical and concentration specific.

The biotransformation of organophosphate compounds is critical to their overall toxicity. There are numerous accounts of metabolism causing dramatic toxicity differences between chemicals and between species (Gaines, 1966; Hutson and Millburn,

1991; Chambers and Carr, 1995). In many cases, metabolism is required for OP bioactivation as well as detoxication. The most common groups requiring bioactivation are the phosphorothioates (e.g. parathion, Table 2.1) and phosphorodithioates (e.g. dimethoate, Table 2.1) which must undergo oxidative desulfuration to the corresponding oxon in order to exhibit any significant anticholinesterase behavior (Meyers and Mendel,

1952; Metcalf and March, 1953; Gaines et al., 1966). Activation is predominantly carried out in mammals by Cytrochrome P-450 isozymes in liver and other tissues

(Davison, 1955; Murphy and Dubois, 1957; Ecobichon, 2001). Chambers and Carr

(1995) found that induction of rat microsomes with phenobarbital increased the rate of chlorpyrifos desulfuration but decreased overall toxicity. Their conclusion was that microsomal induction shifted the activity ratio between two or more P450s competitively carrying out the desulfuration and dealkylation of chlorpyrifos. This case demonstrates the dynamic nature of the bioactivating and/or detoxifying pathways that work in series and in parallel to modify the concentration of different OP species in vivo (Kemp and

Wallace, 1990).

Organophosphate detoxication results from a large collection of enzymatic reactions. Phase I reactions include the activities of Cytochrome P450s, glutathione-S- transferases, and variety of phosphotases, esterases, and amidases (Dicowsky and 10

Morello, 1971; Ecobichon, 1982,). Degradation reactions can involve phosphoester bonds or side chain components (Matsumura, 1985). As with other xenobiotics, these reactions depend greatly upon the physicochemical nature of the substrate with some pathways displaying considerable specificity. Serine esterases and proteases other than

AChE are believed to play a protective role in low to moderate level OP exposures

(Richardson, 1995). As with AChE, phosphorylation of these enzymes is usually irreversible raising concerns over the potential effects of their inactivation. Phase II conjugation reactions have also been reported for OP compounds in the literature, but appear to be much less common than phase I reactions. Hutson et al. (1967) identified glucuronide and glycine conjugates in the breakdown of chlorfenvinphos in rats and dogs. Trichlorfon and some related compounds have also been reported to be susceptible to glucuronidation (Yang, 1976). While accounts can be found in the literature, they seem overshadowed by phase I reactions. A more comprehensive treatment of OP biotransformation is given by Yang (1976).

Clinical Manifestations and Treatments

The physiologic effects of acute OP intoxication are predominantly due to the

buildup of ACh in the central nervous system and at the nicotinic and muscarinic

synapses of the peripheral nervous system and neuromuscular junctions (Marrs, 1993).

The clinical symptoms are numerous and vary to some degree in occurrence and severity

between individuals and type of exposure. Classic symptoms of acute OP poisoning

include miosis, headache, confusion, tremors, convulsions, rhinorhea, lachrymation,

diarrhea, bradycardia, tachycardia, muscular weakness, and fasiculations (Marrs, 1993). 11

Death is usually the result of asphyxiation due to some combination of hypotension,

bronchoconstriction, diaphragmatic paralysis, and depression of the respiratory centers in

the brain (Ecobichon, 1982). Though they seem to bear little correlation to actual

toxicity, depression of erythrocyte AChE, plasma AChE, and butyrylcholinesterase

(BChE) can be measured as confirmation of exposure to an anticholinesterase agent

(Richardson, 1995). Treatment of acute poisoning can entail the treatment of symptoms

as well as attempts to reverse the inhibition of AChE. The muscarinic acetylcholine

receptor blocking agent atropine is often employed to mute the effects of ACh buildup,

and artificial respiration is employed when needed. Attempts can also be made to

chemically promote the dissociation of the phosphodiester-enzyme complex by

administration of oximes such as pralidoxime (Marrs, 1993).

Organophosphates are also reported to have acute effects not clearly linked to

AChE inhibition. Acute cognitive and sensory effects have been described, but it has been unclear whether these symptoms resulted from the direct action of the OPs, or secondarily from anoxia associated with intoxication (Marrs, 1993). Abdel-Rahman et al. (2002) report acute sarin exposures at the LD50 level in rats increased permeability of the blood-brain barrier and caused extensive damage to the cerebral cortex, hippocampus, and cerebellum in addition to inhibiting brain AChE and plasma BChE. Exposure at the

LD50 level also caused degeneration in the motor and sematosensory cortex, and doses down to 0.5 (LD50) caused death of cerebellar Purkinje fibers. With respect to non- nervous system organs, postmortem histological evidence has drawn possible links between acute OP intoxication and cardoiotoxicity (Marrs, 1993). There have also been anecdotal, and somewhat circumstantial, reports of OP disruption of the immune system 12

(Richardson, 1995). Recent experiments lend some mechanistic credibility to these

observations. Li et al. (2002) describe in vitro tests where dichlorvos exposure resulted

in the dose dependent inhibition of natural killer cells, cytotoxic T lymphocytes, and

lymphokine-activated killer cells. These cells function in directed cell death in part by

the release of cytolytic granzymes containing serine proteases. Tests by Li et al. (2002)

also showed the dose dependent inhibition of granzymes offering an explanation for

decreased immunocell function.

A so called intermediate syndrome has been described as resulting from some cases of acute OP intoxication. The common symptoms of this condition include the onset of proximal muscle weakness and paralysis 1-4 days following intoxication (Mars,

1993; Ray, 1998). The syndrome commonly culminates in acute respiratory failure, but is unresponsive to the usual treatments (atropine and oximes) of OP poisoning (Marrs,

1993).

Organophosphate Induced Delayed Neuropathy (OPIDN) is another syndrome

associated with acute poisoning by some OP compounds. OPIDN is characterized by

retrograde axonal degeneration of the long, large diameter nerve fibers approximately 10-

14 days following an episode of acute distress. Secondary to the axonal degeneration,

demyelination, Schwann cell proliferation, and macrophage accumulation usually occur

(Barrett et al., 1985). The result is a long term, usually permanent paralysis of the legs

and less commonly the arms (Richardson, 1995). The onset of OPIDN has been linked to

the phosphorylation and aging of NTE (neurotoxic esterase or neuropathy target

esterase), but this relationship is still poorly understood (Marrs, 1993). NTE is an

esterase protein primarily of neuronal origin with decreasing concentrations in the brain, 13 spinal cord, peripheral nerves, and some other tissues (Richardson, 1995). NTE is less vulnerable to phosphorylation than AChE explaining the development of neuropathy almost exclusively after exposures sufficient to produce acute intoxication (Marrs, 1993).

A thorough treatment of OPIDN and the compounds and processes associated with its onset is given by Barrett et al. (1985).

Chronic Effects of Organophosphate Chemicals

There is a general lack of clinical evidence indicating persistent effects from chronic or subchronic OP exposures. The absence of clear long term effects may in part be due to the difficulty of drawing conclusions from uncontrolled and poorly characterized exposures, but even long term clinical usage of select OPs has failed to raise significant cause for concern (Ray, 1998). A few studies, however, have shown persistent changes in brain electrical activity. Korsak and Sato (1977) performed EEGs and a battery of neuropsychological tests on a group of workers had received chronic and clinically nontoxic exposures to organophosphate compounds. Their findings indicated asymptomatic EEGs in the exposed group, distinguishing the left frontal lobe as the brain region most highly impacted. Duffy et al. (1979) also conducted EEGs on a group which had been occupationally exposed one or more times at least one year previous. Using computer analysis of the EEGs, they were able to distinguish increased beta-wave activity suggestive of a long term impact on brain ACh receptors. In both studies, it seems that any effects were too subtle to produce gross clinical signs. In a well defined subchronic study, Sheets et al. (1997) found no additional effects on brain ChE after 13 weeks of feeding exposures compared to levels after 1 and 2 years. No additional 14

behavioral effect was found after the first 4 weeks. This evidence supports the idea that

OPs have little long term effect.

Sublethal Organophosphate Exposure and Mammalian Behavior

Comparisons between studies characterizing the behavioral response of mammals

to sublethal OP exposures can be difficult not only because they differ in chemical and

animal species, but also because of a general lack of standardization in behaviors

measured, measurement unit, route of administration, and time course for observation.

Despite these differences, a picture of overall behavioral depression emerges from the

literature. One of the more consistent behavioral changes observed in response to acute

sublethal exposures is decreased locomotor activity (Lynch et al., 1986; Padilla et al.,

1992; Llorens et al., 1993; Dell‘Omo and Shore, 1996; Dell’Omo et al., 1997; Sheets et

al., 1997) . Padilla et al. (1992) saw decreased movement along with miosis, diarrhea,

and tremors in response to intraperitoneal (IP) injections of paraoxon. Two studies of

behavior in the field following a single dimethoate administration to wood mice

(Dell’Omo and Shore, 1996) and shrews (Dell’Omo et al., 1997) observed decreased

locomotor activity along with grooming, sniffing, and rearing behaviors for a six hour

period followed by a total behavioral recovery. In both cases there was a temporal

correlation between the behavioral changes and decreased ChE activity measurements

taken in the laboratory. They also found that behavior returned to normal six hours after

administration which is significantly faster than the recovery of AChE activity. Subacute

and subchronic studies have produced similar findings to acute experiments. Llorens et

al. (1993) observed the coincidence of classic clinical signs and decreased locomotion in 15

rats following repeated exposures to disulfoton. While the clinical signs quickly

developed a tolerance response, movement behavior remained depressed over the 30 day

exposure period. These results were in keeping with those of Padilla et al. (1992).

Sheets et al. (1997) performed a study where six OPs were incorporated into rat diets for

13 weeks. Increasing behavioral and enzymatic depression was observed for 4 weeks,

but no further cumulative response was detected.

Due to the high risk of subjectivity in measuring behavioral changes, several

researchers have used standardized observation procedures and automated measurement

systems in determining behavioral response. Several researchers have used the

Functional Observational Battery (FOB) of Moser (1988, 1989) to characterize certain

behaviors in different settings or following certain stimuli (Padilla et al., 1992; Sheets et

al., 1997). Sheets et al. (1997) found the combination of clinical observations with the

FOB to be effective at detecting affects on the central nervous system as well as in the

periphery. Rodent movement behavior has been effectively quantified using a figure- eight maze equipped with infrared motion sensors automated to keep track of beam interruptions (Padilla, 1992; Llorens et al., 1993; Sheets et al., 1997). The FOB and automated measurement were able to detect behavior changes correlating to brain ChE depression of approximately 20% leading to the judgment that these techniques are more

sensitive than simple observation (Padilla et al., 1992; Sheets et al., 1997). 16

Caenorhabditis elegans

The nematode Caenorhabditis elegans is a free living bactiverous soil nematode

of the order Rhabditida. In the mid to late 1960s, it was isolated and characterized by

Sidney Brenner for use in the laboratory research. Since that time, it has become one of

the most studied animal species in history being used as a biological model in genetics,

molecular biology, developmental biology, and neurobiology.

Caenorhabditis elegans has many traits that make it suitable for use as a model

organism. The adult is approximately 1mm in length allowing large numbers of

individuals to be kept in a limited space. C. elegans has a life span of approximately

three weeks and a life cycle from egg to reproductive adult of 3-4 days at 200C. It easily

lives in aqueous culture or on agar filled petri dishes with Escherichia coli as a food

source (Wood, 1988). The adult is typically a self fertilizing hermaphrodite with a

lifetime fecundity of approximately 300 eggs per individual. Males infrequently develop and can mate with hermaphrodites allowing for some gene flow within a population

(Wood, 1988). Age synchronous populations are easily obtained through the collection and culturing of eggs. C. elegans is highly tolerant to ranges of temperature (12-250C),

pH (3.2-11.8), salinity (up to 15.46 g/L NaCl), and water hardness (0.236-0.246 g/L

NaHCO3) (Kahanna et al, 1997). In the absence of food, C. elegans can exist for several

months as dauerlarvae, a facultative dormant life stage from which it can reemerge with

the restoration of sufficient resources.

17

Pertinent aspects of C. elegans anatomy, physiology, and behavior

The basic anatomy of C. elegans consists of two concentric tubes. The inner tube

is made up of the intestine and gonad while the outer tube includes the cuticle,

hypodermis, musculature, and nervous tissues (Wood, 1988). The two tubes are

separated by the potential space of the pseudocoelom. The adult hermaphrodite has 810

somatic cells. Each cell has been characterized from differentiation to death by Sulston et al. (1983) and found to show invariant patterns of cell development and migration

(Schnabel and Priess, 1997).

C. elegans is one of the simplest organisms with a centralized nervous system. As in higher organisms, the system’s basic functional unit is the neuron. The adult hermaphrodite has 302 neurons along with 56 glial and associated support cells (Chalfie and White, 1988). The entire system has been mapped by White et al. (1986) through the reconstruction of serial dissection electron micrographs. As a whole, the nervous system includes approximately 2000 neuromuscular junctions, 5000 interneural chemical synapses and 600 gap junctions (Thomas, 1994). Most of the neurons are bundled into secondary structural units. The most important of these are the nerve ring, and dorsal and ventral nerve cords (Rand and Nonet, 1997).

C. elegans neural morphology is quite simple. Neurons typically possess one or two unbranched processes that make connections en passant with little specialization of the process terminals (Thomas, 1994). Presynaptic terminals show little structural distinction, but characteristically display areas concentrated with neurotransmitter filled vesicles. Post synaptic terminals also show little structural specialization. The type of 18

structural complimentarity displayed by terminal invagination in mammalian neuromuscular junctions is completely absent in C. elegans (Rand and Nonet, 1997).

