Historia Venom

background image

Review

Pharmacology and biochemistry of spider venoms

Lachlan D. Rash, Wayne C. Hodgson*

Monash Venom Group, Department of Pharmacology, PO Box 13E, Monash University, Victoria 3800, Australia

Abstract

Spider venoms represent an incredible source of biologically active substances which selectively target a variety of vital

physiological functions in both insects and mammals. Many toxins isolated from spider venoms have been invaluable in helping

to determine the role and diversity of neuronal ion channels and the process of exocytosis. In addition, there is enormous

potential for the use of insect speci®c toxins from animal sources in agriculture. For these reasons, the past 15±20 years has seen

a dramatic increase in studies on the venoms of many animals, particularly scorpions and spiders. This review covers the

pharmacological and biochemical activities of spider venoms and the nature of the active components. In particular, it focuses

on the wide variety of ion channel toxins, novel non-neurotoxic peptide toxins, enzymes and low molecular weight compounds

that have been isolated. It also discusses the intraspeci®c sex differences in given species of spiders. q 2001 Elsevier Science

Ltd. All rights reserved.

Keywords: Spider venom; Neurotoxins; Ion channels; Enzymes; Necrotic arachnidism; Sex differences

Contents
1. Spiders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

2. Spider venoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

3. Venom components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

3.1. Neurotoxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226

3.2. Glutamate receptor toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227

3.3. Calcium channel toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

3.4. Sodium channel toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

3.5. Potassium channel toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233

3.6. Chloride channel toxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234

3.7. Toxins that stimulate the release of neurotransmitters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

3.8. Toxins affecting cholinergic transmission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235

3.9. Other neurotoxins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236

4. Non neurotoxic peptides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236

5. Enzymes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238

6. Low molecular weight components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240

6.1. Biogenic amines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240

6.2. Free amino acids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

6.3. Other low MW components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

7. Necrotic arachnidism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

7.1. Necrotic arachnidism in Australia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242

8. Sex-linked variation in venom composition and activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

9. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245

Toxicon 40 (2002) 225±254

0041-0101/01/$ - see front matter q 2001 Elsevier Science Ltd. All rights reserved.

PII: S0041-0101(01)00199-4

www.elsevier.com/locate/toxicon

* Corresponding author. Tel.: 161-3-99054861; fax: 161-3-99055851.

E-mail address: wayne.hodgson@med.monash.edu.au (W.C. Hodgson).

background image

1. Spiders

Spiders and humans have been sharing habitats for as

long as humans have existed. In that time spiders have

evoked numerous and varied responses from people includ-

ing fascination, wariness and, as any bona ®de arachno-

phobe will attest to, debilitating terror. They have secured

a place in the mythology of many cultures throughout the

world as well as in modern popular culture, from nursery

rhymes to horror movies.

Spiders are an ancient group with the oldest known fossil

records dating 300 million years, from the Carboniferous

period. Excluding insects, spiders are the most successful

invertebrates on land, with nearly 40,000 species having

been described world-wide (Platnick, 1993). Spiders can

be found in nearly every terrestrial habitat from seashores

to alpine regions above the tree line, from deserts to tropical

rain forests, and are found in all levels of these environ-

ments. Many have readily adapted to living in close associa-

tion with man. Most spiders are not naturally aggressive and

are generally harmless to humans. However, several species

of spider are responsible for harmful and even fatal bites.

When spider bites do occur they are usually accidental

and/or defensive, such as when a spider hiding in clothing

or bed linen is trapped against the skin.

There are two main groups of spiders: the Orthognatha or

mygalomorph (primitive or trapdoor spiders), whose cheli-

cera project forward from the cephalothorax, while the fangs

move downwards; and the Labidognatha or araneomorphs,

whose chelicera are positioned vertically, and together with

the fangs move laterally like pincers. Most spiders feed

predominantly on insects and other arthropods with some

larger species readily catching and eating frogs, lizards,

snakes, small birds and rodents. Prey is either ambushed

or tackled and grasped with the fangs (hunting spiders) or

caught in silk snares (web builders). Whatever the method of

capture the ultimate demise of the prey is often brought

about by the injection of venom. Once the prey is secured

all spiders begin digestion externally, regurgitating diges-

tive ¯uids onto or into the prey before swallowing the liquid

meal using their strong sucking stomach.

2. Spider venoms

Given the age of spiders as an evolutionary group and the

minimal changes in their morphology over time, it is possi-

ble that the use of venom was developed very early (Bettini

and Brignoli, 1978). The primary purpose of spider venom

is to paralyse or kill prey, and it may also play a role in

pre-digestion of the intended meal. It may be that a spiders

venom also serves as a self defence mechanism against

predators. Indeed, it has been suggested that some spiders

do not inject venom when subduing prey but only defen-

sively (Minton, 1974). Whatever its purpose the use of

venom has been integral in the successful evolution of

spiders, possibly more so than their use of silk (Bettini

and Brignoli, 1978).

In general, spider venoms, especially those from tarantu-

las, are clear, colourless liquids that are easily soluble in

water (Bucherl, 1971; Norment and Foil, 1979; Savel-

Neimann, 1989). Minton (1974) states that most spider

venoms are neutral or alkaline, although some are acidic.

The latter is the case with the mygalomorphs Atrax robustus,

Dugesiella hentzi and Eurypelma californicum, all of

which have venoms with pHs of 4±5.5 (Wiener, 1961;

Schanbacher et al., 1973a; Savel-Neimann, 1989). Upon

lyophilisation, spider venoms have been reported as ranging

in colour from whitish to whitish-pink, -yellow or -grey

(Wiener, 1961; Bucherl, 1971). The dried venom of true

spiders (araneomorphs) is reportedly stable for years in

vacuum while mygalomorph venoms are more hygroscopic

and will liquefy if left in contact with air, but will keep for a

few months in vacuo (Bucherl, 1971).

3. Venom components

Like the venoms of other animals, such as snakes and

scorpions, venoms from spiders are heterogeneous, not

only between species but also within species. They are

made up of complex mixtures of biologically active and

inactive substances. The major constituents of spider

venoms are protein, polypeptide and polyamine neurotox-

ins, enzymes, nucleic acids, free amino acids, monoamines

and inorganic salts (Jackson and Parks, 1989).

3.1. Neurotoxins

As the primary purpose of their venom is to paralyse prey,

spiders produce a variety of toxins which affect the nervous

system. Most, if not all, spider neurotoxins characterised to

date have been found to be either proteins, peptides or acyl-

polyamines (McCormick and Meinwald, 1993), and have a

variety of actions throughout the nervous system. Two

important characteristics of nervous tissue are the excitabil-

ity of the cell membrane and the ability to transmit the

electrical signal across a synapse. The majority of spider

toxins are therefore targeted to either neuronal receptors,

neuronal ion channels or presynaptic membrane proteins

involved in neurotransmitter release. Thus neurotoxins

isolated from spider venoms can be classi®ed according to

their mode of action: toxins affecting glutamatergic trans-

mission; calcium (Ca

21

), sodium (Na

1

), potassium (K

1

)

and chloride (Cl

2

) channel toxins; toxins that stimulate

transmitter release and toxins blocking postsynaptic choli-

nergic receptors. Following is a description of the normal

physiology of these target sites and the known spider toxins

affecting them.

The structure of neurotoxins isolated so far and their

molecular evolution are beyond the scope of this review

and have recently been reviewed by Escoubas et al. (2000).

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

226

background image

3.2. Glutamate receptor toxins

L-Glutamate (glutamic acid) is the principal excitatory

transmitter in the mammalian central nervous system

(CNS) (Greenamyre and Porter, 1994), the main sources

being glucose, via the Krebs cycle, and glutamine

synthesised in glial cells. It is stored in synaptic vesicles

and released by Ca

21

-dependent exocytosis. The action of

glutamate is terminated by its removal from the synapse via

high af®nity transporters located on both neurones and glial

cells. There are several types of glutamate receptor broadly

classi®ed as the ionotropic and metabotropic glutamate

receptors, iGluR and mGluR respectively. The ionotropic

receptors are further divided into NMDA (N-methyl-D-

aspartate) and non-NMDA (consisting of AMPA

(a-amino-3-hydroxy-5-methyl-isoxazole)

and

kainate)

subtypes, being named after their sensitivity to selective

agonists. The ionotropic receptors are ligand-gated ion

channels that are cation (K

1

, Na

1

and Ca

21

) selective and

are oligomeric assemblies of either four or ®ve subunits

each with three transmembrane domains (the exact number

of subunits making up functional receptors is a matter of

intense debate and is reviewed by Dingledine et al., 1999).

The NMDA receptor consists of NR1 and NR2 subunits, has

slow kinetics, is highly permeable to Ca

21

and has a modu-

latory site sensitive to the endogenous polyamines spermine

and spermidine, the presence of which potentiates the

response to glutamate. AMPA and kainate receptors are

made up of the GluR1-7 and KA1-2 types of subunit, are

both sensitive to the amino acid quisqualic acid, have fast

kinetics and low permeability to Ca

21

. NMDA and AMPA

receptors are generally co-localised and widely distributed

whereas kainate receptors have a more restricted

distribution. The metabotropic receptors are G-protein-

coupled receptors linked to either inositol (1,4,5) tris-

phosphate (IP

3

) formation and intracellular Ca

21

release or

inhibition of adenylate cyclase and are found both pre- and

post-synaptically.

Glutamate is also an important transmitter in the CNS

and the main transmitter in the neuromuscular junction of

invertebrates (Osborne, 1996) with both an excitatory and

inhibitory function. There are several types of glutamate

receptors recognised in invertebrates. The excitatory, or

depolarising, D-response is mediated by a cation (Na

1

)

selective channel with fast kinetics that is sensitive to quis-

qualate, being similar to the mammalian AMPA receptor.

The inhibitory, or hyperpolarising, H-response is mediated

by anion (Cl

2

) selective channels with a slow prolonged

action and no known mammalian counterpart. Both

receptors are found in the CNS and neuromuscular junc-

tion. The ®rst description of an NMDA receptor in an

invertebrate, found in cray®sh visual interneurons, was

reported by Pfeiffer-Linn and Glantz (1991), suggesting a

third type of invertebrate glutamate receptor. Given the

importance of glutamate as a transmitter in invertebrates,

particularly its involvement in locomotion, it is not surpris-

ing that spiders possess toxins that target glutamatergic

transmission.

In the early 1980s, a toxin (JSTX, Joro spider toxin) was

isolated from the venom of the orb weaving spider Nephila

clavata. In the lobster neuromuscular junction, JSTX

irreversibly suppressed excitatory post synaptic potentials

(EPSP) and depolarisation elicited by exogenous glutamate

without affecting inhibitory postsynaptic potentials or

depolarisation due to aspartate (Kawai et al., 1982a).

JSTX was also found to block glutamatergic synapses in

mammalian brain (Kawai et al., 1982b). Similar activity

was found in venom gland extracts from the spider Araneus

ventricosus (Kawai et al., 1983). Shortly after, JSTX and

another glutamate receptor blocker NSTX (Nephila spider

toxin from the spider Nephila maculata) were chemically

characterised (Aramaki et al., 1986). At about the same time

Usmanov et al. (1983) found that venom from Argiope

lobata irreversibly blocked responses of locust muscle to

exogenous glutamate. A 636 Da toxin (named argiopine)

(Fig. 1) was subsequently isolated and found to be closely

related to JSTX and NSTX (Grishin et al., 1986).

Since then, several glutamate receptor antagonists (called

argiotoxins) have been isolated from the venoms of the

North American orb weaving spiders Argiope trifasciata,

Argiope ¯orida and Araneus gemma, one of which (Arg

636; Fig. 1) was subsequently found to be identical to

argiopine (Usherwood et al., 1984). Usmanov et al. (1985)

found that postsynaptic inhibition of glutamatergic trans-

mission at the locust neuromuscular junction was a common

property of eight other spiders belonging to the family

Araneidae (orb weavers).

The chemical structures of many of these toxins have now

been elucidated and they form a family of closely related

compounds, the acylpolyamines, or polyamine amides

(Fig. 1). These toxins contain an aromatic acyl end group

and a polyamine chain (McCormick and Meinwald, 1993).

To date, in all polyamine amide toxins isolated from

Araneidae spiders, the acyl head group is connected to the

polyamine tail by one or two of the amino acids asparagine,

ornithine and v-methyllysine (McCormick and Meinwald,

1993; Schulz, 1997). In general the amino acid containing

polyamine amides appear to have a similar action at inver-

tebrate neuromuscular junctions, causing an open channel

(use-dependent), irreversible or slowly reversible, non-

competitive block of quisqualate sensitive glutamate recep-

tors (Jackson and Parks, 1989). However, there have been

con¯icting reports on their selectivity and speci®city for

vertebrate glutamate receptors. JSTX-3 (Fig. 1) has been

reported to be selective for non-NMDA ionotropic receptors

(Kawai et al., 1991) whereas argiotoxin-636 has been found

to preferentially inhibit both NMDA (Priestly et al., 1989)

and non-NMDA (Ashe et al., 1989) receptors. Nemeth et al.

(1992) found that the argiotoxins and a-agatoxins (see

below) are speci®c inhibitors of NMDA receptors in rat

brain, whereas JSTX was an equipotent inhibitor of

NMDA- and kainate-evoked [Ca

21

]

i

increase in rat

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

227

background image

cerebellar granule cells. In contrast, argiotoxin-636 equally

inhibited responses to both NMDA and kainate in Xenopus

oocytes injected with total rat brain RNA (Usherwood et al.,

1992). More recent work suggests that the sensitivity of

AMPA receptors to argiotoxins is subunit dependent

(Herlitze et al., 1993; reviewed by Usherwood and

Blagbrough, 1994).