It has been shown that C. elegans neurons do not exhibit classical action

potentials, and do not possess the sodium channels or myelin sheaths needed for rapid

electrochemical conduction over long distances (Goodman et al., 1998). It has been

hypothesized that C. elegans lost its sodium channels over time because action potentials

were not required for the short distances of transmission (Davis and Stretton, 1989).

Instead, membrane currents are conducted using a wide variety of potassium and calcium

channels (Wei et al., 1996). These membrane currents mediate tonic neurotransmitter

release capable of graded increases or decreases in response to shifts in electric potential

(Davis and Stretton, 1989). This type of constant electrochemical signal modification

may support less drastic but more sensitive responses to sensory inputs.

C. elegans uses wide array of classical neurotransmitters to chemically span the

synaptic cleft. These neurotransmitters are produced and handled in a manner very

similar to vertebrates (Bargmann, 1998). Enzymatic biosynthesis of neurotransmitters

from common metabolites is followed by microtubule mediated vesicular transport to

neuronal terminals and calcium induced exocytosis (Rand and Nonet, 1997).

Neurotransmitters that have been described in C. elegans include acetylcholine (Lewis et

al., 1980; Hosono et al., 1987), GABA (McIntire, 1993), serotonin (Horvitz et al., 1982),

dopamine (Sulston et al., 1975), glutamate (Li et al., 1997), and several neurogenic

peptides (Rand and Nonet, 1997).

Acetylcholine (ACh) is the primary excitatory neurotransmitter utilized in C.

elegans motor function. The genes controlling its biosynthesis by choline 19

acetyltransferase (cha-1) and vesicular transport (unc-17), show significant similarities

between nematodes and higher vertebrates (Hosono et al., 1987; Erickson et al., 1994;

Bargmann, 1998). Proteins involved in ACh exocytosis including syntaxin, synaptobrevin, and synaptotagmin also show homologies to mammals (Bargmann, 1998).

The work of Lewis et al. (1980) demonstrated the presence of nicotinic acetylcholine

receptors (nAChRs) in C. elegans. Since that time genetic analysis has predicted the

presence of approximately 40 different subunits belonging to the nicotinic class

(Bargmann, 1998). Muscarinic ACh receptors (mAChRs) have also been identified in C.

elegans. These receptors are similar but not identical to vertebrate types in their

pharmacological characteristics (Culotti and Klein, 1983).

As in higher animals, ACh in C. elegans is degraded between neurons or in the

neuromuscular junction by acetylcholinesterase (AChE). Both C. elegans and vertebrates

have more than one form of AChE, but whereas vertebrates have one AChE gene whose

product undergoes alternative splicing posttranscriptionally, C. elegans possesses four

distinct AChE genes (Combes et al., 2001). The ace-1 and ace-2 genes encode the two

major forms of AChE in C. elegans which account for approximately 95% of the overall

AChE activity. The ace-1 gene has been traced to chromosome X (Arpagaus et al.,

1994). A study linking green fluorescent protein (GFP) production to the ace-1 promoter

localized its expression almost exclusively to the worm’s musculature (Combes et al.,

2001). The ace-2 gene is part of chromosome I, and is expressed in the sensory and

motor neurons of the ganglia and nerve cords (Combes et al., 2001). The ace-3 gene was

discovered by Johnson et al. (1988), and encodes a minor form of AChE. A fourth ace

gene was found by Grauso et al. (1998). Its presence has been confirmed through RT- 20

PCR analysis of transcription products, but it seems to be translated into little, if any, active product in vivo (Combes, 2001).

Studies using C. elegans strains exhibiting null mutations in the ace-1, ace-2, and ace-3 genes have revealed different aspects of their relationships to each other. A high degree of functional overlap has been described between ACE-1 and ACE-2 enzymes.

Inactivation of either gene alone results in very little phenotypic change while the ace-1, ace-2 double mutant displays highly impaired behavior (Culotti et al., 1981; Johnson et al., 1981; Johnson et al, 1988). This functional redundancy is interesting given the

differing nature and distribution of ACE-1 and ACE-2 enzymes. The level of

incoordination displayed by ace-1, ace-2 double mutants compared to either single

mutant reveals the subordinate role ACE-3 plays compared to the two major forms. The

ace-1, ace-2, ace-3 triple mutant is not viable (Rand and Nonet, 1997). This demonstrates that some AChE activity is essential for C. elegans survival, and that the activity of the ace-4 gene product is negligible.

The protein products of the different AChE genes have also been studied.

Johnson and Russell (1983) described five protein forms of various sizes which they

divided into two classes (traditionally denoted as A and B) according to physicochemical

characteristics. Further study has shown the "class A" proteins are components of the

ACE-1 form of AChE and products of the ace-1 gene. Likewise, the class B proteins

have been shown to be components of the ACE-2 enzyme encoded by the ace-2 gene

(Combes, 2001). Johnson and Russell (1983) found the ACh binding affinities for ACE-

1 and ACE-2 to be similar to vertebrate AChE, but both were less substrate specific.

They also demonstrated ACE-1 and ACE-2 to be vulnerable to a variety of cholinesterase 21

inhibitors including two organophosphate pesticides. ACE-3 accounts for approximately

5% of C. elegans’ AChE activity, and displays different pharmacological characteristics

than ACE-1 or ACE-2. Kolson and Russell (1985a, 1985b) found ACE-3 to have a much

greater affinity for ACh than the two major forms, and also to be considerably less vulnerable to inhibition by anticholinesterase agents. ACE-1, ACE-2, and ACE-3 account for virtually all AChE activity in C. elegans. Difficulties in isolating the ACE-4 protein have lead to the speculation that ace-4 is essentially a non-functional gene

(Combes et al., 2000; Combes et al., 2001).

Early investigations caused researchers to think there would be little commonality

between C. elegans and vertebrate AChE (Johnson and Russell, 1983). Since that time

an increasing number of similarities have emerged between the different forms. Using

genetic analysis, Arpagaus et al. (1994) discovered 42% identity between ACE-1 and

mammalian AChEs. The highest degree of homology is between ACE-1 and the C-

terminal region of vertebrate T-peptide. Sedimentation studies by Combes et al. (2001)

found that the major form of ACE-1 in vitro was a tetramer likely bearing some

association with a smaller monomer. This was consistent with previous investigations by

Arpagaus et al. (1992) which showed the product of ace-1 in the nematode Stinernema

carpocapsae (a rhabditidae related to C. elegans) was a catalytic subunit which

oligomerized into an amphiphilic tetramer. This structural arrangement bears

resemblance to the T-peptide G4 form of AChE found in the vertebrate brain (Combes et

al., 2001). Evidence from studies using ace-1 mutants indicate that the product of ace-2

is a hydrophilic monomer which can dimerize and associate with a glycolipid moiety

(Combes et al., 2001). This quaternary structure is similar to dimeric H-peptides of 22

vertebrates which are anchored to membrane surfaces by a glycolipid unit added to the

dimer posttranslationally (Combes et al., 2001). The differential expression of ace genes by cell type is also interestingly similar to the AChE arrangement in vertebrates.

Whether these and other similarities are due to homology or analogy remains to be determined.

Despite its simplicity, C. elegans’ neuromuscular system supports a variety of behaviors. More than 250 genes have been identified with some role in regulating behavior (Thomas, 1994). Muscular contractions resulting in sine wave type motion allow for forward and backward movement. The worms feed, defecate, lay eggs, and in the case of males exhibit a fairly complex mating behavior (Wood, 1988). C. elegans also displays complex chemosensory behavior exhibiting attraction and repulsion responses to a well over 100 chemical stimuli (Bargmann, 1993; Thomas, 1994).

Reviews of these behaviors and the research concerning them can be found in Riddle et al. (1997).

Metabolism

Most research on how C. elegans deals with xenobiotics has been devoted to the

detoxification of heavy metals. It is only in recent years that pathways for the

metabolism of organic contaminants have received attention. Unlike higher organisms,

C. elegans does not have an organ primarily devoted to the biotransformation of

xenobiotics. Much of this activity is concentrated in the intestine (McGhee, 1987;

Menzel et al., 2001). Only reports of phase I type biotransformations in C. elegans have

been published, but there has been at least one account of gene induction for phase II 23 enzymes like UDP-glucuronosyltransferases (Menzel et al., 2001). Approximately 80 cytochrome P450 genes have been identified in C. elegans (Nelson, 1999). Although little is known about C. elegans’ CYP450s, it is believed that many of them developed in order to cope with the variety of conditions nematodes face (Gotoh, 1998). A large portion of these enzymes belong to a pure group of nematode P450 isozymes designated as the C. elegans clan. This clan appears to share a common ancestor with the 2 clan of

P450s which include many mammalian enzymes known to be active in drug metabolism

(Nelson, 1999). Using RT-PCR to characterize RNA expression, Menzel et al. (2001) looked for evidence of P450 induction following exposures to known xenobiotic inducers. They found four subfamilies of C. elegans CYP450s to be particularly responsive to such compounds as beta-napthoflavone, phenobarbital, atrazine, clofibrate, lansoprazole, and fluoranthene. Other metabolic enzymes likely to be important to OP toxicity have also been identified. McGhee et al. (1987) characterized a nonspecific esterase from the gut of C. elegans that showed little affinity for ACh, but was stoichiometrically inhibited by several organophosphates. Several forms of glutathione-S- transferase have also been identified (van Rossum et al., 2001).

The use of C. elegans as a Model Organism

In thinking about the directions he should pursue in his research, Sydney Brenner in the early 1960s had the idea of using genetic techniques in a simple model organism to study development and the nervous system (Brenner, 1988). In his initial description of the genetics of C. elegans, Brenner (1974) cites the potential for determining the structure of its entire nervous system as a primary reason for choosing this nematode as his model 24 organism. Since that time, all cell lineages of C. elegans have been traced (Sulston et al.,

1983), its nervous system has been mapped (White et al., 1986), and its genome has been sequenced (C. elegans Sequencing Consortium, 1998). These important accomplishments have combined in a unique way to allow for Brenner’s initial ambitions to become a reality. In the study of developmental and neurological questions, these databases can be used in conjunction with mutant and transgenic strains of C. elegans to decipher what structures are affected and what processes are altered.

C. elegans has been used as a model for ageing for more than twenty years

(Zuckerman and Himmelhoch, 1980). The transgenic capabilities of C. elegans have also made it useful as a model for human degenerative neurological diseases. Invertebrate models allow researchers to use a variety of powerful genetic tools in the analysis of transgenic lines expressing disease associated proteins over a short lifespan (Link et al.,

2001). Human beta-amyloid peptide, a peptide associated with the onset of Alzheimer’s disease, has been expressed in C. elegans by Link et al. (2001). Worms expressing the protein displayed a progressive onset of paralysis. A gene encoding multiple polyglutamine repeats of the sort associated with Huntington’s disease have also been expressed in C. elegans sensory neurons (Faber et al., 1999). Cells expressing the longest of such repeats were found to exhibit an age dependent decrease in function. The deleterious impacts of human beta-amyloid peptide and polyglutamine repeats in nematode cells strongly imply that the pathology of these diseases lie to a great extent in the basic cellular processes conserved between vertebrates and invertebrates (Link,

2001). 25

C. elegans has also found use as a model in the biological response to anesthetic

agents. A number of researchers have used behavioral endpoints to characterize the

response of C. elegans to a host of volatile anesthetic agents (Morgan and Cascorbi,

1985; Crowder et al., 1996; Kayser et al., 1998). Work with anesthetics in higher animals

has long established a correlation between anesthetic potency and lipophilicity called the

Meyer-Overton correlation. Early work by Morgan and Cascorbi (1985), and Morgan et

al. (1988) confirmed that the wild type N2 strain of C. elegans adheres to the Meyer-

Overton relationship, but also described mutant strains deviating from this response.

Subsequent work with these mutant strains demonstrated that contrary to the traditional paradigm, the anesthetic response was controlled by several pathways at multiple sites of action (Morgan and Sedensky, 1994, 1995; Kayser et al., 1998). Working with a more common anesthetic, Anton et al. (1992) found ethanol to have a general effect on worm movement very similar to that observed in higher animals including mammals. Morgan and Sedensky (1995) were able to replicate these results. Critical evaluations of C. elegans by Anton et al., (1992) and Morgan and Sedensky (1995) lead to the conclusion that despite certain limitations, C. elegans’ strict adherence to important aspects of the basic anesthetic response makes it a useful as a mammalian model.

Efforts have also been made to evaluate C. elegans as a model for mammalian

xenobiotic toxicity. Williams and Dusenbery (1988) found that acute lethality values in

C. elegans exposed to eight heavy metal salts predicted overall mammalian acute

lethality as consistently as tests in rats and mice individually. Williams and Anderson

(2000) conducted blind tests on five gadolinium-based compounds used in magnetic

resonance imaging utilizing movement behavior as the measured endpoint. These tests 26

correctly ordered the toxicity for four out of the five compounds compared to median lethal doses for the same compounds in mice. Anderson et al., (submitted for publication) found five compounds across three chemical classes used in behavioral

toxicity tests to have an identical order of toxicity in C. elegans as in mammals.

Transgenic and wild type strains of C. elegans have been used in environmental

toxicity testing. Most common has been its use in assessing the toxicity of heavy metal

contamination in soil (Black and Williams, 2001; Boyd and Williams, 2003) and aquatic

environments (Hitchcock et al., 1997). Williams and Dusenbery (1990) found salts of

lead (Pb) and aluminum (Al) to display toxicity dynamics consistent with neurotoxicity.