A second class of acylpolyamine toxins has been

identi®ed in the venom of a variety of spiders outside the

family of orb weavers. Non-amino acid containing

acylpolyamines have been isolated from the funnel-web

spiders Agelenopsis aperta and Hololena curta, a trapdoor

spider (Hebestatis theveniti), and two tarantulas,

Harpactirella sp. and Aphonopelma chalcodes. The

a-agatoxins, isolated from A. aperta, are selective and

reversible, non-competitive antagonists of NMDA recep-

tors in mammalian brain (Parks et al., 1991). In the chick

CNS, venom from Hololena curta irreversibly blocked

non-NMDA mediated transmission, the responsible toxin

reported as having a molecular weight of 5±10 kDa (Jack-

son et al., 1987). Ten paralytic acylpolyamines and one

insecticidal peptide (collectively known as curtatoxins)

were subsequently isolated from the venom of H. curta

(Quistad et al., 1991), two of the acylpolyamines (HO

489

and HO

505

; Fig. 1) and the peptide (m-AgaII) having

previously been described in the venom of A. aperta. The

polyamine toxins isolated from Hebestatis theveniti

(Hettoxins) were paralytic to lepidopteran larvae.

However, the mechanism of action was not investigated

(Skinner et al., 1990). The structures and pharmacology

of the acylpolyamine spider toxins have been thoroughly

reviewed by Usherwood and Blagbrough (1991);

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

228

Fig. 1. Generalised structure of polyamine amide spider toxins (modi®ed from Usherwood and Blagbrough, 1991) and structures of several key

polyamine toxins from spider venoms.

background image

McCormick and Meinwald (1993). In addition, Hisada et

al. (1998) recently outlined the generalised structures of

toxins (types A±D) isolated from the Nephila spiders in a

paper reporting the identi®cation of a ®fth structure type

(type-E).

In addition to glutamate receptor antagonists, a new class

of toxin affecting glutamatergic transmission was recently

reported (Mafra et al., 1999). PhTx-4, a family of seven

individual toxins, was isolated from the venom of the

Brazilian armed spider (Phoneutria nigriventer) and found

to inhibit glutamate uptake in rat cerebrocortical synapto-

somes in a dose-dependent manner. One of these toxins

Tx4(5±5) with potent insecticidal activity has subsequently

been found to reversibly antagonise the NMDA receptor

while having little effect on AMPA, kainate or GABA

induced currents in rat hippocampal neurons (Figueiredo

et al., 2001).

3.3. Calcium channel toxins (Table 1)

Calcium channels are present in many tissues throughout

the body, and are of particular importance in the nervous

system where the entry of Ca

21

into the nerve terminal plays

a key role in the release of neurotransmitters. In addition to

voltage-activated calcium channels (the focus of this

section), the permeability of membranes to calcium is

increased via ligand-gated ion channels such as the

NMDA receptor (discussed in Section 3.2) and by a family

of intracellular calcium release channels referred to as

ryanodine receptors (Sutko et al., 1997). This family of

receptors are involved in the release of intracellular Ca

21

from the sarcoplasmic (muscle) or rough endoplasmic

reticulum (non-muscle).

There are currently six recognised subtypes of voltage-

activated Ca

21

channel (T-, L-, N-, P-, Q- and R-types)

based on their electrophysiological properties and sensitiv-

ity to various activators/inhibitors and ions (Uchitel, 1997).

However, in some cells it is dif®cult to separate the P and Q

components of a current therefore the term `P/Q-type'

current is often used when referring to either one (Alexander

and Peters, 1999). T-type channels are activated by small

depolarisations (low-voltage activated) and undergo fast

inactivation but slow deactivation and are involved in the

generation of pacemaker activity in neurons and cardiac

muscle. L- and N-type channels are characterised by high-

voltage activation and fast deactivation. L-type channels

inactivate slowly, are blocked by dihydropyridines and are

found in cardiac and vascular smooth muscle, neuronal cell

bodies and proximal dendrites of many central neurons. In

contrast, N-type (neuronal) channels have a moderate rate of

inactivation, are characterised by their sensitivity to

v-conotoxin GVIA (from the venom of the cone snail

Conus geographus) and mediate neurotransmitter release

at synaptic endings. P- and Q-type channels are moderate

voltage-activated Ca

21

channels with fast deactivation, are

found in central and peripheral neurons and are also

involved in transmitter release. Q-type channels have a

moderate rate of inactivation while P-type channels are

non-inactivating. The most recent addition to the voltage-

activated Ca

21

channels is the R-type, which is also acti-

vated at moderate voltages but undergoes fast inactivation

and deactivation and is resistant to known pharmacological

agents. Mammalian voltage-gated calcium channels are

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

229

Table 1

Spider toxins affecting calcium channels

Toxin

Spider

Channel selectivity

Reference/s

PLTXs I±III

P. tristes

Non-selective Drosophila Ca

21

currents

Branton et al., (1987); Leung et al., (1989)

Hololena toxin

H. curta

Non-inactivating Drosophila Ca

21

currents Bowers et al., (1987); Leung et al., (1989)

v-Aga IA-IC

A. aperta

L

Bindokas and Adams, (1989); Venema et al., (1992)

v-Aga IIA and IIB

A. aperta

N

Adams et al., (1990); Venema et al., (1992)

v-Aga IIIA

A. aperta

L and N (P)

Mintz et al., (1991); (1992a); Ertel et al., (1994)

v-Aga IIIB and IIID A. aperta

N (L and P)

Ertel et al., (1994)

v-Aga IVA and IVB A. aperta

P/Q

Mintz et al., (1992b); Adams et al., (1993)

FTX

A. aperta

P

Llinas et al., (1989); (1992)

v-GsTx-SIA

G. spatulata

P, N and Q

Lampe et al., (1993); Piser et al., (1995)

PhTx3); 3-3 and 3-4 P. nigriventer

P/Q

Guatimosin et al., (1997); Miranda et al., (1998)

PhTx3); 3-2 and 3-5 P. nigriventer

L

LeaÄo et al., (1997)

CNS2103

D. okefenokiensis L and R

Kobayashi et al., (1992)

CSTX1

C. salei

L

Cruz et al. in Kuhn-Nentwig and Nentwig, (1997)

SNX-325

S. ¯orentina

N-type selective at nM []s

Newcomb et al., (1995)

Agelenin

A. opulenta

P

Hagiwara et al., (1991); (1999)

DW13.3

F. hibernalis

Non selective

Sutton et al., (1998a)

SNX-482

H. gigas

E

Newcomb et al., (1998)

v-atracotoxin

H. versuta

Insect VACCs

Wang et al., (1999)

HWTX-1

S. huwena

N

Peng et al., (2001)

Hanatoxin 1 and 2

G. spatulata

rabbit brain a

1A

subunit (non-L-type)

Li-Smerin and Swartz, (1998)

background image

made up of several subunits; a1, a2, b, d. Skeletal muscle

channels also contain a g subunit. The a1 subunit of calcium

channels forms the calcium conducting pore. Several a1

subunits have now been cloned from a variety of inverte-

brates and show high structural homology with their verte-

brate counterparts (i.e. the same four repeat structure, each

containing six transmembrane domains) (Hall et al., 1994).

The insect a1 subunits characterised so far share some

sequence identity and pharmacology with the mammalian

L-type channels, for example their sensitivity to phenylalk-

ylamines (Zheng et al., 1995; Lee et al., 1997; MacPherson

et al., 2001). However, as with insect sodium channels

(discussed in Section 3.4), there are suf®cient structural

differences for insecticidal compounds to discriminate

between phyla (Hall et al., 1994).

The presence of Ca

21

channel blocking toxins in spider

venoms was suggested with the discovery that the venoms

of three spiders, including Agelenopsis aperta, caused an

irreversible presynaptic block at the Drosophila neuromus-

cular junction (Branton et al., 1987). Several peptide toxins

were isolated from the spider Plecteurys tristes (PLTXs

I±III, MW 6±7 kDa) (Branton et al., 1987) and one from

Hololena curta (Hololena toxin, MW 16 kDa) (Bowers et

al., 1987) the activities of which were consistent with long

lasting speci®c block of presynaptic Ca

21

channels. This

mode of action was con®rmed when Leung et al. (1989)

found that Hololena toxin selectively blocked non-inactivat-

ing Ca

21

current while PLTX±II blocked both inactivating

and non-inactivating Ca

21

currents in cultured Drosophila

neurones. The complete primary structure of PLTX±II has

been determined revealing a unique post-translational

modi®cation, palmitoylation of the C-terminal threonine

(Branton et al., 1992).

Further work on venom from A. aperta resulted in the

discovery of the v-agatoxins, a family of unrelated peptide

toxins with differing selectivity for neuronal Ca

21

channels

(reviewed by Olivera et al., 1994; Uchitel, 1997). Type I

agatoxins (v-AgaIA, IB and IC) are heterodimeric peptides

of similar size (MW 7.5 kDa) with ®ve disulphide bonds.

v-AgaIA potently blocks voltage sensitive Ca

21

channels in

insects, blocks neurotransmission at the frog neuromuscular

junction (Bindokas and Adams, 1989) and suppresses high

threshold Ca

21

currents in rat dorsal root ganglion cells

(Scott et al., 1990) thereby affecting Ca

21

channels from

both vertebrates and invertebrates. However, v-AgaIA and

IB failed to inhibit binding of v-conotoxin GVIA to chick

synaptosomal membranes (Adams et al., 1990). In addition

to v-AgaIA having no effect on potassium-stimulated Ca

21

entry into chick synaptosomes (Venema et al., 1992), this

suggests that the type-I agatoxins have no activity at N-type

high voltage-activated Ca

21

channels making them

relatively selective for L-type channels.

Type II v-agatoxins (IIA and IIB, ~10 kDa), which have

limited homology with type I toxins (only 43% at NH

2

terminal end), also cause presynaptic block of insect neuro-

muscular transmission. Unlike type I v-agatoxins, type II

v-agatoxins inhibit the binding of v-conotoxin GVIA to

chick synaptosomes (Adams et al., 1990) and potently

block the potassium-stimulated Ca

21

entry into chick

synaptosomes (Venema et al., 1992) hence their classi®ca-

tion as N-type channel blockers. v-Agatoxin IIIA, an

8.5 kDa peptide, was found to be an equipotent blocker of

L- and N-type Ca

21

channels (Mintz et al., 1991), also

causing a partial (~40%) voltage-dependent block of

P-type channels (Mintz et al., 1992a). During the puri®ca-

tion of v-AgaIIIA, several homologous peptides

(v-AgaIIIB, IIIC and IIID) were discovered (Ertel et al.,

1994). v-AgaIIIB and IIID have similar actions to IIIA

with no activity at the insect neuromuscular junction, no

effect on atrial type-T channels and are equipotent inhibitors

of v-conotoxin GVIA binding to rat synaptic membranes.

On the contrary, v-AgaIIIA is a more potent blocker of

locust Ca

21

channels and 100-fold more potent at blocking

L-type channels, making v-AgaIIIB and IIID 100-fold more

selective for N-type Ca

21

channels. Finally, v-AgaIVA, a

48 amino acid peptide unrelated to v-agatoxins I±III, is a

potent blocker of P/Q-type Ca

21

channels (K

d

of ~2 nM for

P-type and ~90 nM for Q-type) in mammalian systems,

however has no effect on T-, L- or N-type channels

(Mintz et al., 1992b). Although having similar potency at

P-type channels (K

d

3 nM) v-AgaIVB, which shares 71%

homology with IVA, is eight times slower in blocking the

current (Adams et al., 1993).

In addition to the peptide v-agatoxins, a low molecular

weight (200±400 Da) fraction of A. aperta venom blocks

Ca

21

channels (now known as P-type after the tissue in

which they were ®rst described) in guinea-pig cerebellar

Purkinje cells, in squid giant synapse and in Xenopus

oocytes injected with rat brain mRNA (Llinas et al., 1989;

Lin et al., 1990). The active toxin, named FTX, was identi-

®ed as a polyamine (Llinas et al., 1992) and a structure

proposed but not con®rmed. A toxin (FTX-3.3; Fig. 1)

claimed to have the identical structure as FTX, along with

an amide-containing analogue (sFTX-3.3), have been

synthesised (Blagbrough and Moya, 1994; Moya and

Blaghbrough, 1994) and shown to block P-type Ca

21

chan-

nels in rat cerebellar Purkinje cells (Dupere et al., 1996).

v-Grammotoxin SIA (v-GsTx-SIA) is a 36 amino acid

peptide (MW 4.1 kDa) with three disulphide bridges

isolated from the venom of the South American Rose taran-

tula, Grammostola spatulata. v-GsTx-SIA produced

concentration-dependent inhibition of K

1

-evoked

45

Ca

21

in¯ux into rat and chick brain synaptosomes suggesting

that it is a blocker of presynaptic N- and P-type voltage-

activated channels in vertebrates (Lampe et al., 1993). Piser

et al. (1995) con®rmed the action of v-GsTx SIA at N- and

P-type channels, in rat cultured hippocampal neurons, and

found that it also blocked Q-type but not L-type Ca

21

chan-

nels. Using rat cerebellar Purkinje neurons, as well as rat and

frog sympathetic neurons, McDonough et al. (1997)

concluded that v-grammotoxin potently inhibits P- and

N-type channels by impeding channel gating and that it

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

230

background image

binds to different or additional sites than v-AgaIVA on

P-type channels. In addition to a relatively non-speci®c

action at voltage-activated Ca

21

channels, v-GsTx SIA

has subsequently been shown to interact with and modify

the gating of voltage-gated K

1

channels, raising the possi-

bility of a voltage-sensing domain common to voltage-gated

ion channels (Li-Smerin and Swartz, 1998).

Venom from the Brazilian `armed' spider contains

another family of peptides with Ca

21

channel blocking

activity. Three fractions toxic to mice, PhTx1, 2 and 3,

were isolated from P. nigriventer venom (Rezende et al.,

1991). Fraction PhTx3 was further separated into six neuro-

toxic peptides (Tx3-1±6) which induced varying types and

degrees of paralysis after intracerebroventricular (i.c.v.)

injection in mice (Cordeiro et al., 1993). Tx3-3 and Tx3-4

potently inhibited Ca

21

in¯ux into rat cerebrocortical synap-

tosomes (IC

50

s of 0.32 and 7.9 nM, respectively) via chan-

nels sensitive to v-AgaIVA but not v-conotoxin GVIA,

suggesting an action at P/Q-type Ca

21

channels (Guatimosin

et al., 1997; Miranda et al., 1998). Toxins Tx3-2 and Tx3-5

had no effect on T-type Ca

21

current, however irreversibly

inhibited L-type current by 80 and 100% respectively, in the

pituitary cell line GH3 (LeaÄo et al., 1997). Using cDNA

sequence analysis and patch clamp studies, Kalapothakis

et al. (1998a) con®rmed the previously published (Cordeiro

et al., 1993) amino acid sequence of Tx3-2 and its action on

L-type Ca

21

channels.

Many other Ca

21

channel toxins, with varying selectivity,

from a wide spectrum of spider species have been isolated in

the last 8 years. A polyamine Ca

21

channel antagonist

(CNS2103) isolated from the venom of the hunting spider

Dolomedes okefenokiensis reversibly blocks both L-type

and dihydropyridine resistant (R-type) channels in

N1E-115 neuroblastoma cells with no effect on T-type

currents or voltage-activated Na

1

or K

1

channels

(Kobayashi et al., 1992). Venom from another hunting

spider, Cupiennius salei, contains a 74 amino acid peptide

toxin (CSTX-1) which inhibits high-threshold Ca

21

chan-

nels (L-type) at glutamatergic synapses (unpublished obser-

vations of Cruz, J., Lacerda BeiraÄo, P.S., and LeaÄo, R.M.,

cited in Kuhn-Nentwig and Nentwig, 1997). Venom from

Segestria ¯orentina yielded the peptide toxin SNX-325,

which has no effect on Na

1

or K

1

channels but inhibits

most Ca

21

channels at micromolar concentrations.

However, at nanomolar concentrations SNX-325 is a selec-

tive blocker of N-type Ca

21

channels (Newcomb et al.,

1995).

Agelenin, isolated from the venom of Agelena opulenta,

is a 35 amino acid peptide containing six cysteine residues

which appears to block P-type Ca

21

channels (Hagawira et

al., 1990, 1991). DW13.3, a novel peptide from Filistata

hibernalis, causes potent blockade of most native Ca

21

channels with the exception of the T-type channel (Sutton

et al., 1998a). The 41 amino acid peptide SNX-482, from the

African tarantula Hysterocrates gigas, blocks human class E

Ca

21

channels (made up of a1E ion-conducting subunits

believed to be responsible for the characteristics of R-type

channels) expressed in a mammalian cell line. SNX-482

also inhibits a native R-type Ca

21

current in rat neurohypo-

physeal nerve terminals but has no effect on R-type currents

in several other types of rat central neurons (Newcomb et

al., 1998). A family of peptide neurotoxins (36±37 amino

acids in length), the v-atracotoxins, isolated from the Blue

Mountains funnel-web spider, Hadronyche versuta, display

selectivity for blocking insect voltage-gated Ca

21

channels

making them promising lead compounds for the develop-

ment of new pesticides (Wang et al., 1999).

HWTX-I is a 33 amino acid peptide with three disulphide

bonds isolated from the Chinese bird spider Selenocosmia

huwena (Liang et al., 1993; Zhang and Liang, 1993).

HWTX±I was found to be toxic to mice (i.p. LD

50

0.7 mg/

kg) and inhibited indirectly-evoked twitches in the mouse

phrenic nerve diaphragm preparation without affecting

directly-evoked twitches. In addition to the amino acid

sequence, the three-dimensional structure of HWTX±I

was determined using

1

H NMR (Quet al., 1995, 1997).

The action of HWTX±I on mouse phrenic nerve diaphragm

was found to be slowly reversible upon prolonged repeated

washing with the recovery time being decreased when

HWTX±I and tubocurarine were added together. It was

suggested that HWTX±I was a blocker of postsynaptic nico-

tinic receptors (Zhouet al., 1997). However, these ®ndings

were not supported by binding studies (Liang et al., 2000)

and subsequent experiments on three electrically stimulated

in vitro preparations suggested that the toxin acted presy-

naptically to inhibit the release of neurotransmitters (Liang

et al., 2000). Using patch-clamp studies, Peng et al. (2001)

recently found that HWTX±I potently and selectively inhi-

bits N-type high-voltage-activated Ca

21

channels in prosta-

glandin E1 differentiated NG108-15 cells. A second toxin,

huwentoxin-II, with neuromuscular blocking activity in the

mouse phrenic nerve diaphragm and reversible paralytic

activity in cockroaches, was recently isolated from

S. huwena, however the mechanism of action has not yet

been determined (Shuand Liang, 1999).

3.4. Sodium channel toxins (Table 2)

The movement of sodium ions across excitable cell

membranes, either via ligand-gated channels (eg. nicotinic

cholinoceptors and non-NMDA glutamate receptors) or

voltage-gated channels, is responsible for the rapid trans-

mission of impulses along and between excitable cells.

Voltage-gated sodium channels allow the co-ordination of

processes such as locomotion and cognition, and are often

the most abundant ion channels present in nerve and muscle

cells (Marban et al., 1998). This makes Na

1

channels prime

targets for paralytic neurotoxins. This section focuses on

spider toxins that target voltage-gated Na

1

channels while

toxins interacting with ligand-gated Na

1

channels are

discussed in Sections 3.2 and 3.8. Like most other

voltage-gated ion channels, Na

1

channels are made up of

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

231

background image

several subunits. The main a subunit, consisting of four

repeats each with six transmembrane domains, determines

the selectivity and properties of the channel and is the only

subunit required for function (Marban et al., 1998). At rest-

ing membrane potentials Na

1

channels are in a closed state,

activating or opening upon depolarisation of the membrane.

Once open, the channels undergo inactivation, or assume a

non-conducting state despite continued depolarisation of the

membrane, before deactivating and returning to a resting

state (Conley and Brammar, 1999). There are currently

nine recognised subtypes of mammalian Na

1

channel,

however there is no of®cial scheme for their classi®cation

(Alexander and Peters, 1999), with an unof®cial scheme

only recently proposed (Conley and Brammar, 1999;

Goldin, 1999). The majority of known Na

1

channels are

selectively and reversibly blocked by nanomolar concentra-

tions of tetrodotoxin (TTX) and saxitoxin (STX). There are,

however, several isoforms of voltage-gated Na

1

channels

that are relatively resistant to TTX, the cardiac Na

1

channel

NaH1 (Conley and Brammar, 1999) and two channels, SNS

and NaN, expressed in sensory neurons of rat, mouse and

human dorsal root ganglion (Did-Hajj et al., 1999). Studies

on insect Na

1

channels have revealed that they are structu-

rally, physiologically and pharmacologically very similar to

their vertebrate counterparts. However, there are also phar-

macological differences as evidenced by the isolation of

insect speci®c Na

1

channels toxins from the venoms of

several species of scorpion (reviewed by Zlotkin, 1999).

For the purpose of this review Na

1

channels will be broadly

classi®ed as either TTX-sensitive (TTX-S) or TTX-resistant

(TTX-R).

Three families of spider toxins are known to act at Na

1

channels (Adams et al., 1989; ArauÂjo et al., 1993a;

Nicholson et al., 1994, 1998), with the venom or toxins of

several other spiders displaying activity consistent with Na

1

channel interactions.

In conjunction with the discovery of the use-dependent

glutamate receptor antagonists the a-agatoxins, a second

class of peptide toxins was isolated from the venom of the

spider Agelenopsis aperta. The m-agatoxins (m-Aga-I±VI)

cause a slowly induced, irreversible, excitatory paralysis in

house¯ies. They were found to act presynaptically, inducing

repetitive ®ring of the nerve, an action blocked by TTX

(Adams et al., 1989). The solution structures of m-Aga-I

and -IV were elucidated and found to have the same overall

fold as many phylogenetically diverse peptide toxins selec-

tive for ion channels (Omecinsky et al., 1996). Fraction

PhTx2, isolated from the venom of P. nigriventer, is toxic

to mice causing excitatory symptoms including salivation,

lachrimation, priapism, convulsions, spastic paralysis of the

limbs and death upon i.c.v. injection (Rezende et al., 1991).

PhTx2 was subsequently shown to inhibit Na

1

channel inac-

tivation and shift the activation voltage to more negative

potentials in frog skeletal muscle (ArauÂjo et al., 1993a).

Nine peaks were obtained upon further separation of the frac-

tion PhTx2 (Tx2-1±9) of which toxins 2-1, 2-5, 2-6 and 2-9

were sequenced (Cordeiro et al., 1992). Toxins Tx2-5 and

Tx2-6 were found to have the same effects on Na

1

channel

inactivation and activation, voltage dependence and inactiva-

tion time constant as the whole fraction (ArauÂjo et al., 1993b).

The cDNAs of Tx2-1 and Tx2-5 were recently cloned

con®rming the previously published amino acid sequences

and identifying two new putative toxins Pn2-1A and Pn2-

5A (Kalapothakis et al., 1998b). Fraction PhTx2 has also

recently been found to cause progressive myonecrosis of

the mouse phrenic nerve diaphragm (Mattiello-Sverzuta and

Cruz-Hoȯing, 2000). The morphological changes induced by

PhTx2 include swelling of the sarcoplasmic reticulum, mito-

chondrial damage, disorganisation of sarcomeres, zones of

hypercontraction and rupture of the plasma membrane. The

authors suggest that the ability of PhTx2 to increase the

permeability of Na

1

channels can account for the myonecro-

sis as the morphological changes observed were similar to

those caused by osmotic disturbances.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

232

Table 2

Spider toxins affecting sodium channels

Toxin

Spider

Na

1

channel action

Reference/s

m-Aga-I±VI

A. aperta

Induce repetitive ®ring of nerves;

TTX sensitive (TTX-S)

Adams et al., (1989);

Omecinsky et al., (1996)

PhTx2; (Tx2-5 and 2-6)

P. nigriventer

inhibit inactivation

Araujo et al., (1993a)

Pn2-1A and Pn2-5A

P. nigriventer

putative Na

1

channel toxins

Kalapothakis et al., (1998b)

d-atracotoxin Hv1a

H. versuta

inhibit inactivation

Nicholson et al., (1994); (1996)

d-atracotoxin Ar1a

A. robustus

inhibit inactivation

Nicholson et al., (1996); (1998)

d-atracotoxin like

M. bradleyi

inhibit inactivation

Rash et al., (2000)

Curtatoxin II

H. curta

identical structure to m-Aga-II

Quistad et al., (1991)

Lyc IV

L. erythrognatha

probable Na

1

channel activity

Cruz et al., (1994)

DTX9.2

D. canities

TTX-S insect Na

1

channels

Bloomquist et al., (1996)

Arg-636/argiopine

Argiope sp.

inhibits Na

1

currents

Scott et al., (1998)

whole venom

Ancylometes sp.

" freq of MEPPs and depolarise

muscle ®bre membrane; TTX±S

GreÂgio et al., (1999)

PcTx1

P. cambridgei

blocks H

1

-gated Na

1

channels

Escoubas et al., (2000)

background image

Two mammalian selective toxins, robustoxin (RTX) and

versutoxin (VTX), from the Australian funnel-web spiders

Atrax robustus and Hadronyche versuta, respectively, are 42

amino acid peptides with four disulphide bonds which also

interact with Na

1

channels. In rat dorsal root ganglion cells

under voltage clamp conditions, VTX had no effect on TTX-

R Na

1

currents or K

1

currents. However, VTX did cause

dose-dependent slowing or removal of TTX-S Na

1

current

inactivation, reduced peak TTX-S Na

1

current and

increased the rate of recovery from inactivation (Nicholson

et al., 1994). VTX also shifted the voltage dependence of

Na

1

channel activation in the hyperpolarising direction with

both VTX and RTX causing partial activation of

22

Na

1

¯ux

as well as inhibiting batrachotoxin-activated

22

Na

1

¯ux in

rat DRG neurons (Nicholson et al., 1996). RTX was

subsequently found to have similar effects on Na

1

channel

activation and inactivation as VTX and the toxins renamed

d-atracotoxin-Ar1 and -Hv1a, respectively, re¯ecting the

subfamily (i.e. Atracinae) to which the spiders belong

(Nicholson et al., 1998).