Nematode behavior has been shown to be much more sensitive than lethality in the

characterization of toxins (Dhawan, et al., 2000; Anderson et al., 2001). The

development of computer automated tracking systems has made possible the accurate

characterization of nematode movement as a toxic endpoint (Dusenbery, 1996; Dhawan

et al., 1999; Boyd et al., 2000). With the idea that movement behavior may be

preferentially sensitive to neurotoxicants, several researchers have used locomotion in the

examination of chemicals known to be neurotoxic. Anderson et al. (2001) demonstrated

that movement in C. elegans was more sensitive to Pb than to Cu over a 4-hr exposure

period while this difference in sensitivity was masked at exposures lasting 24-hr. Recent

work by Anderson et al. (submitted) indicated that this preferential sensitivity may be

indicative of mechanistic differences between the known neurotoxin Pb and Cu which is

not believed to be neurotoxic. They also conclude that the combination of 4-hr exposures

with locomotor measurements is well suited to the detection of neurotoxicity. 27

The research reviewed here leaves outstanding the question of how similar the

dynamics of neurotoxicity are between C. elegans and mammals. Answering this question will require a more rigorous evaluation of C. elegans’ response to neurotoxic chemicals than has been undertaken to this point. The purpose of this thesis work was to begin to address this question of comparability by characterizing the toxic response of C.

elegans to a class of neurotoxins and making comparisons to the corresponding response

displayed by mammals. 28

References

Abdel-Rahman, A., Shetty, A.K., and Abou-Donia, M.B. 2002. Acute exposure to sarin

increases blood brain barrier permeability and induces neuropathological changes

in the rat brain: Dose-response relationships. Neuroscience 113(3), 721-741.

Anderson, G.L., Boyd, W.A., and Williams, P.L. 2001. Assessment of sublethal

endpoints for toxicity testing with the nematode Caenorhabditis elegans. Environ.

Toxicol. Chem. 20, 833-838.

Anton, A.H., Berk, A.I., Nicholls, C.H. 1992. The “Anesthetic” effect of alcohols and

alkanes in Caenorhabditis elegans (C.e.). Res. Commun. Chem. Path. Pharmacol.

78(1), 69-83.

Arpagaus, M., Richier, P., Berge, J., and Toutant, J.-P. 1992. Acetylcholinesterases of

the nematode Steinernema carpocapsae. Characterization of two types of

amphiphilic forms differing in their mode of membrane association. Eur. J.

Biochem. 207, 1101-1108.

Arpagus, M., Fedon, Y., Cousin, X., Chatonnet, A., Berge, J., Fournier, D., and Toutant,

J. 1994. cDNA sequence, gene structure, and in vitro expression of ace-1, the gene

encoding acetylcholinesterase of class A in the nematode Caenorhabditis elegans.

J. Biol. Chem. 269(13), 9957-9965.

Bargmann, C.I. 1998. Neurobiology of the Caenorhabditis elegans genome. Science

282, 2028-2033.

Bargmann, C.I., Hartwieg, E., and Horvitz, H.R. 1993. Odorant-selective genes and

neurons mediate olfaction in C. elegans. Cell 74(3), 551-527. 29

Barrett, D.S., and Oehme, F.W. 1985. A review of organophosphorus ester-induced

delayed neurotoxicity. Vet. Hum. Toxicol. 27, 22-37.

Benke, G.M., Cheever, K.L., Mirer, F.E., and Murphy, S.D. 1974. Comparative toxicity,

anticholinesterase action and metabolism of methyl parathion and parathion in

sunfish and mice. Toxicol. App. Pharmacol. 28, 97-109.

Black, M.C. and Williams, P.L. 2001. Preliminary assessment of metal toxicity in the

middle Tisza River (Hungary) flood plain. J. Soil Sed. 1, 213-216.

Boyd, W.A. and Williams, P.L. 2003. Availability of metals to the nematode

Caenorhabditis elegans: Toxicity based on total concentrations in soil and extracted

fractions. Environ. Toxicol. Chem. 22, 1100-1106.

Boyd, W.A., Anderson, G.L., Dusenbery, D., Williams, P.L. 2000. Computer tracking

method for assessing behavioral changes in the nematode Caenorhabditis elegans.

In: Price, F.T., Brix, K.V., and Lane, N.K., (Eds.), Environmental Toxicology and

Risk Assessment, Vol. 9. STP 1381. American Society for Testing and Materials,

Philadelphia, P.A. 225-238.

Brenner, S. 1974. The genetics of Caenorhabditis elegans. Genetics 77, 71-94.

Brenner, S. 1988. Foreward. In: Wood, W.B. (Ed.) The Nematode Caenorhabditis

elegans. Cold Spring Harbor Laboratory, Plainview, NY, USA, pp. ix-xiii.

C. elegans Sequencing Consortium. 1998. Genome sequence of the nematode C.

elegans: a platform for investigating biology. Science 282, 2012-2018.

30

Camara, A.L., Braga, M.F., Rocha, E.S., Santos, M.D., Cortes, W.S., Cintra, W.M.,

Aracava, Y., Maelicke, A., and Alburquerque, E.X. 1997. Methamidophos: an

anticholinesterase without significant effects on postsynaptic receptors or

transmitter release. Neurotoxicology 18(2), 589-602.

Chalfie, M., and White, J. 1988. The nervous system. In: Wood, W.B. (Ed.) The

Nematode Caenorhabiditis elegans. Cold Spring Harbor Laboratory, Plainview,

NY, USA, pp. 1-16.

Chambers, H.W. 1992. Organophosphorus compounds: an overview. In: Chambers, J.,

and Levi, P. (Eds.), Organophosphates Chemistry, Fate, and Effects. Academic

Press, Inc., San Diego, CA, USA, pp. 3-17.

Chambers, J.E., and Carr, R.L. 1995. Biochemical mechanisms contributing to species

differences in insecticidal toxicity. Toxicology 105, 291-304.

Combes, D., Fedon, Y., and Toutant, J. 2001. Acetylcholinesterase genes in the

nematode Caenorhabditis elegans. Int. Rev. Cytol. 209, 207-239.

Combes, D., Fedon, Y., Grauso, M., Toutant, J.-P., and Arpagaus, M. 2000. Four genes

encode acetylcholinesterases in the nematodes Caenorhabditis elegans and

Caenorhabditis briggsae. cDNA sequences, genomic structures, mutations, and in

vivo expression. J. Mol. Biol. 300, 727-742.

Crowder, C.M., Shebester, L.D., and Schedl, T. 1996. Behavioral effects of volatile

anesthetics in Caenorhabditis elegans. Anesthesiology 85, 901-912.

Culotti, J.G. and Klein, W.L. 1983. Occurrence of muscarinic acetylcholine receptors in

wild type and cholinergic mutants of Caenorhabditis elegans. J. Neurosci. 3(2),

359-368. 31

Culotti, J.G., Von Ehrenstein, G., Culotti, M.R., and Russell, R.L. 1981. A second class

of acetylcholinesterase-deficient mutants of the nematode Caenorhabditis elegans.

Genetics 97, 281-305.

Davis, E.D., and Stretton, A.D.W. 1989. Signaling properties of ascaris motorneurons:

Graded active responses, graded synaptic transmission, and tonic transmitter

release. J. Neurosci. 9(2), 415-425.

Davison, A.N. 1955. The conversion of (OMPA) and parathion to inhibitors of

cholinesterase by mammalian liver. Biochem. J. 61, 203-209.

Dell’Omo, G., and Shore, R.F. 1996. Behavioral and physiological effects of acute

sublethal exposure to dimethoate on wood mice, Apodemus sylvaticus [I-laboratory

studies]. Arch. Environ. Contam. Toxicol. 31, 91-97.

Dell’Omo, G., Breyenton, R., and Shore, R.F. 1997. Effects of exposure to an

organophosphate pesticide on behavior and acetylcholinesterase activity in the

common shrew, Sorex araneus. Environ. Toxicol. Chem. 16(2), 272-276.

Dhawan, R., Dusenbery, D.B., and Williams, P.L. 1999. Comparison of lethality,

reproduction and behavior as toxicological endpoints in the nematode

Caenorhabditis elegans. J. Toxicol. Environ. Health, Part A 58, 451-462.

Dhawan, R., Dusenbery, D.B., and Williams, P.L. 2000. A comparison of metal-induced

lethality and behavioral responses in the nematode Caenorhabditis elegans.

Environ. Toxicol. Chem. 19, 3061-3067.

Dicowsky, L., and Morello, A. 1971. Glutathione-dependent degradation of 2,2

dichlorovinyl dimethyl phosphate (DDVP) by the rat. Life Sci. Part II 10, 1031-

1037. 32

Duffy, F.H., Burchfiel, J.L., Bartels, P.H., Gaon, M., and Sim, V.M. 1979. Long-term

effects of an organophosphate upon the human electroencephalogram. Toxicol.

App. Pharmacol. 47, 161-176.

Dusenbery, D. 1996. NIH Image, version 1.59 computer tracking program modified by

Dr. David Dusenbery, School of Biology, Georgia Institute of Technology, Atlanta,

GA.

Ecobichon, D.J. 1982. Organophosphorus ester insecticides. In: Joy, R.M. and

Ecobichon, D.J. Pesticides and Neurological Diseases. CRC Press, Boca Raton,

FL, USA, pp. 151-203.

Ecobichon, D.J. 2001. Toxic effects of pesticides. In: Klassen, C.D. (Ed.) Casarett and

Doull’s Toxicology: The Basic Science of Poisons McGraw-Hill, New York, pp.

763-810.

Erickson, J.D., Varoqui, H., Schafer, M.K., Modi, W., Diebler, M., Weihe, E., Rand, J.,

Eiden, L.E., Bonner, T.I., and Usdin, T.B. 1994. Functional identification of a

vesicular acetylcholine transporter and its expression from a “cholinergic” gene

locus. J. Biol. Chem. 269(35), 21929-21932.

Faber, P.W., Alter, J.R., MacDonald, M.E., and Hart, A.C. 1999. Polyglutamine-

mediated dysfunction and apoptotic death of a Caenorhabditis elegans sensory

neuron. Proc. Natl. Acad. Sci. USA 96, 179-184.

Gaines, T.B., Hayes, W.J., Linder, R.E. 1966. Liver metabolism of anticholinesterase

compounds in live rats: relation to toxicity. Nature 209, 88-89. 33

Goodman, M.B., Hall, D.H., Avery, L., and Lockery, S.R. 1998. Active currents

regulate sensitivity and dynamic range in C. elegans neurons. Neuron 20(4), 763-

772.

Gotoh, O. 1998. Divergent structures of Caenorhabditis elegans cytochrome P450

genes suggest the frequent loss and gain of introns during the evolution of

nematodes. Mol. Biol. Evol. 15(11), 1447-1459.

Grauso, M., Culetto, E., Combes, D., Fedon, Y., Toutant, J.-P., and Arpagaus, M. 1998.

Existence of four acetylcholinesterase genes in the nematodes Caenorhabditis

elegans and C. briggsae. FEBS Lett. 424, 279-284.

Hansch, C. 1969. A quantitative approach to biochemical structure-activity

relationships. Accounts Chem. Res. 2, 232-239.

Hitchcock, D.R., Black, M.C., and Williams, P.L. 1997. Investigations into using the

nematode Caenorhabditis elegans for municipal and industrial wastewater toxicity

testing. Arch. Environ. Contam. Toxicol. 33, 252-260.

Horvitz, H.R., Chalfie, M., Trent, C., Sulston, J.E., and Evans, P.D. 1982. Serotonin and

octopamine in the nematode Caenorhabditis elegans. Science 216, 1012-1014.

Hosono, R., Sassa, T., and Kuno, S. 1987. Mutations affecting acetylcholine levels in

the nematode Caenorhabditis elegans. J. Neurochem. 49, 1820-1823.

Hutson, D.H. and Millburn, P. 1991. Enzyme-mediated selective toxicity of an

organophosphate and pyrethroid: some examples from a range of animals.

Biochem. Soc. Trans. 19, 737-740. 34

Hutson, D.H., Akintowa, D.A., and Hathway, D.E. 1967. The metabolism of 2-chloro-1-

(2’, 4’-dichlorophenyl) vinyl diethylphosphate in the dog and rat. Biochem. J. 102,

133-142.

Johnson, C.D., and Russell, R.L. 1983. Multiple molecular forms of acetylcholinesterase

the nematode Caenorhabditis elegans. J. Neurochem. 41, 30-46.

Johnson, C.D., Duckett, J.G., Culotti, J.G., Herman, R.K., Meneely, P.M., and Russell,

R.L. 1981. An acetylcholinesterase-deficient mutant of the nematode

Caenorhabditis elegans. Genetics 97, 261-279.

Johnson, C.D., Rand, J.R., Herman, R.K., Stern, B.D., and Russell, R.L. 1988. The

acetylcholinesterase genes of C. elegans: Identification of a third gene (ace-3) and

mosaic analysis of a synthetic lethal phenotype. Neuron 1, 165-173.

Kayser, B., Rajaram, S., Thomas, S., Morgan, P.G., and Sedensky, M.M. 1998. Control

of anesthetic response in C. elegans. Toxicol. Lett. 100-101, 339-346.

Kemp, J.R., and Wallace, K.B. 1990. Molecular determinants of the species-selective

inhibition of brain acetylcholinesterase. Toxicol. App. Pharmacol. 104, 246-258.

Khan, W.A., Dechkovskaia, A.M., Herrick, E.A., Jones, K.H., and Abou-Donia, M.B.

2000. Acute sarin exposure causes differential regulation of choline

acetyltransferase, acetylcholinesterase, and acetylcholine receptors in the central

nervous system of the rat. Toxicol. Sci. 57, 112-120.

Khanna, N., Cressman, C.P., Tatara, C.P., and Williams, P.L. 1997. Tolerance of the

nematode Caenorhabditis elegans to pH, salinity and hardness in aquatic media.

Arch. Environ. Contam. Toxicol. 32, 110-114. 35

Kolson, D.L., and Russell, R.L. 1985a. New acetylcholinesterase-deficient mutants of the

nematode Caenorhabditis elegans. J. Neurogenet. 2, 69-91.

Kolson, D.L., and Russell, R.L. 1985b. A novel class of acetylcholinesterase, revealed

by mutations, in the nematode Caenorhabditis elegans. J. Neurogenet. 2, 93-110.

Korsak, R.J., and Sato, M.M. 1977. Effects of chronic organophosphate pesticide

exposure on the central nervous system. Clin. Toxicol. 11, 83-95.

Lewis, J.A., Wu., C.H., Vevine, J.H., and Berg, H. 1980. Levamisole-Resistant mutants

of the nematode Caenorhabditis elegans appear to lack pharmacological

acetylcholine receptors. Neuroscience 5, 967-989.

Li, H., Avery, L., Denk, W., and Hess, G.P. 1997. Identification of chemical synapses in

the pharynx of Caenorhabditis elegans. Proc. Natl. Acad. Sci. USA 94, 5912-

5916.

Li, Q., Nagahara, N., Takahashi, H., Takeda, K., Okumura, K., and Minami, M. 2002.

Organophosphorus pesticides markedly inhibit the activities of natural killer,

cytotoxic T lymphocyte and lymphokine-activated killer: a proposed inhibiting

mechanism via granzyme inhibition. Toxicology 172, 181-190.

Link, C.D., 2001. Transgenic invertebrate models of age-associated neurodegenerative

diseases. Mech. Ageing Dev. 122, 1639-1649.

Llorens, J., Crofton, K.M., Tilson, H.A., Ali, S.F., and Mundy, W.R. 1993.

Characterization of disulfoton-induced behavioral and neurochemical effects

following repeated exposure. Fundam. Appl. Toxicol. 20, 163-169. 36

Lynch, M.R., Rice, M.A., and Robinson, S.E. 1986. Dissociation of locomotor

depression and ChE activity after DFP, soman, and sarin. Pharmacol. Biochem.

Behav. 24(4), 941-947.

Marrs, T.C. 1993. Organophosphate poisoning. Pharmacol. Ther. 58, 51-66.

Matsumura, F. 1985. Metabolism of insecticides by animals and plants. In: Toxicology

of Insecticides, Second Edition. Plenum Press, New York, NY, USA, pp. 203-298.

McGhee, J.D. 1987. Purification and characterization of a carboxylesterase from the

intestine of the nematode Caenorhabditis elegans. Biochem. 26, 4101-4107.

McGhee, J.D., Birchall, J.C., Chung, M.A., Cottrell, D.A., Edgar, L.G., Svendsen, P.C.,

and Ferrari, D.C. Kolson, D.L., and Russell, R.L. 1990. Production of null mutants

in the major intestinal esterase gene (ges-1) of the nematode Caenorhabditis

elegans. Genetics 125, 505-514.

McIntire, S.L., Jorgensen, E., Kaplan, J., and Horvitz, H.R. 1993. The GABAergic

nervous-system of Caenorhabditis elegans. Nature 364, 337-341.

Menzel, R. Bogaert, T., and Achazi, R. 2001. A systematic gene expression screen of

Caenorhabditis elegans Cytochrome P450 genes reveals CYP35 as strongly

xenobiotic inducible. Arch. Biochem. Biophys. 395(2), 158-168.

Metcalf, R.L., and March, R.B. 1953. Reversed phase paper chromatography of

parathion and related phosphate esters. Science 117, 527-528.

Meyers, D.K., and Mendel, B. 1952. Oxidation of thiophosphate insecticides in the rat.

Nature 170, 805-807.

Morgan, P.G., and Cascorbi, H.F. 1985. Effect of anesthetics and a convulsant on

normal and mutant Caenorhabditis elegans. Anesthesiology 62, 738-744. 37

Morgan, P.G., and Sedensky, M.M. 1994. Mutations conferring new patterns of

sensitivity to volatile anesthetics in Caenorhabditis elegans. Anesthesiology 81,

888-898.

Morgan, P.G., and Sedensky, M.M. 1995. Mutations affecting sensitivity to ethanol in

the nematode, Caenorhabditis elegans. Alcohol. Clin. Exp. Res. 19(6), 1423-1429.

Morgan, P.G., Sedensky, M.M., Meneely, P.M., and Cascorbi, H.F. 1988. The effect of

two genes on anesthetic response in the nematode Caenorhabditis elegans.

Anesthesiology 69, 246-251.

Moser, V.C. 1988. Comparison of chlordimeform and carbaryl using a functional

observational battery. Fundam. Appl. Toxicol. 11, 189-206.

Moser, V.C. 1989. Screening approaches to neurotoxicity: A functional observational

battery. J. Am. Coll. Toxicol. 8, 85-93.

Moss, D.E., and Fahrney, D. 1978. Kinetic analysis of differences in brain

acetylcholinesterase from fish or mammalian sources. Biochem. Pharmacol. 27,

2693-2698.

Murphy, S.D., and Dubois, K.P. 1957. Enzymatic conversion of the dimethoxy ester of

benzotriazine dithiophosphoric acid to and anticholinesterase agent. Journal of

Pharmacol. Exp. Ther. 119, 572-583.

Nelson, D.R. 1999. Cytochrome P450 and the individuality of species. Arch. Biochem.

Biophys. 369(1), 1-10.

O’Brien, R.D. 1960. Introduction. In: Toxic Phosphorus Esters. Chemistry,

Metabolism and Biological Effects. Academic Press, New York, NY, USA, pp.

1-28. 38

Padilla, S., Moser, V.C., Pope, C.N., and Brimijoin, W.S. 1992. Paraoxon toxicity is not

potentiated by prior reduction in blood acetylcholinesterase. Toxicol. Appl.

Pharmacol. 117, 110-115.

Pope, C.N. 1999. Organophosphorus pesticides: do they all have the same mechanism of

toxicity? J. Toxicol. Environ. Health, Part B 2, 161-181.

Rand, J.B., and Nonet, M.L. 1997. Synaptic transmission. In: Riddle, D.L., Blumenthal,

T., Meyer, B.J., and Priess, J.R. (Eds.), C. ELEGANS II. Cold Springs Harbor

Laboratory Press, New York, pp. 611-644.

Ray, D.E. 1998. Chronic effects of low level exposure to anticholinesterases – a

mechanistic review. Toxicol. Lett. 102-103, 527-533.

Richardson, R.J. 1995. Assessment of the neurotoxic potential of chlorpyrifos relative to

other organophosphorus compounds: a critical review of the literature. J. Toxicol.

Environ. Health 44, 135-165.

Riddle, D.L., Blumenthal, T., Meyer, B.J., and Priess, J.R. (Eds.). 1997. C. ELEGANS II.

Cold Spring Harbor Laboratory Press, Plainview, NY, USA.

Schnabel, R., and Priess, J.R. 1997. Specification of cell fates in the early embryo. In:

Riddle, D.L., Blumenthal, T., Meyer, B.J., and Priess, J.R. (Eds.), C. ELEGANS II.

Cold Springs Harbor Laboratory Press, New York, pp. 361-382.

Sheets, L.P., Hamilton, B.F., Sangha, G.K., and Thyssen, J.H. 1997. Subchronic

neurotoxicity screening studies with six organophosphate insecticides: an

assessment of behavior and morphology relative to cholinesterase inhibition.

Fundam. Appl. Toxicol. 35, 101-119. 39

Sulston, J., Dew, M., and Brenner, S. 1975. Dopaminergic neurons in the nematode

Caenorhabditis elegans. J. Comp. Neur. 163, 215-226.

Sulston, J.E., Schierenberg, E., White, J.G., and Thompson, J.N. 1983. The embryonic

cell lineage of the nematode Caenorhabditis elegans. Dev. Biol. 100, 64-119.

Thomas, J.H. 1994. The mind of a worm. Science 264, 1698-1699. van Rossum, A.J., Brophy, P.M., Tait, A., Barrett, J., and Jefferies, J.R. 2001. Proteomic

identification of glutathione S-transferases from the model nematode

Caenorhabditis elegans. Proteomics 1, 1463-1468.

Wallace, K.B. 1992. Species-selective toxicity of organophosphorus insecticides: a

pharmacodynamic phenomenon. In: Chambers, J. and Levi, P. (Eds.),

Organophosphates Chemistry, Fate, and Effects. Academic Press, Inc., San

Diego, CA, USA, pp. 3-17.

Wei, A., Jegla, T., and Salkoff, L. 1996. Eight potassium channel families revealed by

the C. elegans genome project. Neuropharmacol. 35(7), 805-829.

White, J.G., Southgate, E. Thompson, J.N., and Brenner, S. 1986. The structure of the

nervous system of the nematode Caenorhabditis elegans. Philos. Trans. R. Soc.

London B. 314, 1-340.

Williams, P.L., and Anderson, G.L. 2000. Caenorhabditis elegans as an alternative

animal species. J. Toxicol. Environ. Health, Part A 61, 641-647.

Williams, P.L., and Dusenbery, D.B. 1988. Using the nematode Caenorhabditis elegans

to predict mammalian acute lethality to metallic salts. Toxicol. Ind. Health 4, 469-

478. 40

Williams, P.L., and Dusenbery, D.B. 1990. A promising indicator of neurobehavioral

toxicity using the nematode Caenorhabditis elegans and computer tracking.

Toxicol. Ind. Health 6, 425-440.

Wood, W.B. Introduction to C. elegans Biology. 1988. In: Wood, W.B. (Ed.), The

Nematode Caenorhabiditis elegans. Cold Spring Harbor Laboratory, Plainview,

NY, USA, pp. 1-16.

Yang, R.S. 1976. Enzymatic conjugation and insecticide metabolism. In: Wilkinson,

C.F. (Ed.), Insecticide Biochemistry and Physiology. Plenum Press, New York,

NY, USA, pp. 177-225.

Zuckerman, B.M., and Himmelhoch, S. 1980. Nematodes as models to study aging. In:

Zuckerman, B.M. (Ed.), Nematodes as Biological Models. Vol. 2. Aging and other

models. Academic Press, New York, USA, pp. 4-28. 41

Table 2.1. Basic Description of Chemicals Tested.

Compound aUse/Accepted bRat LD50 aWater Solubility cStructure Mechanism (mg/kg) (mg/L)

Demeton-S-methylsulfone dInsecticide C6H15O5PS2 Metabolite 32.4 > 2.2x104 FW: 262.3 Cholinesterase Inhibitor CAS: 1740-19-6

Dichlorvos Insecticide and C4H7Cl2O4P Acaricide 17 1.8x104 FW: 221.0 Cholinesterase Inhibitor CAS: 62-73-7

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com). d International Program on Chemical Safety. Environmental Health Criteria 197 (http://www.inchem.org/documents/ehc/ehc/ehc197.htm) 42

Table 2.1. Basic Description of Chemicals Tested (continued).

Compound aUse/Accepted bRat LD50 aWater Solubility cStructure Mechanism (mg/kg) (mg/L)

Dimethoate Systemic Insecticide & C5H12NO3PS2 Acaricide 60 2.3x104 FW: 229.2 Cholinesterase Inhibitor CAS: 60-51-5

Ethephon Plant Growth C2H6ClO3P Regulator 3400 1.0x106 FW: 144.5 Ethylene Production CAS: 16672-87-0

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com). 43

Table 2.1. Basic Description of Chemicals Tested (continued).

Compound aUse/Accepted bRat LD50 aWater Solubility cStructure Mechanism (mg/kg) (mg/L)

Ethyl Paraoxon dInsecticide

C10H14NO6P Metabolite 1.8 e3632 FW: 275.2 Cholinesterase Inhibitor CAS: 311-45-5

Fenamiphos Systemic nematicide for C13H22NO3PS plant application 8 400 FW: 303.4 Cholinesterase Inhibitor CAS: 22224-92-6

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com). d National Toxicology Program Chemical Health and Safety Data. (http://ntp-server.niehs.nih.gov/htdocs/CHEM_H&S/NTP_Chem3/Radian311-45-5.html). e Bowman, B.T., and Sans, W.W. 1983. J. Environ. Sci. Health Part B 18, 667-683 44

Table 2.1. Basic Description of Chemicals Tested (continued).

aUse/Accepted bRat LD50 aWater Solubility cStructure Compound Mechanism (mg/kg) (mg/L)

Fensulfothion dSystemic Nemataicide C11H17O4PS2 & Insecticide d 2.2 1540 FW: 308.3 Cholinesterase Inhibitor CAS: 115-90-2

Glyphosate Systemic C3H8NO5P Herbicide 4873 1.16x104 FW: 169.1 Inhibitor of plant aromatic acid CAS: 1071-83-6 biosynthesis

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com). d Milne, G.W.A. (Ed.) 1995. CRC Handbook of Pesticides. CRC Press, Boca Raton, FL, USA, p.89.