Venom from the male Eastern mouse spider M. bradleyi

(Fig. 2), of Australia, appears to contain a neurotoxin that

facilitates neurotransmitter release in both smooth and

skeletal muscle preparations. The activity of the venom in

the chick biventer cervicis muscle preparation was inhibited

by A. robustus antivenom. In addition, experiments in rat

dorsal root ganglion cells under voltage clamp conditions

suggested that a component in the venom modi®es

TTX-sensitive sodium channel gating in a manner similar

to the d-atracotoxins (Rash et al., 2000a).

The curtatoxins (I±III) are peptide toxins from another

Agelenid spider (Hololena curta) which have been isolated,

sequenced and found to cause rapid, irreversible ¯accid

paralysis in crickets (Stapleton et al., 1990). Curtatoxin II

has an identical sequence to the peptide toxin m-Aga III

from the venom of A. aperta, however the mechanism of

action of the toxins from H. curta has not been con®rmed

(Quistad et al., 1991). The venom of the Brazilian wolf

spider Lycosa erythrognatha increased the duration of

compound action potentials and produced long-lasting

post-potentials in the bullfrog sciatic nerve, actions which

were essentially irreversible and attributed to a peptide (Lyc

IV) of approximately 8 kDa with a probable action at Na

1

channels (Cruz et al., 1994). An insecticidal peptide,

DTX9.2, from the venom of a weaving spider Diguetia

canities induced rapid paralysis in insects, caused hyperex-

citation of sensory nerve and neuromuscular preparations

from house¯y larvae, and depolarised the membrane of

cockroach giant axons. All the in vitro actions of DTX

were prevented or reversed by TTX suggesting an action

at voltage-sensitive Na

1

channels of insect nerve

membranes (Bloomquist et al., 1996). In addition to antag-

onising glutamate receptors, argiotoxin-636 was shown to

reduce the neuronal excitability of rat cultured DRG cells

partly by inhibiting voltage-activated Na

1

currents in a

manner dependent on repeated activation (Scott et al.,

1998). In the rat phrenic nerve diaphragm, venom from a

South American hunting spider (Ancylometes sp.), increased

the amplitude of indirectly evoked twitches, depolarised

muscle ®bre membranes and increased the frequency of

miniature endplate potentials. The depolarisation of muscle

®bre membranes was prevented by low Na

1

Tyrode solution

and TTX, which also inhibited the increase in miniature

endplate potentials suggesting the presence in the venom

of another toxin that activates voltage-sensitive Na

1

channels (GreÂgio et al., 1999).

A novel class of sodium channels appears to be blocked

by a recently isolated toxin (PcTX1) from the venom of the

tarantula Psalmopoeus cambridgei (Escoubas et al., 2000).

The 40-amino acid toxin potently and selectively blocks the

ASIC1a subclass proton-gated sodium channels (a member

of the Acid Sensing Ion Channel family) that are expressed

in sensory neurons from dorsal root ganglion and in the

central nervous system.

3.5. Potassium channel toxins (Table 3)

The largest and most diverse family of ion channels are

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

233

Fig. 2. The male (a) and female (b) Eastern Mouse spider (Missulena bradleyi). scale bar ˆ 2 cm.

background image

those selective for K

1

(Garcia et al., 1997), which is not

surprising given that K

1

channels regulate a wide variety of

cellular processes throughout the body. Potassium channels

are divided into two broad families, voltage-gated and

inwardly rectifying channels (Jan and Jan, 1994). Potassium

channels, like voltage-gated Na

1

and Ca

21

channels, are

made up of four repeats (in this case individual subunits),

each with six transmembrane domains, forming the ion pore

and several accessory subunits (Robertson, 1997). Voltage-

gated K

1

currents of native cells have been classi®ed as

either rapidly inactivating, transient outward `A-type' or

delayed rectifying, non-inactivating currents. Cloned

voltage-gated (K

V

) K

1

channel subunits are divided into

several subfamilies, those best characterised represent

mammalian counterparts of the Drosophila channels Shaker

(K

V

1), Shab (K

V

2), Shaw (K

V

3) and Shal (K

V

4) (Garcia et

al., 1997;Conley and Brammar, 1999). Although the genes

for many subunits have been cloned, there is still a substantial

gap in the correlation between cloned K

V

channels and native

currents (Robertson, 1997). However, the recent discovery of

toxins selective for K

1

channels from a variety of animal

venoms, including spiders, is helping to close this gap (Harvey

and Anderson, 1991; Garcia et al., 1994; Schweitz et al., 1995;

Sanguinetti et al., 1997; Diochot et al., 1999).

Two peptides, hanatoxins 1 and 2, isolated from the venom

of the Rose tarantula Grammostola spatulata (now Phrixotri-

cus spatulata), were shown to inhibit rat brain Shab- (Kv2.1)

and Shal-related K

1

channels expressed in Xenopus oocytes

whilst Shaker-, Shaw- and eag channels were relatively insen-

sitive (Swartz and Mackinnon, 1995). As mentioned

previously, the Ca

21

-channel blocker v-grammotoxin SIA,

from the same spider, was additionally found to interact

with voltage-gated K

1

channels. Likewise, the hanatoxins

were found to interact with voltage-sensitive Ca

21

channels

(Li-Smerin and Swartz, 1998). The K

V

4 family of channels are

also targeted by peptide toxins from two other spiders. In rat

ventricular myocytes, the heteropodatoxins (HpTx 1±3) from

Heteropoda venatoria prolonged the action potential duration

by blocking the transient outward K

1

current (I

to

) (Sanguinetti

et al., 1997). They were also found to block cloned K

V

4.2 but

not K

V

1.4 channels expressed in Xenopus oocytes and have

similar effects on channel gating as many of the Na

1

channel

toxins previously described, slowing current activation and

inactivation, as well as shifting the voltage-dependence of

inactivation to more positive potentials. The I

to

current in rat

ventricular myocytes and underlying channels, K

V

4.2 and

K

V

4.3, were also potently and selectively blocked by two

toxins (phrixotoxins, PaTx1 and PaTx2) from the venom of

the Chilean ®re tarantula Phrixotrichus auratus (Diochot et

al., 1999). Marvin et al. (1999) suggested that SGTx1, puri®ed

from the venom of the African tarantula Scodra griseipes, had

a similar mechanism of action to the hanatoxins after ®nding

the peptide partially (~40%) and reversibly inhibited both fast

transient and delayed recti®er K

1

currents in rat cerebellar

granule cells.

Using whole-cell patch clamping on GH3 cells,

Kushmerick et al. (1999) showed that toxin Tx3-1 from P.

nigriventer reversibly inhibits A-type K

1

currents without

affecting delayed rectifying, inward-rectifying and Ca

21

-

sensitive K

1

currents or T- and L-type Ca

21

channels.

Studies on the intra- and extracellular application of argio-

toxin-636 found that it reduced the excitability of rat DRG

neurons partly by inhibiting Na

1

channels, as mentioned

previously, in addition to dose-dependently inhibiting

voltage-activated K

1

channels without affecting the voltage

dependence of activation or steady-state inactivation (Scott

et al., 1998). More recently, argiotoxin was found to block

outward currents through the strong inward rectifying chan-

nel Kir2.1, probably via the insertion of the polyamine tail

lengthwise into the channel pore (Lee et al., 1999).

3.6. Chloride channel toxins (Table 3)

Chloride channels are involved in a multitude of

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

234

Table 3

Spider toxins affecting potassium and chloride channels

Spider

Channel/current inhibited

Reference/s

Potassiun channel Toxins

Hanatoxin 1 and 2

G. spatulata (now Phrixotrichus

spatulata)

Shab related (Kv2.1) and Shal related K

1

channels

Swartz and MacKinnon, (1995)

v-grammotoxin SIA

G. spatulata

voltage gated K

1

channels

Li-Smerin and Swartz, (1998)

HpTx (1±3)

H. venatoria

Kv4.2 channels

Sanguinetti et al., (1997)

PaTx 1 and 2

H. versuta

Kv4.2 and 4.3 channels

Diochot et al., (1999)

SGTx1

S. griseipe

fast transient and delayed recti®er currents

Marvin et al., (1999)

Tx3-1

P. nigriventer

A-type K

1

current

Kushmerick et al., (1999)

Arg 636

Argiope sp.

Kir2.1 channels

Scott et al., (1998); Lee et al.,

(1999)

Chloride channel toxins

Arg 636

Argiope sp.

inhibit Ca

21

-activated Cl

2

currents

Sutton et al., (1998b)

sFTX3.3

synthetic analogue of FTX

inhibit Ca

21

-activated Cl

2

currents

Sutton et al., (1998b)

FTX3.3

synthetic A. aperta toxin

inhibit Ca

21

-activated Cl

2

currents

Sutton et al., (1998b)

background image

physiological processes including transepithelial transport,

cell volume control and acidi®cation of intracellular orga-

nelles (Jentsch et al., 1999). They are also important for the

regulation of cellular excitability as components of

receptors for the inhibitory neurotransmitters GABA, in

the vertebrate CNS (Bormann, 1988), and glutamate, in

the invertebrate nervous system (Osborne, 1996). Although

important, Cl

2

channels are not crucial for the initiation and

propagation of action potentials or transmission at the

neuromuscular junction and few toxins have been found

which affect them. However, Sutton et al. (1998b) have

recently shown that the polyamine, sFTX3.3 inhibits both

voltage-activated Ca

21

currents and Ca

21

-activated Cl

2

currents (I

Cl(Ca)

) in rat cultured dorsal root ganglion neurons,

whilst synthesised FTX and argiotoxin-636 inhibit I

Cl(Ca)

but

not the Ca

21

currents.

3.7. Toxins that stimulate the release of neurotransmitters

Spiders of the genus Latrodectus (family Theridiidae) are

found throughout the world (Bucherl, 1971). Likewise,

latrodectism, or envenomation by a spider of this genus, is

a world-wide phenomenon (Maretic, 1978). Human enve-

nomation by Latrodectus sp. spiders consists of a character-

istic set of symptoms including; severe pain radiating from

the bite site to the muscles of the back, abdomen and lower

limbs, restlessness, respiratory dif®culty, abdominal rigid-

ity, transient hypertension, profuse sweating, salivation,

nausea and vomiting (Duchen and Gomez, 1984).

Latrodectus venom has been found to cause massive

neurotransmitter release from a variety of nerves in both

vertebrates and invertebrates (Longnecker et al., 1970;

Frontali et al., 1972; Kawai et al., 1972; Cull-Candy et al.,

1973; Pinto et al., 1974) resulting in blockade of nerve

transmission leading to muscle paralysis (Henkel and

Sankaranarayanan, 1999). The effects on nerves from

vertebrates, insects and crustaceans have been attributed to

several high molecular weight proteins selective for each

type of animal (Frontali et al., 1976; Ornberg et al., 1976;

Fritz et al., 1980). The vertebrate selective toxin, a 130 kDa

protein, was named a-latrotoxin (a-LTX) (Tzeng and

Siekevitz, 1978) and the names latroinsectotoxins (LITs)

and a-latrocrustatoxin (a-LCT) have been proposed for

the insect and crustacean selective toxins, respectively

(Grishin, 1998).

Whole venom from black widow spiders and a-latrotoxin

cause transmitter release and depletion of synaptic vesicles

(Harvey, 1990). This may be explained by several possible

actions of a-latrotoxin at the presynaptic membrane.

Finkelstein et al. (1976) found that the component active

at the neuromuscular junction interacts irreversibly with

arti®cial phospholipid membranes to form non-selective

cation channels, a property which allows the in¯ux of extra-

cellular Ca

21

into the nerve terminal leading to vesicular

exocytosis. It has been noted that a-latrotoxin-stimulated

exocytosis at neuromuscular synapses is independent of

extracellular Ca

21

(Ceccarelli et al., 1979; Henkel and

Betz, 1995) whereas secretion of catecholamines from rat

adrenal chromaf®n cells is dependent on extracellular

calcium (Barnett et al., 1996), suggesting different modes

of action of a-latrotoxin.

A receptor for a-latrotoxin was isolated from bovine

solubilised brain membranes (Petrenko et al., 1990) and

found to bind to synaptotagmin, a synaptic vesicle-speci®c

membrane protein involved in exocytosis (Petrenko et al.,

1991). The isolated receptor was subsequently identi®ed as

a member of the neurexins, a family of single transmem-

brane domain, synaptic cell surface proteins (Ushkaryov et

al., 1992) the interaction with which is dependent on

calcium (Davletov et al., 1995). The role of neurexins,

speci®cally neurexin 1a as the calcium-dependent receptor

for a-latrotoxin has recently been con®rmed (Geppert et al.,

1998; Sugita et al., 1999). A second, calcium-independent

receptor for a-latrotoxin (CIRL or latrophilin) was also

isolated from bovine brain membranes by af®nity

chromatography (Krasnoperov et al., 1996) and found to

be a member of the secretin family of G-protein coupled

receptors (Lelianova et al., 1997). Overall the mechanism of

action of a-latrotoxin appears to consist of several

components, at low concentrations it stimulates exocytosis

by binding to the receptors neurexin and/or latrophilin,

while at higher concentrations it forms cation channels

directly increasing cytosolic calcium (Henkel and

Sankaranarayanan, 1999).

Venom gland extract from another theridiid spider,

Steatoda paykulliana, at mg/ml concentrations, also

increased the conductance of arti®cial lipid membranes

and at higher concentrations stimulated the release of neuro-

transmitter from PC12 cells, but did not contain a-latrotoxin

(Cavalieri et al., 1987).