45

Table 2.1. Basic Description of Chemicals Tested (continued).

aUse/Accepted bRat LD50 aWater Solubility cStructure Compound Mechanism (mg/kg) (mg/L)

Methamidophos Systemic Insecticide C2H8NO2PS & Acaracide 7.5 2x105 FW: 141.1 Cholinesterase Inhibitor CAS: 10265-92-6

Methidathion Non-systemic Insecticide & Acaricide C6H11N2O4PS3 20 (mg/kg) 200 Cholinesterase FW: 302.3 Inhibitor

CAS: 950-37-8

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com).

46

Table 2.1. Basic Description of Chemicals Tested (continued).

aUse/Accepted bRat LD50 aWater Solubility cStructure Compound Mechanism (mg/kg) (mg/L)

Methyl Parathion Non-systemic C8H10NO5PS Insecticide & Acaracide 6 55 FW: 263.2 Cholinesterase Inhibitor CAS: 298-00-0

Mevinphos Systemic Insecticide C7H13O6P & Acaricide 3 FW: 224.1 Cholinesterase miscible Inhibitor CAS: 7786-34-7

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com).

47

Table 2.1. Basic Description of Chemicals Tested (continued).

aUse/Accepted bRat LD50 aWater Solubility cStructure Compound Mechanism (mg/kg) (mg/L)

Monocrotophos Systemic Insecticide C7H14NO5P & Acaracide 8 miscible FW: 223.2 Cholinesterase Inhibitor CAS: 2157-98-4

Omethoate Systemic Insecticide C5H12NO4PS & Acaracide 30 readily FW: 213.2 Cholinesterase soluble Inhibitor CAS: 1113-02-6

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com).

48

Table 2.1. Basic Description of Chemicals Tested (continued).

aUse/Accepted bRat LD50 aWater Solubility cStructure Compound Mechanism (mg/kg) (mg/L)

Parathion Non-systemic C10H14NO5PS Insecticide & Acaracide 2 11 FW: 291.3 Cholinesterase Inhibitor

CAS: 56-38-2

a The Pesticide Manual, 12th ed. 2000. Tomlin, C.D (Ed.). 2000. . British Crop Protection Council, Farnham, UK. b RTECS Database, National Institute for Environmental Health. c Images obtained from Chemfinder.com (http://www.chemfinder.com).

49

CHAPTER 3

THE NEMATODE CAENORHABDITIS ELEGANS AS A MODEL OF

ORGANOPHOSPHATE INDUCED MAMMALIAN NEUROTOXICITY1

______

1Cole, R.D. and P.L. Williams. To be submitted to Toxicology and Applied Pharmacology. 50

ABSTRACT

Fifteen organophosphate (OP) pesticides were tested by computer tracking for their acute behavioral toxicity with the nematode Caenorhabditis elegans. EC50 values

for each OP were compared to the corresponding LD50 acute lethality value in rats and

mice. Order of toxicity was found to be significantly correlated in comparisons of C.

elegans to both rats and mice. Mechanistic investigations were conducted by assaying 8

of the 15 OPs for anticholinesterase activity in C. elegans. Significant cholinesterase

inhibition was confirmed for 4 OPs which had displayed high behavioral toxicity while 3

OPs of low behavioral toxicity showed no significant decrease in cholinesterase activity.

One highly toxic compound assayed yielded equivocal results. Toxicity for 2 chemicals was linked to pH effects. Detailed comparison of individual chemicals and metabolic issues are discussed. These results have positive implications for the use of C. elegans as a mammalian neurological model, and support the use of C. elegans in early rounds of chemical toxicity screening. 51

INTRODUCTION

The study of biological questions often proceeds through the investigation of

model systems. Of those organisms which have become established as biological models few have the scope and breadth of the nematode Caenorhabditis elegans. C. elegans, a

non-parasitic bacterial feeder which lives in soil interstitia, has served as a model in

neurobiology, developmental biology and genetics. The potential of C. elegans as a

model organism was recognized by Brenner (1974) early in the process of its initial

characterization. Several unique advancements in the study of C. elegans biology over

the past 30 years have brought Brenner’s insight to fruition. In 1986, White et al.

published a complete map of the C. elegans nervous system reconstructed from serial

section electron micrographs. A few years later, Sulston et al. (1983) completed a

milestone in developmental biology by tracing the linage of each cell in the adult

hermaphrodite from differentiation to death. More recently, genetic distinction has come

to C. elegans by being the first multicellular organism to have its genome completely

sequenced (C. elegans Sequencing Consortium, 1998). The enormous body of C. elegans

research which is highlighted by these accomplishments arguably makes C. elegans the

most thoroughly characterized organism on earth.

C. elegans is one of the simplest organisms with a centralized nervous system.

Despite the phylogenetic distance between nematodes and vertebrates, they share many

similarities with respect to neural physiology. The genomic sequencing of C. elegans has

brought to light an unexpected level of conservation with vertebrates. Homology and

conserved gene structure have been identified in genes responsible for many neural

components including ion channels, neurotrasnsmitter biosynthetic enzymes, synaptic 52

release mechanisms, neurotransmitter receptors and second messenger systems

(Bargmann, 1998). Strong evidence for neurotransmitter mediated transduction of the

synapse in C. elegans has been demonstrated for acetylcholine, GABA, serotonin,

dopamine, glutamate, and several neurogenic peptides (Rand and Nonet, 1997).

As in higher animals, acetylcholine (ACh) is the primary excitatory

neurotransmitter involved in C. elegans motor function (Rand and Nonet, 1997). Also

conserved is the synaptic degradation of ACh by acetylcholinesterase (AChE) enzymes.

While vertebrates have a single AChE gene whose product can undergo alternative

splicing post-transcriptionally to yield different isoforms, C. elegans has 4 separate genes

yielding 4 AChE enzymes (Combes et al., 2001). Two of these enzymes, ACE-1 and

ACE-2, account for roughly 95% of all AChE activity in C. elegans. Though they share

only 35% identity, ACE-1 and ACE-2 display a high degree of functional redundancy

(Combes et al., 2001). Johnson and Russell (1983) found that dual mutations nullifying

the product of both genes were required to produce a phenotype readily distinguishable

from the wild type. ACE-3 and ACE-4 are minor forms of AChE in C. elegans

accounting for approximately 5% and 0.1% of wild type activity, respectively (Combes et

al., 2001). ACE-3 exhibits enough activity for ace-1, ace-2 double mutants to be viable,

but movement and development are severely degraded.

There are several logistical considerations that make C. elegans appealing as a

neurological model. With a total of 302 neurons in the adult hermaphrodite, C. elegans

offers simplified anatomical complexity compared to higher organisms without

sacrificing its genetic and physiological applicability. The progression of each neural cell

has been characterized (Sulston et al., 1983), and the almost invariant arrangement of its 53

5000 chemical synapses and 2000 neuromuscular junctions is known (White et al., 1986).

This relatively simple nervous system supports a small number of basic behaviors which

have been studied quantitatively (Sulston and Hodgkin, 1988).

C. elegans has been used in testing environmental contaminants known to have

neural toxic effects. By developing a sort of neurotoxicity index from lethality data,

Williams and Dusenbery (1990a) concluded that lead (Pb), mercury (Hg), and two

organophosphate pesticides (malathion and dichlorvos) displayed concentration-response dynamics consistent with neurotoxicity while copper (Cu) did not. The idea that movement behavior may be preferentially sensitive to neurotoxicants has lead several

researchers to use locomotion in the examination of chemicals known to be neurotoxic.

Anderson et al. (2001) demonstrated that movement in C. elegans was more sensitive to

Pb than to Cu following 4-h exposures while this difference in sensitivity was masked

after exposures lasting 24-h.

Efforts have also been made to evaluate C. elegans as a model for mammalian

toxicity. Williams and Dusenbery (1988) found that acute lethality values in C. elegans

exposed to 8 heavy metallic salts predicted overall mammalian acute lethality as

consistently as tests in rats and mice individually. Williams and Anderson (2000)

conducted blind tests on 5 gadolinium-based compounds used in magnetic resonance

imaging measuring movement behavior as the quantified endpoint. These tests correctly

ordered the toxicity for 4 out of the 5 compounds compared to median lethal doses for the

same compounds in mice. Anderson et al. (submitted) found behavioral toxicity testing

following 4-h exposures correctly ordered 5 compounds across 3 chemical classes when

referenced to toxicity in rats and mice. 54

The purpose of this investigation was to explore the use of C. elegans as a model for mammalian neurotoxicity by testing its response to organophosphate pesticides which are a well known class of neurotoxic chemicals. We looked at this question in two ways.

Our first hypothesis was that the relative order of toxicity for a group of OPs tested in C. elegans using behavioral methods would bear significant similarity to their toxicity order in mammals. We also hypothesized that cholinesterase inhibition would be the mechanism by which the OPs exert their toxic effects in C. elegans.

MATERIALS AND METHODS

Exposure Solutions

Fifteen organophosphate chemicals (Table 1) were selected for testing based on two factors: water solubility and mammalian toxicity (lethality). To eliminate the need

for more than one exposure vehicle, only organophosphates which displayed toxic effects

within their published solubility limits were used. Compounds were also chosen to

represent a range of mammalian toxicities. All OPs were ordered from Sigma-Aldrich

(St. Louis, MO, USA) as analytic grade chemicals. Directly prior to testing, a working

stock solution of each compound was made by dissolution in K-medium (0.032M KCl,

0.051M NaCl) (Williams and Dusenbery, 1990b) on a mass/volume basis using a digital

scale and volumetric flasks. The concentration of stock solutions were kept within

published aqueous solubility limits. Immediately before each test, working stock

solutions were diluted in K-medium to the desired test concentrations.

55

Worm Cultures and Exposures

All tests were performed on C. elegans strain N2 (wild type) obtained from the

Caenorhabditis Genetics Center (Minneapolis, MN, USA). Age synchronous

populations of N2s were obtained by the collection and culturing of eggs laid by

emergent dauerlarvae (Donkin and Williams 1995). All developmental stages were

cultured on 115 mm Petri dishes of K-agar (0.032M KCl, 0.051M NaCl, 0.1M CaCl2,

0.1M MgSO4, 2.5% Bacto-peptone, 0.17% Bacto-agar, and 0.01% cholesterol)

(Williams and Dusenbery 1988) seeded with Escherichia coli OP50 to serve as a food

source (Brenner 1974). Since starvation has been shown to impact behavioral toxicity

measures (Anderson et al., 2001; Boyd et al., 2003), a high nutritional status was

maintained by transferring 2-d juveniles to plates with fresh OP50 lawns in a few drops

of K-medium. All exposures were carried out using 3-d adults.

Exposures were performed in 12-well sterile tissue culture plates. Each test

consisted of six 1 ml exposure volumes including one control (K-medium), and 5

different concentrations of the chemical in question. Approximately 100 worms were

transferred in 5µl to each exposure solution using a micropipetter fitted with a non-

heparinized hematocrit tube. All exposures were 4-h in length and were carried out in a

20 0C incubator in the absence of food.

Behavioral Assessment using Computer Tracking

Movement of exposed populations was quantified by computerized tracking in a manner patterned after Boyd et al. (2000). Following exposure, the contents of each exposure well were transferred by pipette to 2ml centrifuge tubes. When the worms had settled 56

into a pellet, the supernatant was siphoned off and replaced with approximately 1.5ml of

fresh K-medium and the worms were resuspended with a Pasteur pipette. This process

was repeated two times for a total of 3 rinses. After the final rinse, approximately 50

worms were transferred in 5µL of K-medium to 1% agar disks on glass slides. The

worms were allowed to disperse over the surface of the gel for approximately 1-h while

inverted over a Petri dish of water to avoid desiccation. The agar pads were then placed

in a weak, humidified air stream and the worms were tracked using the computer

automated methods of Boyd et al. (2000) and NIH Image Tracker software as modified

by David Dusenbery (1996). This software captures the position of up to 300 worms

once every second for 100-sec, and writes this data directly to an Excel spreadsheet. A

movement index (MI) was calculated from this data representing the average distance

(µm) moved by a single worm per second. Movement behavior was characterized as a

percentage of control movement collected on the same day. Each control and exposed

group was tracked for 3 cycles on a given day, and each day’s measurements were

considered one replicate. Three replicates were carried out for each chemical.

The concentration-response data was tested for normality (Chi square), and

homogeneity of variance (Levene’s test) using the statistical modeling program Toxstat® v.3.5 (Gulley, 1996). Once normality and homogeneity of variance were established,

Toxstat was also used to estimate the concentration effectively decreasing the average distance moved per worm by 50% (EC50) for each compound using probit analysis.

The EC50 values from the toxicity testing portion of the experiment were compared to acute oral LD50 values in rats and mice for the same compounds obtained from the Registry of Toxic Effects of Chemical Substances (RTECS) database. Acute 57

oral LD50s in rats and mice were chosen for comparison because they were the most

widely available values for the compounds tested. Rat LD50 values were available for all compounds tested, while mouse oral LD50 values were found for 13 of the 15 OPs.

Because the formula weights for the compounds varied by as much as a factor of two, it

seemed best to use millimoles per liter (mM) as the common concentration unit for

comparisons across compounds. LD50 values for rats and mice were converted from

mg/kg to millimoles per liter body water using values from the Handbook of Biological

Data (Spector, 1956) for average water volume per kg body weight. These values were

660ml/kg and 766ml/kg for rats and mice, respectively.

pH Measurements and Testing

Pre and post-exposure pH measurements were made on the high and low exposure concentrations used for each test compound using an Orion 720A multimeter (Orion

Research, Beverly, MA, USA). While collecting behavioral data, the effect of decreasing pH on behavior was recognized as a factor that needed to be addressed. Behavioral testing with pH adjusted K-medium was carried out in a manner identical to tests with

OPs. The pH levels tested included ~5.8 (the unadjusted pH of K-medium), 3.5, 3.0, 2.5,

2.0, and 1.5. Adjustments to test solution pH were made with HCl.