3.8. Toxins affecting cholinergic transmission

In vertebrates, acetylcholine (ACh) is well established as

a transmitter in autonomic and motor neurons as well as

being an important central neurotransmitter (Florey,

1967). ACh is also a major neurotransmitter in the central

nervous system of invertebrates (Osborne, 1996). Receptors

for ACh are broadly classi®ed into nicotinic and muscarinic

subtypes, depending on their sensitivity to nicotine and

muscarine, respectively. A third type of ACh receptor, the

mixed nicotinic/muscarinic receptor is present in inverte-

brates (Osborne, 1996). Nicotinic receptors are ligand-

gated ion channels made up of ®ve subunits, each with

four membrane spanning domains, whilst muscarinic recep-

tors belong to the G-protein-coupled receptor superfamily.

Venom from the orb weaving spider, Argiope lobata, not

only caused postsynaptic block of glutamatergic synapses

but also reversibly blocked cholinergic transmission in frog

skeletal muscle (Usmanov et al., 1983). In a subsequent

study, Usmanov et al. (1985) found that the ability to post-

synaptically block cholinergic synapses was common

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

235

background image

to seven other spiders from the family Araneida. They also

separated the venom of A. lobata and found seven toxic

fractions, of which one was selective for vertebrate neuro-

muscular transmission while the other six affected glutama-

tergic transmission of the locust. Nicotine-induced, but not

high K

1

-induced, secretion of catecholamines and ATP

from bovine chromaf®n cells was inhibited by v-agatoxin

IVA, which also reversibly blocked nicotine-induced inward

current suggesting that, in addition to blocking P-type Ca

21

channels, it also inhibits the neuronal nicotinic channel

(Granja et al., 1995). The polyamine amide JSTX-3, but

not argiotoxin-636, was also found to affect cholinergic

transmission causing voltage-dependent inhibition of ACh-

activated currents in rat phaeochromocytoma PC12 cells

(Liuet al., 1997).

3.9. Other neurotoxins

The potential use of spider toxins as tools for studying

physiological processes, identifying and characterising

neuronal ion channels and as new selective insecticides

has led to the isolation and characterisation of many novel

peptides. In response to the latter potential use of spider

venoms, there has been a shift in the emphasis of toxicity

assays toward using insects rather than more traditionally

used vertebrate laboratory animals (Bettini and Brignoli,

1978; Quistad et al., 1992; Atkinson et al., 1996; Manzoli-

Palma et al., 2000). However, mouse toxicity assays are still

widely used as a means of detecting novel neurotoxins

active on the vertebrate CNS (Escoubas et al., 1998). In

addition to the neurotoxins previously described, for

which distinct mechanisms of action have been reported,

many neurotoxic peptides have been isolated but as yet,

no mode of action determined.

The isolation from the East African tarantula, Pterino-

chilus sp., of one of at least 12 peptides toxic to mice was

reported by Bachmann (1982). Fraction A4/4 is a 77 amino

acid peptide that, like most of the toxic peptides isolated

from spiders thus far, appears to be tightly folded with

four disulphide bonds and causes death in mice by respira-

tory paralysis. The mode of action was not investigated but a

presynaptic action was suggested as unlikely, due to the

rapid reversibility of sublethal doses.

While characterising the venom of the tarantula

Eurypelma californicum, Savel-Neimann (1989) isolated

and sequenced a 38 amino acid peptide, ESTX, lethal to

cockroaches. A 39 amino acid peptide, identical to an

isoform of ESTX, and another unrelated protein toxin

showing similarities to toxin Tx2-9 from P. nigriventer,

have been isolated from the Mexican red knee tarantula

Brachypelma smithii (Kaiser et al., 1994). Escoubas et al.

(1997) isolated two novel peptide neurotoxins (lasiotoxins 1

and 2) from the venom of the tarantula Lasiodora

parahybana, which were lethal to mice and share 74%

homology with toxins isolated from the tarantulas

E. californicum and B. smithii. Upon initial fractionation

of the venom, this group noticed an interesting partition of

the vertebrate and invertebrate toxicity into different frac-

tions with little overlap. The insecticidal fractions contained

low molecular weight components while the vertebrate

toxicity was concentrated in peptide components of

3.7±7.3 kDa. Nine insecticidal peptides were isolated from

the venom of the trapdoor spider Aptostichus schlingeri, of

which six were sequenced (Skinner et al., 1992). Seven of

the isolated toxins caused ¯accid paralysis in tobacco horn-

worm (Manduca sexta) and beet armyworm (Spodotera

exigua) larvae and the sequenced peptides contained three

or four disulphide bonds. Two insecticidal peptides, Tx4(7)

and Tx4(6-1), which had no effects in mice, were isolated

from toxin fraction four of P. nigriventer venom (Figueiredo

et al., 1997). The peptide Tx4(6-1) and pooled fraction

(Tx4) stimulated the release of glutamate from cockroach

muscle slices. PhTx1, also from P. nigriventer, is lethal to

mice causing neurotoxic symptoms (tail elevation, excita-

tion and spastic paralysis) upon i.c.v. injection (Rezende et

al., 1991). In the mouse phrenic nerve diaphragm prepara-

tion, PhTx1 had no effect on indirectly evoked twitch height,

resting membrane potential or miniature end-plate poten-

tials. However, it did induce morphological changes in

nerves (vacuolization) and 20±30% of muscle ®bres

(vacuolisation with damaged mitochondria and sarcoplas-

mic reticulum) suggesting both neurotoxic and myotoxic

actions, although the mechanism remains unclear

(Mattiello-Sverzut et al., 1998).

4. Non neurotoxic peptides (Table 4)

The majority of the toxic proteins or peptides isolated to

date from spider venoms are either neurotoxins or enzymes.

However, several peptides displaying novel activities have

been discovered.

A necrotoxic peptide of 6.7 kDa appears to be the

predominant toxic component of venom from the female

Arkansas tarantula Dugesiella hentzi (Lee et al., 1974).

The toxin was found to bring about histological changes at

the site of injection and in the heart, where it produced

lesions typi®ed by acute focal areas of myocardial necrosis.

Chan et al. (1975) noticed that the co-administration of

adenosine 5

0

-triphosphate, also present in the venom in

substantial quantities, had a synergistic effect with the

necrotoxin, signi®cantly decreasing the LD

50

values when

injected into mice. The venom of the Singapore tarantula

Corecnemius validus contains another myotoxic peptide,

Covalitoxin-I, which was recently isolated, sequenced and

chemically synthesised, and causes necrosis of mouse

skeletal muscle (Balaji et al., 1999a). Two other peptides

(Covalitoxins II and III) from C. validus venom have been

isolated, sequenced and synthesised, however no toxicity or

functional studies have been reported as yet (Balaji et al.,

1999b).

In addition to the neurotoxic actions of P. nigriventer

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

236

background image

venom, it has been found to contract rabbit vascular smooth

muscle (Antunes et al., 1993) and cause an increase in the

vascular permeability of rabbit and rat skin (Antunes et al.,

1992) by activating the local tissue kallikrein±kininogen±

kinin system (Marangoni et al., 1993). A series of peptides

with activity accounting for these actions on smooth muscle

have been isolated. PNV1 is a 13.9 kDa protein of approxi-

mately 125 amino acids with two disulphide bonds which

causes contractions in rabbit pulmonary arteries and mesen-

teric veins and has an N-terminal sequence different from

other peptides so far isolated from P. nigriventer venom

(Marangoni et al., 1993). A second large peptide, PNV2

(102 amino acids and 12.1 kDa), which contracts rabbit

vascular smooth muscle was isolated and found to have

four disulphide bonds and an N-terminal sequence different

to that of PNV1 (Bento et al., 1993). The peptide PNV3,

isolated by Bento et al. (1995), appears to be responsible for

the increase in microvascular permeability in rabbit skin and

has an N-terminal sequence which shares 60±70% homol-

ogy with toxins Tx2.1, Tx2.5 and Tx2.6 from the same

spider. Finally, PNV4, a 16.6 kDa peptide of 147 residues,

is one of two fractions, generated by reverse-phase HPLC of

P. nigriventer venom, responsible for causing relaxation of

the rabbit corpus cavernosum (Rego et al., 1996).

Short peptides that potentiate the action of kinins, both in

vitro and in vivo, have been found in the venoms of snakes

(Ferreira, 1965; Ferreira et al., 1992, 1995) and scorpions

(Ferreira et al., 1993) and led to the development of angio-

tensin converting enzyme (ACE) inhibitors used clinically

in the treatment of some forms of hypertension (Opie and

Kowolik, 1995). In 1990, Sosnina et al. (1990) isolated two

peptides from the venom of the European black widow

spider Latrodectus tredecimguttatus that inhibited the

activity of ACE. An undecapeptide (BPP-S) that signi®-

cantly potentiates the effects of bradykinin on smooth

muscle and inhibits ACE in vitro was also isolated from

the venom of Scaptocosa raptoria, a Brazilian wolf spider

(Ferreira et al., 1996). Two kinin-like peptides (peptide-S

and peptide-R), which were almost equipotent with brady-

kinin in contracting guinea-pig isolated ileum, were subse-

quently isolated from the venom of the same spider (Ferreira

et al., 1998). Kinin-like peptides have also been found in the

venoms of many other animals including snakes, wasps and

frogs (Schachter and Thain, 1954; Yasahura et al., 1973;

Ferreira and Henriques, 1992).

The lycotoxins (I and II, 25 and 27 amino acids, respec-

tively) are two novel peptides isolated from the venom of

the North American wolf spider Lycosa carolinensis with a

broad range of actions (Yan and Adams, 1998). In insect

body wall muscles, lycotoxin I caused complete loss of cell

membrane potential and blockade of neuromuscular trans-

mission. At high concentrations lycotoxin I lyses rabbit

erythrocytes, causing 55% haemolysis. Both lycotoxins I

and II caused ef¯ux of calcium ions from, and prevented

sequestration of

45

Ca

21

by, rat brain synaptosomes. In

addition, both peptides possess potent antimicrobial activity

against both prokaryotic and eukaryotic cells in plate growth

inhibition assays. The lycotoxins have predicted secondary

structures of amphipathic a-helices, a con®guration

characteristic of the antimicrobial pore-forming peptides

adenoregulin, dermaseptins and magainins. Indeed, the abil-

ity to form pores in biological membranes can account for

all the observed actions of lycotoxins I and II (Yan and

Adams, 1998). An antibacterial peptide was also found in

the venom of the wolf spider Lycosa singoriensis (Xuet al.,

1989). In addition, ®ve antibacterial peptides of 3±4 kDa

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

237

Table 4

Non-neurotoxic peptides from spider venoms

Toxin/Spider

Action/effect

References

D. hentzi

necrotoxin, myocardial necrosis

Lee et al., (1974); Chan et al., (1975)

Covalitoxin1: C. validus

myotoxin, skeletal muscle necrosis

Balaji et al., (1999a)

PNV1: P. nigriventer

contracts vascular smooth muscle

Marangoni et al., (1993)

PNV2: P. nigriventer

contracts vascular smooth muscle

Bento et al., (1993)

PNV3: P. nigriventer

increases microvascular permeability

Bento et al., (1995)

PNV4: P. nigriventer

relaxes rabbit corpus cavernosum

Rego et al., (1996)

L. tredecimguttatus

inhibits angiotensin converting enzyme (ACE)

Sosnina et al., (1990)

BPP-S: S. raptoria

potentiates bradykinin/inhibits ACE

Ferreira et al., (1996)

peptide-S: S. raptoria

kinin-like peptide

Ferreira et al., (1998)

peptide-R: S. raptoria

kinin-like peptide

Ferreira et al., (1998)

Lycotoxins I and II: L. carolinensis

antimicrobial activity

Yan and Adams, (1998)

L. singoriensis

antibacterial activity

Xuet al., (1989)

C. salei

®ve antibacterial peptides

Haeberli et al., (2000)

SHLP-1: S. huwena

lectin-like peptide

Liang and Pan, (1995)

A. schlingeri, Aphonopelma sp., Filistata sp.,

L. deserta, and P. tristis

cytotoxic to mouse neuroblastoma cells

Cohen and Quistad, (1998)

Salticidae (jumping spider venom)

cytotoxic to mouse and insect cells

Cohen and Quistad, (1998)

L. reclusa

cytotoxic to human neutrophils

Majeski et al., (1977)

background image

have recently been isolated from the venom of C. salei

(Haeberli et al., 2000). The activity of antibacterial peptides

is thought to be due to a lytic action on bacterial cell walls

but the mode of lysis remains unknown.

Lectins are carbohydrate-binding proteins, found primar-

ily in plant seeds, which can cause agglutination of erythro-

cytes. Several lectins have been isolated from snake venoms

(Ogilvie et al., 1986; Bruno, 1990). Liang and Pan (1995)

isolated and sequenced the ®rst lectin-like peptide from

spider venom. The venom of the Chinese bird spider S.

huwena was found to contain substantial amounts of a 32

amino acid peptide (S. huwena lectin-like peptide-I,

SHLP-I) that agglutinated both mouse and human erythro-

cytes. SHLP-I showed homology with a fragment of great

nettle lectin as well as the neurotoxin HWTX-I, but

displayed no toxicity when administered to mice.