Cholinesterase Activity Assay

The colorimetric assay of Ellman et al. (1961) has long been used to assess

cholinesterase activity in a wide range of animals including mammals (Blakley and Yole, 58

2002), birds (Hill and Flemming 1982), and crustaceans (Mora et al. 1999). Various researchers have modified the Ellman method over the years to better suit their particular needs. Our protocol most closely resembled the adaptations of Moulton et al. (1996) to better measure ChE activities in low activity tissues.

Cholinesterase activity measurements were carried out for 8 of the 15 chemicals used in behavioral tests. Because the desired comparison was between control and exposed groups for a single chemical, the exposure and collection of samples for each compound was organized as a randomized complete block design with day as the blocking factor. ANOVA was used to test for significant differences between groups using the general linear model in SAS (Cary, NC, USA).

Cholinesterase measurements were collected from worms exposed to dichlorvos, parathion, methyl parathion, methidathion, fensulfothion, glyphosate, ethephon, and demeton-S-methylsulfone. Three exposure levels were used for each chemical including an unexposed control, a mid-level concentration near the behavioral EC50, and the highest concentration used in the behavioral assays. Exposures were conducted in 12- well tissue culture plates with 4 wells devoted to each concentration. Chemical stock solutions were diluted in K-medium to the appropriate concentration in each well so that the final exposure volume was 1ml. Worms used in the ChE assays were 3-d adults reared, handled, and exposed identically to those used in the behavioral studies.

Following the 4-h exposure period, each exposure concentration was transferred to a

15ml centrifuge tube. The worms were rinsed as previously described for behavioral tests, and were transferred to microcentrifuge tubes. The samples were centrifuged at 59

2500rpm and 40C for 10-min. The supernatant was removed, and the worms were frozen

in liquid nitrogen before storage at -800C.

At the time the ChE assays were carried out each sample was thawed in 200µl of

50mM Tris buffer, pH 7.4, and was homogenized with a teflon pestle. The samples were again centrifuged at 40C and 2500rpm for 10-min. A 160µl portion of the supernatant

from each sample was mixed with 1200µl of 0.25mM 5,5’-Dithiobis(2-nitrobenzoic acid)

(DTNB), and 40µl of 156mM acetylthiocholine iodide (ASChI) in a 1.5ml cuvette.

These contents were rapidly mixed and the rate of change in absorbance was measured at

30-s intervals for 90-s using a Shimadzu UV-1601 spectrophotometer (Shimadzu

Scientific Instruments, Columbia, MD, USA). Absorbance was measured at 405nm.

Kinetic measurements were recorded and later converted to total cholinesterase activity

using the extinction coefficient for the colored product, 5-thio-2-nitro-benzoic acid (II)

(Ellman et al. 1961).

Protein content of the worm-Tris homogenate supernatant was also determined

using the Bio-Rad protein assay kit (Bio-Rad Laboratories, Hercules, CA, USA).

Bovine serum albumin (BSA) was used as a protein standard, and the colored product of

the dye binding reaction was measured spectrophotometrically by absorbance at 595nm.

The protein content of each sample was used to standardize the ChE activity to a per mg

protein basis.

60

RESULTS

Concentration-Response Data

Concentration response curves were generated for each compound tested by plotting the level of movement (per cent control) for each group of exposed worms against the concentration at which they were exposed. The 15 compounds tested displayed a wide range of behavioral toxicities with EC50 values covering more than 4.5 orders of magnitude (Table 3.1). This is similar to rats and mice which exhibited toxicity ranges of roughly 3.5 and 4 orders of magnitude, respectively. If the movement data are plotted by compound (Fig. 3.1), the overall trends of progressing toxicity can be examined. Data points for C. elegans and the mammals are tightly clustered for some chemicals (e.g. parathion) while for others they are not (e.g. dichlorvos).

There were a few compounds whose EC50 values could not easily be distinguished from one another. This was problematic since the values were to be used in a rank order comparison. To address this concern, the values in question were tested for significant difference (α=0.05) using a paired t-test. No significant difference was found between the EC50s of ethyl paraoxon and methyl parathion, or for methyl parathion and parathion. EC50s for ethyl paraoxon and parathion, however, were found to differ significantly. The toxicity rankings for ethyl paraoxon, methyl parathion, and parathion in C. elegans were therefore designated as 2.5, 3, and 3.5 respectively (Table 3.1). This procedure also indicated a lack of any significant difference between the EC50s for ethephon and methamidophos which were also given equal ranking. Since no estimates of variability or uncertainty were given for the values in the RTECS database, these estimates were taken at face value and ranked accordingly. 61

The toxicity values (EC50 or LD50) for each test organism were ranked, and the rank orders were compared using Spearman’s Rank Correlation Coefficient. The results

of these comparisons display a significant correlation between the orders of toxicity for

the tested compounds in C. elegans, rats and mice (Table 3.1). In the comparison of C.

elegans with rats, the ranks were highly correlated meeting the standard for significance

at the α=0.01 level. The correlation between C. elegans and mice was significant at

α=0.05.

Cholinesterase Activity Assay

Significant reductions in ChE activity were observed for parathion, methyl

parathion, methidathion, and fensulfothion. In each case, both the mid and high exposures were found to be significantly different from controls (α=0.05). No chemical

displayed a difference in ChE activity between the mid and high exposure levels.

Dichlorvos seemed to display a large concentration dependent decrease in cholinesterase

activity (Fig. 3.2), but contained too much variability to achieve significance at the

α=0.05 level (p=0.061). Exposures to the remaining compounds, glyphosate, ethephon,

and demeton-S-methylsulfone, failed to result in any significant decrease in ChE activity.

Mid level exposures to glyphosate and ethephon appear to have no impact what-so-ever

on ChE activity (Fig. 3.2), but tests did show a qualitative decrease for both chemicals.

At the time of sample collection, noticeable worm mortality was observed at the high

exposure level for these two chemicals.

62

Exposure Solution pH

Measurements of exposure solution pH were taken before and after worm

exposure. In all cases OP addition caused an initial decrease in solution pH followed by an increase over the 4-h exposure. Most pH decreases were moderate with reductions

from approximately 5.8 (K-medium) to a range of 4.4-5.7. The high concentrations of

glyphosate and ethephon, however, reduced the exposure pH to 2.2 and 2.0, respectively.

These measurements identified exposure pH as a possible confounding variable. For this reason, tracking experiments were performed using pH as the experimental variable to characterize the effect of decreasing pH on movement behavior. Following initial testing,

5 pH levels decreasing in 0.5 pH unit increments were tested between 3.5-1.5.

Movement in C. elegans appeared unaffected as pH decreased to pH 3.5, then began a linear decline to virtually no movement at pH 1.5 (Fig. 3.3).

DISCUSSION

Several studies comparing the effects of xenobiotics in C. elegans to those

observed in mammals have found strong similarities. Williams and Dusenbery (1988)

found that toxicity tests using C. elegans produced acute LC50 values for 8 metallic salts

that significantly predicted the order of toxicity of the same salts in mammals. Anton et

al. (1992) and Morgan and Sedensky (1995) found the influence of ethanol on C. elegans

to be qualitatively parallel to its action in mammals. These investigators also deemed C.

elegans to be useful as a mammalian model for studying the biologic response to volatile

anesthetics. Most recently, Anderson et al. (submitted) found that results from behavioral

testing with C. elegans correctly predicted the order of toxicity of 3 organic solvents and 63 two organic pesticides in mammals. The present study has produced a similar finding.

Fifteen organophosphate compounds tested for behavioral toxicity in C. elegans yielded

EC50 values whose rank order was significantly correlated to the order of rat LD50s at an α=0.01, and mouse LD50s at α=0.05. Genetic and physiologic similarities have been established between the cholinergic nervous system of C. elegans and higher organisms

(Bargmann, 1998). Our results support the usefulness of C. elegans as a mammalian biological model for systems where a physiologic basis for such comparison exists.

In this investigation we hypothesized that the mechanism of acute OP toxicity in

C. elegans would be the inhibition of AChE. We found that just as with higher animals,

OPs in C. elegans display a full range of anticholinesterase potential. Eight of the 15 OPs subjected to behavioral toxicity testing were assayed for anticholinesterase activity.

Parathion, methyl parathion, methidathion, and fensulfothion significantly inhibited ChE activity (Fig. 3.2). All of these compounds also exhibited high behavioral toxicity (EC50

< 1mM). Dichlorvos showed qualitative signs of ChE inhibition, but significant anticholinesterase activity could not quantitatively be established.

The levels of cholinesterase inhibition we observed are in agreement with levels that have been reported in other species. Exposures in our experiment were tailored to significantly reduce movement without being lethal. The 4 chemicals which showed significant ChE inhibition reduced ChE activity between 15% and 45% (Fig. 3.4). This is consistent with the work of Hill and Flemming (1982) who supported using 20% reduction in bird brain ChE activity as an indicator of field exposure to an anticholinesterase agent, and a reduction of 50% or more as grounds for attributing death to AChE inhibition. Sheets et al. (1997) also used a 20% reduction in the activity of rat 64 brain AChE as the cutoff in judging the biological significance of exposures. Moulton et al. (1996) recommend a similar but slightly more conservative 30% reduction from control levels as reflecting significant inhibition in freshwater mussel adductor tissue.

Our data indicate a 20% reduction in ChE activity in C. elegans as a threshold for biological significance. By these criteria, dichlorvos also seems to be implicated as an anticholinesterase agent (Fig 3.4). While this may be true, variability in the results for dichlorvos caused the data to be quantitatively inconclusive. Of the chemicals judged not to significantly inhibit ChE activity, only the high exposure of glyphosate decreased activity below 80%. This decrease was borderline (79.3%), and the mortality also observed at this exposure concentration raises doubts about the nature of the reduction.

Movement data for the ChE inhibiting compounds showed a decrease in locomotor activity of 25-30% between the EC50s and high exposure concentrations. It therefore seems odd that no significant decrease in ChE activity was observed between mid level and high exposure groups. Hill and Flemming (1982) argue that bird brain ChE inhibition of 50% or more can prove lethal. If this is also the case in C. elegans, cessation of movement should correspond to a ChE activity level at or above 50% controls. There also seems to be a leveling off in the response of ChE activity to increasing OP concentration (Fig 3.4) while movement continues to decrease. Our data seem to indicate a nonlinear relationship between ChE activity and movement with a given decrease in ChE activity resulting in proportionately larger reductions in locomotion as overall ChE activity levels decrease.

The mean ChE activity of the pooled control samples was 12.5nmol/mg protein/min with a standard deviation of 2.1. This control activity is similar to previously 65

reported values for other invertebrates. Mora et al. (1999) reported whole organism ChE

activities of approximately 13 and 20nmol/mg protein/min for the bivalves Mytilus

galloprovincialis and Corbicula fluminea, respectively. Variability in estimates of C. elegans’ background ChE activity (CV=17%; N=27) are also consistent with M. galloprovincialis (CV=30%; N=3) and C. fluminea (CV= 34%; N=6) (Mora et al.,

1999). Although background activities were higher for brain tissues in vertebrate species,

Blakley and Yole (2002) observed similar variability estimates for cattle (CV=30.7%;

N=33) and bald eagles (C =18.4%; N=29). These reports form a context in which our

results for background ChE activity in C. elegans seem reasonable.

Glyphosate, ethephon, and demeton-S-methylsulfone showed no significant

inhibition of ChE activity at the highest exposure levels which reduced movement in

behavioral tests to 17%, 7%, and 33% controls, respectively. This result indicates that

these compounds do not exert their effect on movement by poisoning the cholinergic

system. Some qualitative decrease in ChE activity was visible at the highest level of

glyphosate and ethephon exposure (Fig. 3.2), but this was most likely due to mortality

observed in these two groups over the course of the exposure period. This conclusion is

supported by the complete absence of reduction in ChE activity in worms exposed to

concentrations near the movement EC50s for these compounds. It is not surprising that

glyphosate and ethephon do not exert their toxicity by ChE inhibition. Glyphosate is the

active ingredient in Roundup®, a broad spectrum herbicide, while ethephon is used as a

regulator of plant growth and fruit maturation (Agrochemicals Handbook, 1987).

Bababunmi et al. (1978) site the uncoupling of oxidative phosphorylation and pulmonary hyperemia as being the major toxic lesions displayed in glyphosate toxicity in rats and 66

mice. Ethephon has been shown to inhibit butyrylcholinesterase (BChE), but appears to possess little affinity for AChE (Haux et al., 2000).

Behavioral toxicity in C. elegans appears to be better suited for predicting relative potency of OPs in mammals than predicting potency itself. Linear regression of log transformed C. elegans EC50s against log transformed rat LD50s yields a weak but not totally insignificant correlation (r2= 0.346). Worm EC50s are less predictive of mouse

LD50s (r2=0.241). Five compounds displayed considerably reduced toxicity in C. elegans compared to rats and mice. Fensulfothion was more than 30 times less toxic in

C. elegans than in rats (Table 1), but still significantly decreased ChE activity (Fig. 3.2).

Demeton-S-methylsulfone, in contrast, is reported to have a relatively high toxicity in

rodents (LD50rat =0.19mM), but showed low toxicity in C. elegans (EC50 =12.8 ±0.26

mM) with no evidence of significant anticholinesterase activity (Fig. 3.2). Three other

compounds (dimethoate, methamidophos, and omethoate) yielded EC50s 1.3 to 2.1

orders of magnitude higher than the corresponding LD50s. These OPs are listed by

Tomlin (2001) as cholinesterase inhibitors but were not tested here for their effect on

ChE activity in C. elegans.