Cohen and Quistad (1998) conducted a study examining

the cytotoxic activity of venom from 26 species of spider

and four other arthropods on mammalian and insect cultured

cells. The insect cell line Sf9 was more sensitive to the

venom of spiders from the families Araneidae, Lycosidae

and Oxyopidae whilst venom from Aptostictnus schlingeri,

Aphonopelma sp., Filistata sp., Loxosceles deserta and

Plectreurys tristis were preferentially toxic to the mouse

neuroblastoma cell line N1E-115. Venom from the jumping

spiders (family Salticidae) were found to be the most

cytotoxic, being equipotent in insect Sf9 and mouse N1E-

115 cells with IC

50

values less than 1 mg venom protein per

ml in both lines. Although no toxic components were

isolated and characterised, this study introduced another

novel activity of spider venoms, which may be in part due

to enzymes or toxins already described. Previously, Majeski

et al. (1977) had reported that venom from the brown recluse

spider, Loxosceles reclusa, was directly cytotoxic to human

neutrophils at high concentrations as well as inhibiting

neutrophil chemotaxis toward complement derived

chemotaxins.

5. Enzymes (Table 5)

Enzymes are an important and common component of the

venom of many animals including bees, wasps, snakes and

lizards, with several possible functions. In this respect the

venom of spiders is no different, containing a wide variety of

enzymatic activities.

Hyaluronidase or hyaluronidase activity has been found

in the venom of many spiders, both mygalomorphs and

labidognaths. Nearly 50 years ago the venoms of the

Brazilian spiders Lycosa raptoria and Ctenus nigriventer

(now Phoneutria nigriventer) were reported to contain a

considerable amount of a hyaluronidase-like substance simi-

lar to the testicular enzyme (Kaiser, 1956). Shortly after,

Cantore and Bettini (1958) found hyaluronidase activity in

the venom of the European widow spider L. tredecimgutta-

tus (cited in Geren and Odell, 1984). Schanbacher et al.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

238

Table 5

Enzyme activities found in spider venoms

Enzyme activity

Spider

Reference/s

Hyaluronidase

L. raptoria (now S. raptoria)

Kaiser, (1956)

P. nigriventer

Kaiser, (1956)

L. tredecimguttatus

Cantore and Bettini, (1958) (cited in Geren and Odell, (1984))

D. hentzi

Schanbacher et al., (1973a)

L. reclusa

Wright et al., (1973); Geren et al., (1976)

L. laeta

Schenone and Suarez, (1978)

L. rufescens

Young and Pincus, (2001)

C. salei

Kuhn-Nentwig et al., (1994)

A. robustus

Sutherland, (1978)

E. californicum

Savel-Niemann, (1989)

L. cylindrata/murina and L. godeffroyi

Rash and Hodgson, unpublished observations

Phosphodiesterase

A. robustus, Aphonopelma cratus and L. mactans Russell, (1966)

Alkaline phosphatase

L. reclusa

Heitz and Norment, (1974); Norment et al., (1979)

Esterase

L. reclusa

Wright et al., (1973); Norment et al., (1979)

ATPase

L. reclusa

Geren et al., (1976)

L. laeta

Schenone and Suarez, (1978)

Sphingomyelinase D

L. reclusa

Forrester et al., (1978)

Kininase (endopeptidase)

L. tredecimguttatus

Akhunov et al., (1981)

Collagenase

N. edulis, E. transmarina and I. Immanis

Atkinson and Wright, (1992)

Peptide isomerase

A. aperta

Shikata et al., (1995)

Phospholipase A

female H. versutus

Sheumack et al., (1984)

Eresus niger

Usmanov and Nuritova, (1994)

Proteases

see text for discussion on the presence of

protease activity in spider venoms

background image

(1973a) identi®ed hyaluronidase as a major constituent of

venom from the tarantula D. hentzi. They subsequently

isolated the enzyme and found it to have a molecular weight

of approximately 39 kDa (Schanbacher et al., 1973b). Since

then hyaluronidase activity has been found in the venom of

Loxosceles reclusa, L. laeta and L. rufescens (Wright et al.,

1973; Geren et al., 1976; Schenone and Suarez, 1978;

Young and Pincus, 2001), the wandering spider Cupiennius

salei (Kuhn-Nentwig et al., 1994), Lycosa godeffroyi and

Lampona cylindrata/murina (Rash and Hodgson, unpub-

lished observations). In addition, hyaluronidases have

been isolated from the venom of the Sydney funnel-web

spider A. robustus (Sutherland, 1978) and the tarantula

E. californicum (Savel-Neimann, 1989). The substrate for

hyaluronidase is the mucopolysaccharide, hyaluronic acid,

which is a major constituent of the extracellular matrix

(Kreil, 1995). This has lead several researchers to postulate

that the hyaluronidase in animal venoms facilitates the

spread of other venom components by hydrolysing connec-

tive tissue (Schanbacher et al., 1973b; Sutherland, 1978).

Venoms from L. raptoria and C. nigriventer were

reported to possess proteolytic activity similar to that of

trypsin hydrolysing casein and ®brin (Kaiser, 1956). In addi-

tion, Kaire (1963) found a heat stable, neutral protease in the

venom of A. robustus. Three fractions, possibly different

forms of the one enzyme, displaying proteolytic but not

esterase activity were puri®ed from the venom of Pampho-

beteus roseus (Mebs, 1972). There have been several

con¯icting reports as to the presence of protease activity

in brown recluse spider venom. L. reclusa venom, obtained

by extraction of the dissected venom apparatus, was

reported to be devoid of proteolytic activity using casein

and haemoglobin as substrates (Geren et al., 1973). In

contrast, Jong et al. (1979), using venom obtained by elec-

trostimulation, isolated a proteolytic enzyme of 29.6 kDa

which hydrolysed the amide linkage of amino acids contain-

ing aliphatic, aromatic or basic side chains. Using the same

assay system as Jong et al. (1979); Rekow et al. (1983)

found protease activity in the cephalothorax and abdomen

extracts but not in venom apparatus extracts or the puri®ed

toxin from L. reclusa. Perret (1977) showed that the proteo-

lytic activity of tarantula venoms was due to the contamina-

tion of milked venom with digestive secretions expelled at

the time of milking. In support of this, Kuhn-Nentwig et al.

(1994) found proteolytic activity in the venom of C. salei

upon `normal' milking, which was completely lacking when

the chelicera had been carefully cleaned or when venom

from dissected glands was tested. The authors suggest that

the lack of venom protease activity is common to all spiders

and that when it is detected it is due to contamination during

the milking process.

In contrast to this view, two metalloproteinases from the

venom of Loxosceles intermedia have recently been identi-

®ed and characterised (Feitosa et al., 1998). Zymogram

analysis of venom, obtained by electrical stimulation, indi-

cated a 28 kDa band with ®bronectinolytic and ®brinogen-

olytic activity and a 35 kDa molecule with gelatinolytic

activity. Furthermore, two gelatinolytic molecules were

detected in the venom of L. intermedia following treatment

with trypsin (Veiga et al., 2000). The enzymes were

identi®ed as large serine-proteases with molecular weights

of 85- and 95-kDA. The authors suggested that upon enve-

nomation these zymogens may be activated by tissue

proteases and may act synergistically with other venom

components to induce necrosis. The venom used in both

of these studies was from the same source and obtained by

electrostimulation of the cephalothorax again raising the

question of contamination with digestive secretions.

However, Veiga et al. (2000) suggested that the narrow

proteolytic effect of the enzymes they have characterised

supports the fact that they are venom enzymes, as digestive

enzymes originating in the stomach of the spider would be

expected to have broad substrate speci®cities. In support of

proteolytic enzymes being constituents of spider venoms,

using zymogram analysis and venom gland extracts, we

have identi®ed several bands of activity (from ,18.5 to

around 32.5 kDa) in the venoms of the Australian spiders

Lycosa godeffroyi and Lampona cylindrata/murina against

both casein and gelatine, but not ®brinogen (Authors unpub-

lished observations). The use of venom gland extracts mini-

mises the likelihood of contamination with digestive

enzymes, although the possibility of contamination from

glandular tissue or adhering muscle remains. At this point

the studies providing evidence that spider venom itself

contains proteolytic enzymes still harbour the possibility

of some kind of contamination, be it from surrounding

tissues or digestive secretions. Perhaps in the not too distant

future the ever expanding cDNA libraries will answer this

question once and for all.

In a study examining the enzyme activity of several

species of snakes and arthropods, Russell (1966) found

phosphodiesterases in the venom of A. robustus,

Aphonopelma cratus, and the North American black

widow, Latrodectus mactans. The venoms of the brown

spiders L. laeta and L. reclusa were found to lack both

phosphodiesterase and collagenase activity (Suarez et al.,

1971; Wright et al., 1973; Schenone and Suarez, 1978).

However, they did contain enzymes with the ability to

hydrolyse the nucleotide ATP (Geren et al., 1976; Schenone

and Suarez, 1978). Wright et al. (1973) reported the

presence of esterase activity in the venom of L. reclusa

which was also shown by ¯uorometric assay to contain

alkaline phosphatase activity (Heitz and Norment, 1974).

The presence of these enzyme activities in L. reclusa

venom was later con®rmed when Norment et al. (1979),

using electrophoretic assays, detected two protein bands

with alkaline phosphatase activity and three bands exhibit-

ing esterase activity. In a study on L. reclusa venom-induced

lysis of red blood cells, Forrester et al. (1978) partially

puri®ed a haemolytic fraction that caused degradation of

the sphingomyelin component of red blood cell membranes.

This observation was found to be due to sphingomyelinase

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

239

background image

D-type activity rather than phospholipase C or phospholi-

pase A

2

activity. This toxic enzyme will be discussed further

in the section on necrotic arachnidism.

A bradykinin-inactivating enzyme (kininase) has been

isolated from the venom of L. tredecimguttatus (Akhunov

et al., 1981). Further characterisation of the enzyme found

it to be a thiol endopeptidase which cleaved the -Pro(7)-

Phe(8)-bonds of both bradykinin and angiotensin I. In the

same study it was noted that whole venom had no proteoly-

tic activity against casein or haemoglobin (Akhunov et al.,

1996). These results are interesting as speci®c non-toxic

kininases are useful tools for studying the mammalian

kallikrein±kinin system and have potential as therapeutic

agents.

Studies on the enzymic activities of venom from several

species of spider have failed to detect the presence of

collagenase activity (Kaiser and Raab, 1967; Suarez et al.,

1971; Wright et al., 1973; Schenone and Suarez, 1978).

Conversely, Atkinson and Wright (1992) reported the

presence of measurable amounts of collagenase activity in

the milked venom of three Australian spiders, Nephila

edulis, Eriophora transmarina (orb weavers) and Isopeda

immanis (huntsman spider), and in midgut extracts of all 13

species tested.

In biological systems the translation and synthesis of

proteins only involves the use of L-form amino acids

(Branden and Tooze, 1991), however D-amino acid contain-

ing peptides have been isolated from a variety of higher

organisms (Kreil, 1994). Following the discovery of a

D-form amino acid (Ser

46

) in the P-type calcium channel

toxin v-agatoxin-TK (also designated as v-AgaIVB)

(Kuwada et al., 1994), a novel peptide isomerase, which

speci®cally inverts the chirality of Ser

46

, was isolated and

characterised (Shikata et al., 1995).

In several studies on venom from Loxosceles species

spiders and D. hentzi, neither phospholipase A nor phospho-

lipase C activities were found (Suarez et al., 1971;

Schanbacher et al., 1973a; Schenone and Suarez, 1978).

Conversely, in a study comparing the venoms of several

species of Australian funnel-web spiders, Sheumack et al.

(1984) detected phospholipase A activity only in the venom

of female Atrax versutus (now Hadronyche versutus).

Another phospholipase A, named component EnPA, has

subsequently been identi®ed in Eresus niger spider venom

and found to have anticoagulant activity (Usmanov and

Nuritova, 1994).

Therefore, spider venoms appear to contain a wide variety

of enzyme activities. Several of these enzymes, such as the

hyaluronidases of many venoms and the peptide isomerase

of A. aperta venom, may play a speci®c role in the enveno-

mation process or the post-translational modi®cation of

peptide toxins. Minton (1974) suggests that, as spiders

feed on the juices and lique®ed tissues of their prey, spider

venoms may have originally been digestive secretions and

may retain this function to varying degrees, thereby

explaining the abundance of enzymes present.

6. Low molecular weight components

In addition to the wide spectrum of biologically active

proteins, peptides and polyamine amides, spider venoms

contain a variety of low molecular weight components

displaying pharmacological activities, including biogenic

amines, free amino acids, polyamines, nucleotides and

inorganic salts.

6.1. Biogenic amines

Welsh and Batty (1963) found the amine 5-hydroxytryp-

tamine (5-HT) in the venoms or venom apparatus of a

variety of arthropods including seven species of Brazilian

spiders from the families Theraphosidae, Lycosidae and

Ctenidae. Venom gland extracts of the black widow spider

L. tredecimguttatus were also reported to contain 5-HT

(Pansa et al., 1972), as was the venom from male A.

robustus, the female of which lacked 5-HT but contained

trace amounts of 5-methoxytryptamine (Duf®eld et al.,

1979). Using gas chromatography and mass spectrometry,

Duf®eld et al. (1979) also detected and quantitated the

amines tyramine and octopamine in the venoms of both

male and female A. robustus.

Histamine is another biogenic amine commonly found in

arthropod venoms (Bettini and Brignoli, 1978; von Sicard et

al., 1989; Matuszek et al., 1992) and substantial amounts

have been detected in spider venoms. For example, L.

erythrognatha, P. nigriventer (Fischer and Bohn, 1957),

C. salei (Kuhn-Nentwig et al., 1994), Lampona cylindrata/

murina

1

(Rash et al., 2000b) and Lycosa godeffroyi (Rash et

al., 1998).

The vasoconstrictor action of Hololena curta venom on

rat thoracic aorta was found to be due to the presence of a

substance identi®ed by HPLC and electrochemical detection

as the catecholamine noradrenaline (Frew et al., 1994).

Noradrenaline has also been detected in the venom of

male, but not female, white-tailed spiders (L. cylindrata/

murina) (Rash et al., 2000b). Venom from female A.

robustus, in addition to the Australian huntsman spiders

Delana cancerides and Isopeda immanis, have also been

reported to display direct activity at adrenoceptors.

However, the agents responsible for this activity were not

identi®ed (Morgans and Carroll, 1976; Korszniak and Story,

1993).

Most of these biogenic amines have demonstrated roles as

neurotransmitters in the insect nervous system (Osborne,

1996), their presence in the venoms of spiders serving

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

240

1

The original publication refers to the white-tailed spider as

Lampona cylindrata. However, Platnick (2000) revised the genus

raising Lampona to family status, noted that the distribution of the

two most common species of Lampona (cylindrata and murina)

overlaps and that the spiders are almost indistinguishable. For this

reason spiders previously identi®ed as L. cylindrata will be referred

to as L. cylindrata/murina.

background image

several possible functions. Amines such as 5-HT and hista-

mine are well established agents of pain production suggest-

ing a possible role in self defence, or in the facilitation of

uptake of other venom components by increasing local

blood ¯ow and cell permeability (Welsh and Batty, 1963;

Kuhn-Nentwig et al., 1998). As many of the polyamine

toxins so far isolated from spiders are use-dependent block-

ers of neurotransmission, the presence of a variety of known

insect neurotransmitters, in particular glutamate and

g-aminobutyric acid, in their venoms may accelerate the

paralytic actions of these toxins.

6.2. Free amino acids

A wide variety of free amino acids have been detected in

the venoms of many spiders. Particularly common are the

amino acid neurotransmitters g-aminobutyric, glutamic and

aspartic acids, which have been found in the venom of

several mygalomorphs, D. hentzi, E. californicum and

A. robustus (GABA only) (Gilbo and Coles, 1964;

Schanbacher et al., 1973a; Savel-Neimann, 1989) as well

as araneomorphs, P. nigriventer and L. tredecimguttatus

(Fischer and Bohn, 1957; Bettini and Maroli, 1978).

The amino acid taurine occurs in the venom of L.

tredecimguttatus (Bettini and Maroli, 1978) and is present

in high concentrations in C. salei venom (Kuhn-Nentwig et

al., 1994). Kuhn-Nentwig et al. (1998) subsequently discov-

ered that higher levels of taurine and histamine in insect

haemolymph increased the sensitivity of the insect to the

peptide toxin CSTX-1 from the venom of C. salei. They

concluded that the role of these compounds was to act

synergistically with other toxins increasing their lethality.

Other amino acids so far detected in spider venoms include

glycine, serine, threonine, lysine, glutamine, alanine,

arginine, asparagine, leucine and histidine (Fischer and

Bohn, 1957; Gilbo and Coles, 1964; Bettini and Maroli,

1978; Duf®eld et al., 1979).

6.3. Other low MW components

The venoms of three tarantulas, D. hentzi, an

Aphonopelma sp. from Arizona, and E. californicum, have

been described as containing the adenine nucleotides ATP,

ADP and AMP (Chan et al., 1975; Savel-Neimann, 1989).

As mentioned previously, ATP was found to have a syner-

gistic effect with the necrotoxin isolated from D. hentzi

venom (Chan et al., 1975). The purine derivatives adeno-

sine, guanosine, inosine and 2,4,6-trihydroxypurine were

recently reported in the venom of Latrodectus menavodi

(Horni et al., 2001). Venom from female L. cylindrata/

murina has been shown to have activity at adenosine recep-

tors which was inhibited upon exposure to adenosine deami-

nase but this venom component has not yet been isolated

(Rash et al., 2000b).

Substantial amounts of citric acid are present in the

venoms of A. robustus (Duf®eld et al., 1979), the tarantula

Grammostola cala and brown recluse spider L. reclusa,

among other arthropods (Fenton et al., 1995). In addition,

Fenton et al. (1995) found that citrate inhibits bee venom

phospholipase A

2

. However, no role for citrate in the venom

of spiders was suggested.

In addition to the polyamine amide compounds discussed,

the free polyamines spermine, spermidine, cadaverine and

putrescine have been detected in the venoms of A. robustus

(Duf®eld et al., 1979), D. hentzi, Aphonopelma sp. and

Aphonopelma emilia (Cabbiness et al., 1980). Savel-

Neimann (1989) detected trace amounts of free cadaverine

and spermidine in female E. californicum venom as well as

compounds containing spermine and aromatic molecules.

However, the author also points out that in the above studies,

which reported substantial concentrations of free polya-

mines, the venom used was hydrolysed with 6 M HCl and

that it cannot be excluded that those polyamines are released

from polyamine amides upon hydrolysis.

Other low molecular weight components found in spider

venoms thus far include glucose, lactic and phosphoric

acids, glycerol, urea and the inorganic ions sodium, potas-

sium, calcium, magnesium, chloride and phosphorous

(Wiener, 1961; Savel-Neimann, 1989; Kuhn-Nentwig et

al., 1994) some of which may be present to stabilise other

venom components, particularly neurotoxic peptides or

enzymes.

7. Necrotic arachnidism

Necrotic arachnidism is the development of necrotic cuta-

neous lesions following the bites of several species of

spider, particularly of the genus Loxosceles. The link

between skin ulceration and spider bites was suggested by

Schmaus (1929), but was not con®rmed in South America

until 1947 when Macchiavello (1947) associated the bite of

Loxosceles laeta to the `gangrenous spot of Chile', and 1957

in North America when Atkins et al. (1957) attributed necro-

tic arachnidism in the midwest to Loxosceles reclusa.

The envenomation syndrome of Loxosceles sp. spiders

consists of local necrotic lesions and, less commonly,

systemic reactions. The initial bite is reportedly painless,

often going unnoticed, with mild to severe pain (probably

due to ischaemia), itching, swelling and tenderness preced-

ing the development of a blister surrounded by an area of

ischaemia. Over a few days the lesion becomes a dull, blue-

violet colour and the centre sinks, hardens and usually drops

off leaving an ulcer which can take 6±8 weeks to heal. Bites

in areas of thicker fatty tissue such as the thighs, abdomen

and buttocks result in the most severe cases of local necrosis

and scarring. The rare cases of lethal Loxosceles envenoma-

tion, mainly in children, are due to systemic reactions, the

symptoms of which include fever, malaise, weakness,

nausea, vomiting and haematological disturbances such as

haemolytic anaemia, thrombocytopenia and consumption

coagulopathy (Futrell, 1992).

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

241

background image

Early studies on the necrotic actions of venom from

L. reclusa found that rats were relatively resistant, while

mice, guinea-pigs and rabbits were susceptible to the

venom, with only rabbits and guinea-pigs developing the

dermonecrotic lesions seen in cases of human envenomation

(Morgan, 1969). As a result, rabbits have been used exten-

sively in studying the pathology and underlying mechanism

of Loxosceles induced dermonecrosis. The histopathologic

changes resemble cutaneous Arthus reaction (a type III

immune response involving the development of an in¯am-

matory lesion, characterised by induration, erythema,

oedema, haemorrhage and necrosis) and include oedema,

thickening of blood vessel endothelium, polymorphonuclear

leukocyte in®ltration, intravascular coagulation, vasodilata-

tion, destruction of blood vessel walls and haemorrhage

(Futrell, 1992). The accumulation of polymorphonuclear

leucocytes is required for the development of necrotic

lesions (Smith and Micks, 1970) and appears to be related

to the presence of circulating complement (Ward and

Cochrane, 1965). Exactly how the complement pathway is

involved is a topic of much debate (discussed by Futrell,

1992).

As mentioned previously, Forrester et al. (1978) partially

puri®ed a haemolytic fraction from L. reclusa venom that

also displayed sphingomyelinase D activity, both activities

being coincident within the electrophoretic pattern upon

further characterisation. The sphingomyelinase D

component (MW 31±32 kDa) has subsequently been

puri®ed and can account for most of the toxic actions of

whole L. reclusa venom, being able to induce necrotic

lesions, haemolyse red blood cells, cause platelet aggrega-

tion and cause death in laboratory animals (Babcock et al.,

1981; Kurpiewski et al., 1981). L. reclusa venom-induced

platelet aggregation appears to involve serotonin release

(Kurpiewski et al., 1981) and serum amyloid P component

(Rees, 1989; Gates and Rees, 1990), not C-reactive protein

as originally suggested (Rees et al., 1988). However,

C-reactive protein was reported to be required for venom-

induced lysis of human erythrocytes (Hufford and Morgan,

1981). The mechanism of action of L. reclusa venom and the

roles of C-reactive protein, serum amyloid P and serotonin

release are still not fully understood. However, the

dermonecrosis has been suggested to result from a direct

action of sphingomyelinase D on cell membranes at the

site of envenomation (Futrell, 1992).

There are seven species of Loxosceles found in Brazil

(Mota and Barbaro, 1995) with L. intermedia, L. gaucho

and L. laeta being the most common and clinically

important (Braz et al., 1999). The venoms of these three

species are antigenically cross-reactive and upon

SDS±PAGE analysis the dermonecrotic activity lies in the

major protein band of each with molecular weights of

35 kDa for L. gaucho and 32 kDa for both L. intermedia

and L. laeta (Barbaro et al., 1997). Tambourgi et al.

(1997) found the necrotoxic protein from L. intermedia

venom to be 35 kDa and further separated it into three

components P1, P2 and P3. The N-terminal sequence of

P2 shares 78% identity and 91% similarity with the sphin-

gomyelinase D isolated from the venom of L. reclusa,

suggesting that venoms of Loxosceles sp. spiders contain a

relatively homogenous family of toxic proteins.

Experimental envenomation of rabbits, by inducing the

Agelenid spider Tegenaria agrestis to bite a shaved patch of

skin, caused necrotic lesions closely resembling those

produced by Loxosceles venom (Vest, 1987a). T. agrestis,

also known as the hobo spider and the aggressive house

spider, has been implicated in several cases of human enve-

nomation resulting in dermonecrosis from the north west

United States (Vest, 1987b). Lesions resulting from

T. agrestis envenomation are characterised by induration,

blistering and necrotic ulcers requiring from 45 days to 3

years to heal. Systemic symptoms can include headache,

nausea, weakness, fatigue, dizziness, temporary memory

loss and vision impairment, with more severe cases some-

times resulting in aplastic anaemia, intractable vomiting and

profuse diarrhoea (Vest, 1987b).

The venom of two South African crab spiders of the

species Sicarius (testaceus and albospiosus) have also

been reported to cause tissue necrosis and haemorrhagic

lesions in rabbits suggesting that these spiders are also

potentially harmful to humans (Newlands, 1982; Van Aswe-

gen et al., 1997).

7.1. Necrotic arachnidism in Australia

The possibility that necrotising arachnidism occurs in

Australia was ®rst raised in the late 1970s and early 1980s

with the report of several cases of full thickness skin loss

following presumed envenomation (Sutherland, 1983a). In

no case was a creature seen to bite the victim. However,

circumstantial evidence suggested spiders as the likely

cause with Lycosa and Lampona species being singled out

as likely suspects (Sutherland, 1983b). Reports of three

further cases of substantial skin loss (one involving massive

dermonecrosis covering much of the victims right leg)

following suspected spider bites (Spring, 1987; Ibrahim et

al., 1989) highlighted the case for necrotic arachnidism

being a potential health issue in Australia (Sutherland,

1987). In these reports, as above, no offending creature

was felt or seen to bite the victims, sparking a large debate

as to the true cause of such a syndrome (Kemp, 1990;

Sutherland, 1990; Beardmore, 1991; Harvey and Raven,

1991; Hayman and Smith, 1991; Raven and Harvey, 1991;

Sutherland, 1991). After catching a small, brown, cigar

shaped spider (possibly L. cylindrata/murina) in the act of

biting her inner thigh, a 33 year-old woman developed an

acutely painful ulcerous lesion which required 3 weeks to

begin healing (Gray, 1989). Gray (1989) also reported treat-

ing three other cases of local ulceration following the bite of

a small brown spider. In contrast, White et al. (1989)

presented 36 cases of con®rmed spider bites, including

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

242

background image

eight by L. cylindrata/murina, none of which resulted in

local ulcers or necrosis.

To date, cases of positively identi®ed spider bites result-

ing in local tissue damage in Australia are rare. However,

several species of spider are still implicated, the prime

suspects being the white-tailed spider L. cylindrata/murina

(Skinner and Butler, 1995; Pincus et al., 1999; Woo and

Smart, 1999; Fig. 3), and the black house or window spider

Badumna insignis (Macmillan, 1989; Pincus et al., 1999).