There was only one chemical that that was much more toxic in C. elegans than in rats and mice. Though it ranked 10th in rat toxicity and 11th in mice, dichlorvos

(EC50= 0.7 ± 0.02 µM) was the most toxic compound tested in C. elegans by roughly an order of magnitude (Table 1). Dichlorvos accounted for one half of the summed squared difference between C. elegans and rat rankings, and for almost 75% of the difference between rankings of C. elegans and mice. High toxicity of dichlorvos to C. elegans has been described before. Williams and Dusenbery (1990a) exposed worms to vapona 67

(dichlorvos) dissolved in agar plates and reported concentrations that reduced average

movement behavior per worm by 50% (BC50). Not only is this measure analogous to

our EC50, but the BC50 of 0.8µM reported for dichlorvos is almost identical to our EC50

of 0.7µM. This high potency is also comparable to published toxicity values in other

invertebrates (Eisler, 1969). While dichlorvos yielded clear qualitative evidence of ChE

inhibition, it failed to demonstrate a significant quantitative decrease in ChE activity.

This result seems incongruous with its high degree of toxicity. Organophosphates have

been documented to have effects in mammals other than AChE inhibition (Pope, 1999).

It seems possible, therefore, that mechanisms other than or in addition to ChE inhibition

may be active at or below dichlorvos concentrations required to cause significant ChE

inhibition.

Inconsistencies in interspecies toxicity are not simply an effect of the phylogenetic distance between nematodes and mammals. Longstanding efforts to use interspecies correlations to build a predictive model of OP toxicity have met with little success (Wallace, 1992, Chambers and Carr, 1995). The potential of a xenobiotic to exert toxicity within any organism is a complicated interaction of multiple pharmacokinetic and pharmacodynamic parameters. The vast array of chemical characteristics displayed by OPs results in a high degree of individuality with regard to bioavailability, absorption, distribution, degree of AChE inhibition, metabolism, and elimination (Hutson and Millburn, 1991). Of these factors, biotransformation and the pharmacodynamic aspects of AChE inhibition have received the most attention in the literature. There are reports of a general resistance in fish to OP intoxication compared to birds and mammals (Kemp and Wallace, 1990; Chambers and Carr, 1995). Kemp and 68

Wallace (1990) found greater steric hindrance and a weaker nucleophilic center at the

esteratic site of rainbow trout AChE in comparison to the enzyme in hens and rats. This

lead them to hypothesize that greater substrate specificity in fish AChE results in reduced

enzyme phosphorylation, and reduced inhibition. In characterizing the two major C.

elegans AChEs, Johnson and Russell (1983) found binding affinities to be similar to

vertebrates. They also reported relatively low substrate specificity with only small

differences in the rates of hydrolysis for acetylthiocholine, propionly thiocholine, and

butyryl thiocholine. This pharmacodynamic explanation of interspecies variability is

therefore unlikely to be satisfactory for the comparison of C. elegans to mammals.

A possible alternative explanation is differing metabolism between nematodes

and mammals. Metabolism has been identified as a major factor in some cases of

interspecies toxicity difference (Hutson and Millburn, 1991). Chambers et al. (1994)

found that microsomal induction in rats with chemicals like phenobarbital and β-

Napthoflavone could alter the dynamics of parathion toxicity. They concluded that induction modified the relative rates of activation and detoxication reactions carried out by competing cytochrome P450s resulting in increased parathion activation but decreased overall toxicity (Chambers et al., 1994; Chambers and Carr, 1995).

Although our work was not designed to specifically investigate the metabolism of

OPs in C. elegans, some observations can be made. Phosphorothioates are a group of

OPs that must undergo oxidative desulfuration (Fig. 3.5) to the corresponding phosphate before displaying significant anticholinesterase activity (Metcalf and March, 1953). In mammals, this bioactivation is predominantly carried out by the microsomal cytrochrome

P-450 isozymes, primarily in the liver (Davison, 1955). Approximately 80 CYP 450s 69

have been identified in C. elegans (Nelson, 1999), and at lease one class of these have

been shown to be highly inducible by xenobiotics (Menzel et al., 2001). Five of the 15

compounds we tested were phosphothioates or phosphodithioates. Parathion, methyl parathion, fensulfothion (phosphorothioates), and methidathion (a phosphorodithioate) were assayed for anticholinesterase activity, and all were found to significantly decrease worm ChE activity. In addition, our behavioral toxicity tests resulted in an EC50 estimate for parathion (13 ± 0.4 µM) that was only slightly higher than its bioactive

metabolite ethyl paraoxon (11 ± 0.3 µM). These results not only indicate that C. elegans

is able to carry out the bioactivating desulfuration of phosphorothioates, but also that they are capable of doing so at a rate resulting in virtually no overall toxicity difference between some pro-toxicants and their active metabolites.

With respect to the 5 compounds showing a large decrease in potency in C. elegans compared to rats and mice, demeton-S-methylsulfone, dimethoate, methamidophos, and omethoate are structurally similar in that they all have a thioester bond. Three of these 4 also have carbonyl (dimethoate and omethoate) or sulfone

(demeton-S-methylsulfone) groups in the position either β or γ to the thioester sulfur.

These groups are susceptible to oxidation and hydrolytic detoxication in mammals

(Matsumura, 1985), and may be targeted for metabolism in C. elegans as well. The similarities among these compounds are interesting, but conclusions about why they are less toxic in C. elegans than in mammals are beyond the scope of this investigation.

Cholinesterase assays showed that glyphosate and ethephon did not decrease ChE

activity at concentrations causing toxicity. Other aspects of our investigation, however,

shed light on an alternative mechanism. The chemical structures of these two compounds 70

are unique among the chemicals tested in that they have two hydroxyl groups (-OH) as α substituents of the central phosphorous atom. This structure results in highly acidic pKas for both glyphosate (2.2 and 5.8) and ethephon (2.5 and 7.2) (Tomlin, 2000). These are near or below the normal pH of K-medium (approximately 5.8) causing both compounds to be significant proton donors in K-medium.

Glyphosate and ethephon lowered exposure solution pH to 2.2 and 2.0,

respectively. This was considerably greater than reductions caused by the other chemicals tested. Subsequent behavioral tests using pH as the experimental variable showed movement behavior was unaffected between pH 5.75 and 3.5, but exhibited a sharp, linear decrease between pH 3.5 and 2.0 where movement levels approached zero

(Fig. 3.3). This finding is consistent with the work of Kahana et al. (1997) who established the range of pH tolerance for C. elegans to be 3.1 to 11.9 over 24-h, and 3.2 to 11.8 over 96-h. All OPs caused some decrease in pH, but ethephon and glyphosate were the only two which caused a drop in pH below the pH 3.5 threshold even at the lowest concentrations tested. When the movement data for ethephon and glyphosate are plotted against the pH rather than concentration (Fig. 3.6), the response directly overlays the response exhibited to decreasing pH alone. This supports the idea that the behavioral toxicity observed for ethephon and glyphosate was primarily a pH effect. In contrast, all other compounds tested displayed the observed range of behavioral toxicity at pH levels above 3.5.

In conclusion, previous studies have shown C. elegans to possess many traits

favorable to its use as a biological model. It is easy to work with and cost effective as a

testing organism. The basic biology of C. elegans has been thoroughly studied revealing 71

many similarities between the nervous systems of nematodes and vertebrates. It has also

been shown to be sensitive to a variety of known neurotoxic chemicals. The results

presented in this study demonstrate the ability of C. elegans to serve as an alternative model for some types of neurotoxicity in mammals. Behavioral toxicity tests with C. elegans ordered the toxicity of 15 organophosphate pesticides in a manner statistically consistent with testing using rats and mice. Direct measurements of ChE activity confirmed significant ChE inhibition in C. elegans by fensulfothion, methidathion, methyl parathion, and parathion, which are all known to inhibit ChE in mammals.

Cholinesterase assays also verified a lack of significant ChE inhibition by glyphosate and ethephon, two OPs known to be only weak inhibitors of mammalian AChE. The anticholinesterase dichlorvos showed a very high toxicity and qualitative signs of ChE inhibition, but failed to demonstrate clear quantitative depression of ChE activity in C. elegans. Demeton-S-methylsulfone failed to inhibit ChE activity and was the only OP assayed for ChE inhibition which showed opposite dynamics in C. elegans compared to those reported in mammals. It has been suggested that C. elegans may react to neurologically active chemicals with enough similarity to mammals to be useful as a first round screening agent for neurotoxicity. Our results support such a role for C. elegans.

Further testing with more OPs and investigations into C. elegans metabolism should be conducted to further clarify the similarities and differences between OP toxicity in C. elegans and mammals. It would also be useful to conduct behavioral testing using chemicals other than OPs known to result in neurotransmitter related toxicity.

72

ACKNOWLEDGEMENTS

We would like to thank the Interdisciplinary Toxicology Program at the University of Georgia for contributing financial support to this research. We would also like to thank Dr. David Dusenbery for use of the computer tracking software program. Finally, we would like to thank the Caenorhabditis Genetics Center which is funded by the NIH

National Center for Research Resources (NCRR) for supplying the nematode strain used in this research. 73

REFERENCES

Agrochemicals Handbook, The. 1987. Royal Society of Chemistry, Information

Services. Cambridge, England.

Anderson, G.L., Boyd, W.A., and Williams, P.L. 2001. Assessment of sublethal

endpoints for toxicity testing with the nematode Caenorhabditis elegans. Environ.

Toxicol. Chem. 20, 833-838.

Anton, A.H., Berk, A.I., Nicholls, C.H. 1992. The “Anesthetic” effect of alcohols and

alkanes in Caenorhabditis elegans (C.e.). Research Communications in Chemical

Pathology and Pharmacology 78(1), 69-83.

Bababunmi, E.A., Olorunsogo, O.O., and Bassir, O. 1978. Toxicology of Glyphosate in

Rats and Mice. Toxicol. Appl. Pharmacol. 45, 319-320.

Bargmann, C.I. 1998. Neurobiology of the Caenorhabditis elegans genome. Science

282, 2028-2033.

Blakley, B.R., and Yole, M.J. 2002. Species differences in normal brain cholinesterase

activities of animals and birds. Vet. Human Toxicol. 44, 129-132.

Boyd, W.A. and Williams, P.L. 2003. Availability of metals to the nematode

Caenorhabditis elegans: Toxicity based on total concentrations in soil and extracted

fractions. Environ. Toxicol. Chem. 22, 1100-1106.

Boyd, W.A., Anderson, G.L., Dusenbery, D., Williams, P.L. 2000. Computer tracking

method for assessing behavioral changes in the nematode Caenorhabditis elegans.

In: Price, F.T., Brix, K.V., Lane, N.K., (Eds.), Environmental Toxicology and Risk

Assessment, Vol. 9. STP 1381. American Society for Testing and Materials,

Philadelphia, PA, USA, pp. 225-238. 74

Brenner, S. 1974. The genetics of Caenorhabditis elegans. Genetics 77, 71-94.

C. elegans Sequencing Consortium, The. 1998. Genome sequence of the nematode C.

elegans: a platform for investigating biology. Science 282, 2012-2018.

Chambers, J.E., and Carr, R.L. 1995. Biochemical mechanisms contributing to species

differences in insecticidal toxicity. Toxicology 105, 291-304.

Chambers, J.E., Ma, T.., Boone, J.S., and Chambers, H.W. 1994. Role of detoxication

pathways in acute toxicity levels of phosphorothionates insecticides in the rat. Life

Sci. 54, 1357-1364.

Combes, D., Fedon, Y., and Toutant, J. 2001. Acetylcholinesterase genes in the

nematode Caenorhabditis elegans. Int. Rev. Cytol. 209, 207-239.

Davison, A.N. 1955. The conversion of schradan (OMPA) and parathion to inhibitors of

cholinesterase by mammalian liver. Biochem. J. 61, 203-209.

Dhawan, R., Dusenbery, D.B., and Williams, P.L. 1999. Comparison of lethality,

reproduction and behavior as toxicological endpoints in the nematode

Caenorhabditis elegans. J. Toxicol. Env. Health, Part A 58, 451-462.

Donkin, S.G., and Williams, P.L. 1995. Influence of developmental stage, salts and food

presence on various end points using Caenorhabditis elegans for aquatic toxicity

testing. Environ. Toxicol. Chem. 14, 2139-2147.

Dusenbery, D. 1996. NIH Image, version 1.59 computer tracking program modified by

Dr. David Dusenbery, School of Biology, Georgia Institute of Technology, Atlanta,

GA.

Eisler, R. 1969. Acute toxicities of insecticides to marine decapod crustaceans.

Crustaceana 16, 302-310. 75

Ellman, G.L., Courtney, K.D., Andres, Jr. V., Featherstone, R.M. 1961. A new and

rapid colorimetric determination of acetylcholinesterase activity. Biochem.

Pharmacol. 7, 88-95.

Gulley, D.D. 1996. Toxstat® Version 3.5, Western EcoSystems Technology, Inc.

Cheyenne, WY, USA.

Haux, J.E., Quistand, G.B., and Casida, J.E. 2000. Phosphobutyrylcholinesterase:

Phosphorylation of the esteratic site of butyrylcholinesterase by ethephon [(2-

Chloroethyl)phosphonic acid] dianion. Chem. Res. Toxicol. 13, 646-651.

Hill, E.F., and Fleming, W.J. 1982. Anticholinesterase poisoning of birds: field

monitoring and diagnosis of acute poisoning. Environ. Toxicol. Chem. 1, 27-38.

Hutson, D.H. and Millburn, P. 1991. Enzyme-mediated selective toxicity of an

organophosphate and pyrethroid: some examples from a range of animals.

Biochem. Soc. Trans. 19, 737-740.

Johnson, C.D., and Russell, R.L. 1983. Multiple molecular forms of acetylcholinesterase

the nematode Caenorhabditis elegans. J. Neurochem. 41, 30-46.

Kemp, J.R., and Wallace, K.B. 1990. Molecular determinants of the species-selective

inhibition of brain acetylcholinesterase. Toxicol. App. Pharmacol. 104, 246-258.

Khanna, N., Cressman, C.P., Tatara, C.P., and Williams, P.L. 1997. Tolerance of the

nematode Caenorhabditis elegans to pH, salinity and hardness in aquatic media.

Arch. Environ. Contam. Toxicol. 32, 110-114.

Matsumura, F. 1985. Metabolism of insecticides by animals and plants. In: Toxicology

of Insecticides, Second Edition. Plenum Press, New York, NY, USA, pp. 203-298. 76

Menzel, R. Bogaert, T., and Achazi, R. 2001. A systematic gene expression screen of

Caenorhabditis elegans Cytochrome P450 genes reveals CYP35 as strongly

xenobiotic inducible. Arch. Biochem. Biophys. 395, 158-168.

Metcalf, R.L., and March, R.B. 1953. Reversed phase paper chromatography of

parathion and related phosphate esters. Science 117, 527-528.

Mora, P., Michel, X. Narbonne, J. 1999. Cholinesterase activity as potential biomarker

in two bivalves. Environ. Toxicol. Pharmacol. 7, 253-260.

Morgan, P.G., and Sedensky, M.M. 1995. Mutations affecting sensitivity to ethanol in

the nematode, Caenorhabditis elegans. Alcohol. Clin. Exp.Res. 19, 1423-1429.

Moulton, C.A., Fleming, W.J., and Purnell, C.E. 1996. Effects of two cholinesterase-

inhibiting pesticides on freshwater mussels. Environ. Toxicol. Chem. 15, 131-137.

Nelson, D.R. 1999. Cytochrome P450 and the individuality of species. Arch. Biochem.

Biophys. 369, 1-10.

Pope, C.N. 1999. Organophosphorus pesticides: do they all have the same mechanism of

toxicity? J. Toxicol. Environ. Health, Part B 2, 161-181.

Rand, J.B., and Nonet, M.L. 1997. Synaptic transmission. In: Riddle, D.L., Blumenthal,

T., Meyer, B.J., and Priess, J.R. (Eds.), C. ELEGANS II. Cold Springs Harbor

Laboratory Press, New York, pp. 611-644.

Sheets, L.P., Hamilton, B.F., Sangha, G.K., and Thyssen, J.H. 1997. Subchronic

neurotoxicity screening studies with six organophosphate insecticides: an

assessment of behavior and morphology relative to cholinesterase inhibition.

Fundam. Appl. Toxicol. 35, 101-119. 77

Spector, W.S. (Ed.), 1956. Handbook of Biological Data. WB Saunders Company,

Philadelphia, PA, USA. p. 340.

Sulston, J. and Hodgkin, J. 1988. Methods. In: Wood, W.B. (Ed.), The Nematode

Caenorhabditis elegans. Cold Spring Harbor Laboratory, Plainview, NY, USA, pp.

587-606.

Sulston, J.E., Schierenberg, E., White, J.G., and Thompson, J.N. 1983. The embryonic

cell lineage of the nematode Caenorhabditis elegans. Dev. Biol. 100, 64-119.

Tomlin, C.D (Ed.). 2000. The Pesticide Manual, 12th ed. British Crop Protection

Council, Farnham, UK.

Wallace, K.B. 1992. Species-selective toxicity of organophosphorus insecticides: a

pharmacodynamic phenomenon. In: Chambers, J. and Levi, P. (Eds.),

Organophosphates Chemistry, Fate, and Effects. Academic Press, Inc., San

Diego, CA, USA, pp. 3-17.

White, J.G., Southgate, E. Thompson, J.N., and Brenner, S. 1986. The structure of the

nervous system of the nematode Caenorhabditis elegans. Phios. Trans. R. Soc.

London B 314, 1-340.

Williams, P.L., and Anderson, G.L. 2000. Caenorhabditis elegans as an alternative

animal species. J. Toxicol. Environ. Health, Part A 61, 641-647.

Williams, P.L., and Dusenbery, D.B. 1988. Using the nematode Caenorhabditis elegans

to predict mammalian acute lethality to metallic salts. Toxicol. Ind. Health 4, 469-

478. 78

Williams, P.L., and Dusenbery, D.B. 1990a. A promising indicator of neurobehavioral

toxicity using the nematode Caenorhabditis elegans and computer tracking.

Toxicol. Ind. Health 6, 425-440.

Williams, P.L., and Dusenbery, D.B. 1990b. Aquatic toxicity testing using the nematode

Caenorhabditis elegans. Environ. Toxicol. Chem. 9, 1285-1290. 79

Table 3.1. Toxicity values and results for Spearman’s Rank Order Correlation comparison.

C. elegans aRat aMouse C. elegans Rat Mouse Chemical EC50 (sd) (mM) LD50 (mM) LD50 (mM) Rank Rank Rank Dichlorvos 7E-4 (2E-5) 0.117 0.360 1 10 11 Ethyl Paraoxon 0.011 (3E-4) 0.010 0.004 2.5 1 1 Methyl Parathion 0.012 (4E-4) 0.035 0.088 3 5 4 Parathion 0.013 (4E-4) 0.010 0.022 3.5 2 2 Mevinphos 0.015 (4E-4) 0.020 -- 5 4 -- Fenamiphos 0.055 (2E-3) 0.040 0.098 6 6 5 Methidathion 0.080 (3E-3) 0.100 0.108 7 9 6 Fensulfothion 0.350 (7E-3) 0.011 -- 8 3 -- Monocrotophos 0.720 (0.02) 0.054 0.088 9 7 3 Omethoate 4.47 (0.11) 0.213 0.116 10 12 7 Ethephon 6.22 (0.20) 35.7 25.8 11.5 14 13 Methamidophos 6.46 (0.20) 0.081 0.130 11.5 8 8 Glyphosate 9.02 (0.39) 43.7 12.1 13 15 12 Demeton-S-methylsulfone 12.8 (0.26) 0.187 0.149 14 11 9 Dimethoate 42.2 (1.03) 0.397 0.341 15 13 10 a Values from RTECS Database -- Value not available Number of Comparisons 15 13 Spearman’s Rank Order Correlation Coefficient (ρ) 0.709 0.571 Significant ρ (N=15, α=0.01) 0.689 Significant ρ (N=13, α=0.05) 0.566 80

Toxicity Comparison by Chemical

100.0000

10.0000

50 1.0000 D L r 0.1000

50 o 0.0100 C E 0.0010

0.0001

s n n n s s n n s te n s e e e o o io io o o io io o a o o t n t rv x h h h h h h h h h a fo a o o t t p ip t t p o p p s l o l a a a in a fo o th e o o u th h r r r v m id l t e h id h ls e ic a a a e a h u ro t p y D P P P n t s c m E m ly h im l l M e e n o O a G t D y y F M e n th e th th F o e m E e M - M -S M n to e m e Compound D

C. elegans Rat Mouse

Figure 3.1. Comparison of toxicity values for C. elegans (behavioral EC50s) to rats and mice

(LD50s). 81

ChE Activity Comparison

18 n) i 16 m n/

ei 14 ot

pr 12

g Control m

/ 10 l

o Mid Exposure

m 8

n High Exposure ( y

t 6 i v i t

c 4 A

E 2 h C 0

s n n n n te n e o io io io io a o n rv h h h h s h fo lo t t t t o p l h ra ra a fo h e su c a a id ul p th l i P P th s ly E hy D yl e n G t h M e e et F -m M S n- o et em D

Figure 3.2. Graphical comparison of ChE activity levels in nmols activity/mg protein/minute.

82

Effect of pH on Movement

125 ) l o r

100 t n o C 75 % ( t 50 en

25 vem o M 0 6 5 4 3 2 1 0 pH

Rep 1 Rep 2 Rep 3

Figure 3.3. Response of movement behavior to decreasing exposure solution pH.

83

ChE Activity Comparison

120 )

l 100 o r t n

o 80 C Control

(% 60 Mid Exposure ty i

v High Exposure ti

c 40 A E h

C 20

0

s n n n n te n e o o o o o a o n v i i i i s h o r th th th th p lf lo a a a o o e u h r r d lf h h s c a a i u p t l i P P th s ly E y D l e n G th y e e th M e F -m M -S n to e m e D

Figure 3.4. Graphical comparison of ChE activity levels expressed as % control activity. 84

Oxidative Desulfuration

CYP 450

Parathion Paraoxon

Figure 3.5. Oxidative desulfuration of the parathion to the active metabolite paraoxon. 85

pH versus Movement

125 ) l o r 100 t n o

75 C % ( t 50

25 vemen o M 0 6 5 4 3 2 1 pH

pH Composite Ethephon Glyphosate Monocrotophos Methamidophos Fenamiphos

Figure 3.6. Organophosphate induced change in movement versus exposure solution pH.

The two data points for each chemical represent the low and high concentrations used in behavioral toxicity exposures for the compounds shown.

86

CHAPTER 4

CONCLUSIONS

The purpose of this thesis was to explore C. elegans as a model for neurotoxicity in mammals. This question was addressed by choosing a class of known neurotoxic chemicals, conducting toxicity tests for these compounds using C. elegans, and comparing this to toxicity values from the literature for testing in rats and mice. A second goal of this study was to confirm cholinesterase inhibition resulting from organophosphate exposure as a toxic effect in C. elegans. After conducting the described set of experiments and analyzing the resulting data a number of conclusions can be drawn, and a number of outstanding questions are raised.

The data presented in this thesis support the use of C. elegans as a biological model for neurotoxicity in higher animals, including mammals, where there is an anatomical and physiological basis for such a comparison. Similarities between the cholinergic nervous systems of C. elegans and mammals provide the basis for such a comparison with respect to organophosphate toxicity. Toxicity comparisons between C. elegans and rats and C. elegans and mice exhibited statistically significant correlation in rank order of toxicity between 4-hour behavioral EC50s in C. elegans and mammalian

LD50 values recorded in the literature. Future experiments should consider other OPs, other anticholinesterase chemicals (e.g. carbamates), and other organic chemicals known to be exert neurotoxic effects through other shared portions of the nervous system. 87

It can be concluded from this work that organophosphate chemicals display a wide range of potency in C. elegans much as they do in rats and mice. Toxicity values in all species compared spanned a range of 3.5 to 4.5 orders of magnitude. Further testing with more OPs could further characterize the range of OP toxicity in C. elegans.

This work confirms 4-hour aquatic exposures as useful exposure regimes in toxicity testing with C. elegans for compounds which show sensitivity at concentrations below their solubility limits. The use of an organic vehicle in aquatic exposures of C. elegans to hydrophobic toxicants is an area that needs to be further explored.

Cholinesterase assays conducted as part of this work confirm statistically significant depression in C. elegans cholinesterase activity following exposure to OP agents. Almost all significant declines in activity were more than 20% lower than control values. We conclude that these decrements are a primary cause of the observed corresponding decrease in movement. It would be very informative to conduct experiments more rigorously examining concentration-response relationship of ChE activity to OP exposure. This would be especially useful if the same concentrations could be tested for their behavioral effect allowing examination of the relationship between ChE activity and movement behavior.

A limited number of concentration-response experiments quantifying ingestion behavior as the toxic endpoint were carried out in association with this investigation.

The method used for these analyses entailed counting the number of fluorescent microspheres ingested by worms as described in Boyd et al. (In press). These experiments were not conducted on a sufficient number of chemicals to be included in this thesis, but what results were obtained appeared to parallel the corresponding 88

locomotor response. A more in-depth characterization of the ingestion response to OPs,

and its relationship to locomotion and ChE activity dynamics may be worth pursuing in

order as a different behavioral endpoint.

Cholinesterase assays also displayed that not all organophosphate compounds

decrease locomotor activity through AChE inhibition. This conclusion is in keeping with

the OP literature. An OP induced reduction in pH was shown to be the primary cause in

displayed glyphosate and ethephon toxicity. Unexpected differences in potency for some

OP compounds (dichlorvos and demeton-S-methylsulfone) reported to be anticholinesterase in mammals but which failed to show significant anticholinesterase activity in C. elegans raise questions as to the metabolic avenues open to nematodes.

Further exploration of metabolic and mechanistic possibilities in C. elegans is needed.

An important conclusion emerging from the cholinesterase assay portion of this

work is the confirmation that C. elegans is capable of carrying out the oxidative

desulfuration necessary to bioactivate phosphorothioates and phosphorodithioates. Four

OPs requiring bioactivation were shown to cause significant depression in ChE activity.

Highly similar EC50s for parathion and its active metabolite paraoxon in behavioral tests

also imply that bioactivation of some compounds in C. elegans is rapid enough to have

little bearing on overall toxicity. Experiments to determine which nematode enzymes are

responsible for the desulfuration would be a useful place to begin investigating OP

metabolism in C. elegans.

89

References

Boyd, W.A., Cole, R.D., Anderson, G.L., and Williams, P.L. The effects of metals and food availability on the behavior of Caenorhabditis elegans. Environ. Toxicol. Chem.

Accepted for publication, 2003.