In an editorial comment, White (1999) points out that until

Pincus et al. (1999) published 14 cases of de®nite or

suspected spider bites resulting in necrosis, only four

cases of tissue injury associated with spider bite in Austra-

lia had been published, and in none of these cases was the

spider formally identi®ed. He also notes that in only three

of the cases reported by Pincus et al. (1999) was the spider

positively identi®ed. The two cases involving the white-

tailed spider, L. cylindrata/murina, resulted in shallow or

small ulcers that healed within a month and were described

by White as `at the mild end of the necrotising arachnidism

spectrum'. Some authors agree that bites by L. cylindrata/

murina may produce small localised lesions or ulcers

(Raven and Harvey, 1991; Sutherland, 1991) but doubt

that the unexplained extensive skin loss sometimes seen

in Australia can be attributed to the action of spider

venom (Raven and Harvey, 1991). White et al. (1995)

acknowledges that there is `a poorly de®ned clinical entity

consisting of unexplained skin blistering and ulceration' in

Australia and has suggested that some individuals may be

particularly sensitive to spider venoms. In a follow-up

study to Pincus et al. (1999); Young and Pincus (2001)

detected hyaluronidase and protease activity in venom

extracts from Loxosceles rufescens, B. insignis and L.

cylindrata/murina. Interestingly, only venom extract

from Loxosceles rufescens, and the abdominal extracts of

B. insignis and L. cylindrata/murina, displayed sphingo-

myelinase activity.

In Australia, most of the cases where spider bite has been

associated with necrotic lesions, or indeed, a spider seen to

bite before the development of local necrosis, the spider was

identi®ed by the victim, not by an arachnologist. The

problem with this is illustrated in a letter to the editor of

Toxicon by Russell and Gertsch (1983) where they describe

ticks, assassin bugs, and even crickets and grasshoppers as

being brought to the hospital as `spiders' by patients bearing

a necrotic lesion. Clinicians and arachnologists are now

stressing the importance of witnessing the bite and having

the offending animal positively identi®ed by an expert if we

are to determine, without doubt, the creature or creatures

responsible for `necrotic arachnidism' in Australia (White,

1987; Raven and Harvey, 1991; Pincus et al., 1999; White,

1999; Isbister and Gray, 2000a). Interestingly, a recent

prospective study involving 52 bites by L. cylindrata/

murina, where the identity of the spider was con®rmed by

an arachnologist, reported the following clinical effects:

local pain in all victims, the presence of a persistent lump

(44%), red area (56%) or both (27%) but no cases of ulcera-

tion or necrosis (Isbister and Gray, 2000b).

8. Sex-linked variation in venom composition and

activity

The resulting symptoms from envenomation by many

creatures can vary dramatically from case to case. There

are several factors that can contribute to the severity of a

venomous bite or sting, such as the species of the offending

creature, the site of envenomation and the susceptibility of

the victim. The quantity and/or quality of venom injected

are factors that should also be considered. Within a given

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

243

Fig. 3. Female (a) and male (b) white-tailed spider (Lampona cylindrata/murina). scale bar ˆ 1 cm.

background image

species, the two latter aspects have been shown to vary

according to the geographic origin, size, age and sometimes

sex of the creature in question. For example venom from

female Buthus eupeus scorpions was found to be twice as

toxic to mice and guinea-pigs as venom from males

(Kashnikova, 1979). Despite the relative paucity of direct

comparison studies, the venoms of spiders appear to be no

exception, with differences between the venom of male and

female spiders of the same species having been noted for

many years. This is mainly because one sex proves to be

toxic to humans or mammals while the other does not. This

is the case with the theridiid spider Latrodectus mactans, the

notorious black widow, in which the venom of the female

spider was lethal to rats, while venom from males was

essentially non toxic (D'Amour et al., 1936). Maretic et

al. (1964) reported similar ®ndings for another theridiid

spider, Steatoda paykulliana, where guinea-pigs bitten by

female but not male spiders developed symptoms analogous

to those of latrodectism.

In contrast, the reverse is true for the Sydney funnel-web

spider A. robustus. Where the male spider has been respon-

sible for 14 human fatalities (Gray and Sutherland, 1978),

whereas venom from female A. robustus seems to lack

signi®cant amounts of the lethal neurotoxin, robustoxin

(Sheumack et al., 1984). In toxicity assays in new-born

mice, female A. robustus venom was found to be 19 times

less potent than venom from the male. In addition, the

venoms of male and female A. robustus displayed substan-

tially different HPLC pro®les and polyacrylamide gel

isoelectric focusing patterns. This re¯ects the degree of

sex-related variation of M. bradleyi venom, where the

male contained a potent neurotoxin with activity in avian

and mammalian nerve-muscle preparations whereas this

component appeared to be absent from female M. bradleyi

venom (Rash et al., 2000a). Interestingly, venoms from male

and female Blue Mountains funnel-web spiders Hadronyche

versutus (formerly Atrax versutus) were found to have very

similar LD

50

s in new born mice (0.7 and 0.5 mg/kg s.c.,

respectively) and HPLC pro®les (Sheumack et al., 1984).

However, isoelectric focusing patterns of venom from both

sexes revealed several differences in their respective protein

components.

In characterising the venom of an African tarantula

Scodra griseipes bred in captivity, CeÂleÂrier et al. (1993)

noted several gender differences. Female spiders produced

more venom which was of higher toxicity to mice and

contained three more electrophoretic bands than venom

from male spiders. Conversely, male S. griseipes venom

had a higher protein content than female venom. A higher

protein content in venom from male spiders as compared to

female spiders was also reported for both venom gland

extracts and venom obtained by electro-stimulation of the

sac spider Cheiracanthium furculatum (Croucamp and

Veale, 1999). Aside from the differences in biogenic

amine content and adenosine receptor activity of male and

female L. cylindrata/murina venom (Rash et al., 2000b) and

the apparent neurotoxic activity peculiar to venom from the

male, the protein content and SDS±PAGE patterns were

found to be very similar (authors unpublished observations).

Like L. cylindrata/murina, venom from males and females

of an Australian Lycosa sp. had almost identical

SDS±PAGE patterns. However, the female Lycosa sp.

venom contained signi®cantly more protein and myotoxic

activity than venom from the male (authors unpublished

observations). An analogous situation was noted for the

venom of the South American brown spider L. intermedia

(de Oliveira et al., 1999). Both male and female

L. intermedia had a similar pattern of bands upon

SDS±PAGE analysis. However, venom from female spiders

had a two-fold higher protein content, was richer in the

dermonecrotic toxin F35 and induced both complement-

dependent haemolysis and dermonecrosis more potently

than did venom from male L. intermedia. These ®ndings

supported the results of Kent et al. (1984) who reported

that although adult male and female L. reclusa spiders had

the same venom composition, the relative abundance of the

components differed. Finally, in a study examining the

variability of venom from several tarantulas, using HPLC

and MALDI-TOF-MS, Escoubas et al. (1999) found both

qualitative and quantitative differences in six of eight

species examined. These observations indicate that sex is

a substantial factor in the intraspeci®c variation of spider

venoms.

In general, there appears to be a trend toward the female

of the species having more toxic venom than that of the

male, which, given the maternal responsibilities, longer

life span of female spiders and the relatively short life of

the mature male, would make sense for the survival of the

species. However, the toxicity assays used in many of the

above studies were conducted on mammals rather than

insects. The latter constitute the main diet and therefore

presumably the speci®c target of spider venoms. As a result,

one must exercise caution in hypothesising on the possible

reasons for the sometimes marked gender differences

observed in spider venoms. Therefore, until more compar-

isons are made between venoms from male and female

spiders in light of their relevance to prey capture and survi-

val the purpose of the differences observed thus far remains

unclear.

9. Conclusions

Over millions of years spiders have evolved toxins, with

a variety of targets, affecting primarily the nervous systems

of their prey. As illustrated in this review, the majority of

the components of spider venom are either neurotransmit-

ters (excitatory and inhibitory amino acids), neurotoxins

themselves or components that aid in the stability, delivery

or effectiveness of these paralysing toxins to their site of

action such as enzymes, inorganic ions and nucleotides.

Given the dependence of neurotransmission on the

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

244

background image

movement of ions across membranes it is not surprising

that most of the neurotoxins speci®cally target neuronal

ion channels. It is likely that, with further work, neurotox-

ins with as yet unidenti®ed actions will prove to act at ion

channels of some description.

Spider venom neurotoxins are of interest for several

reasons. Due to the similarity of ion channels from both

the natural invertebrate prey and vertebrates, and the amaz-

ing selectivity of spider neurotoxins, they have already been

instrumental in our understanding of channel structure and

function. As the trend to screen spider toxins for more selec-

tive ligands progresses, they will surely continue to provide

new tools and lead compounds for the design of selective ion

channel activators and blockers with therapeutic potential.

In addition, toxins which discriminate between insect and

vertebrate channels may not only give us access to envir-

onmentally friendly insecticides but knowledge of the struc-

tural differences allows the possibility to design new, highly

selective, non-peptide insecticides. As discussed in the ®nal

section of this review, the sex of a spider or indeed many

venomous creatures plays a substantial role in the composi-

tion of venom and should not be overlooked in the search for

the next captopril or novel ion channel.

Acknowledgements

We gratefully acknowledge support from the Monash

University Faculty of Medicine Research Initiatives Scheme

and the Australian Research Council Small Grants Scheme.

We are grateful to David Humfrey for spider photography.

References

Adams, M.E., Bindokas, V.P., Hasegawa, L., Venema, V.J., 1990.

v-agatoxins: Novel calcium channel antagonists of two

subtypes from funnel-web spider (Agelenopsis aperta) venom.

J. Biol. Chem. 265, 861±867.

Adams, M.E., Herold, E.E., Venema, V.J., 1989. Two classes of

channel speci®c toxins from funnel web spider venom. J. Comp.

Physiol. 164A, 333±342.

Adams, M.E., Mintz, I.M., Reily, M.D., Thanabal, V., Bean, B.P.,

1993. Structure and properties of v-agatoxin IVB, a new

antagonist of P-type calcium channels. Mol. Pharmacol. 44,

681±688.

Akhunov, A.A., Golubenko, Z., Sadykov, A.S., 1981. Isolation of

kininase from the spider venom Latrodectus tredecimguttatus.

Dokl. Akad. Nauk. SSSR 25, 1478±1480.

Akhunov, A.A., Makevnina, L.G., Golubenko, Z., Paskhina, T.S.,

1996. Kininase of the Latrodectus tredecimguttatus venom: A

study of its substrate speci®city. Immunopharmacology 32,

160±162.

Alexander S.P.H., Peters J.A. 1999. Receptor and ion-channel

nomenclature supplement. Trends Pharmacol. Sci. (Suppl.).

Antunes, E., Marnagoni, R.A., Borges, N.C., Hyslop, S., Fontana,

M.D., de Nucci, G., 1993. Effects of Phoneutria nigriventer

venom on rabbit vascular smooth muscle. Braz. J. Med. Biol.

Res. 26, 81±91.

Antunes, E., Marnagoni, R.A., Brain, S.D., de Nucci, G., 1992.

Phoneutria nigriventer (armed spider) venom induces increased

vascular permeability in rat and rabbit skin in vivo. Toxicon 30,

1011±1016.

Aramaki, Y., Yashuhara, T., Higashijima, T., Yoshioka, M., Miwa,

A., Kawai, N., Nakajima, T., 1986. Chemical characterisation of

spider toxin, JSTX and NSTX. Proc. Jpn. Acad. Ser. B. 62,

359±362.

ArauÂjo, D.A.M., Cordeiro, M.N., Diniz, C.R., BeiraÄo, S.L., 1993.

Effects of a toxic fraction, PhTx

2

, from the spider Phoneutria

nigriventer on the sodium current. Nauyn-Schmeidebergs Arch.

Pharmacol. 347, 205±208.

ArauÂjo, D.A.M., Cordeiro, M.N., Richardson, M., Diniz, C.R.,

BeiraÄo, S.L., 1993. A novel class of polypeptide toxin modi®es

sodium current inactivation and activation in isolated frog

skeletal muscle. J. Physiol. 467, 365.

Ashe, J.H., Cox, C.L., Adams, M.E., 1989. Argiotoxin-636 blocks

excitatory synaptic transmission in rat hippocampal CA1

pyramidal neurons. Brain Res. 480, 234±241.

Atkins, J.A., Wingo, C.W., Sodeman, W.A., 1957. Probable cause

of necrotic spider bite in the Midwest. Science 126, 73.

Atkinson, R.K., Wright, L.G., 1992. The involvement of

collagenase in the necrosis induced by the bites of some spiders.

Comp. Biochem. Physiol. 102C, 125±128.

Atkinson, R.K., Vornax, E.J., Howden, M.E.H., 1996. Effects of

whole venom and venom fractions from several Australian

spiders, including Atrax (Hadronyche) species, when injected

into insects. Comp. Biochem. Physiol. 1114C, 113±117.

Babcock, J.L., Civello, D.J., Geren, C.R., 1981. Puri®cation and

characterization of a toxin from brown recluse spider

(Loxosceles reclusa) venom gland extracts. Toxicon 19,

677±689.

Bachmann, M., 1982. Isolation and partial characterization of a

toxin from the venom of the East African orthognath spider

Pterinochilius spec. Toxicon 20, 547±552.

Balaji, R.A., Sasaki, T., Gopalakrishnakone, P., Sato, K., Kini, R.M.,

Huat, B.B., 1999. Isolation and characterization of a novel

myotoxic peptide (Covalitoxin-I) from Corecnemius validus

(Singapore tarantula) venom. Proceedings of the 5th Asia-Paci®c

Congress on Animal, Plant and Microbial Toxins, 39.

Balaji, R.A., Sasaki, T., Gopalakrishnakone, P., Sato, K., Kini,

R.M., Huat, B.B., 1999. Isolation, characterization and chemical

synthesis of toxins from Singapore tarantula (Corecnemius

validus). Proceedings of the 5th Asia-Paci®c Congress on

Animal, Plant and Microbial Toxins, 38.

Barbaro, K.C., Ferreira, M.L., Cardoso, D.F., Sousa, M.V., Morhy,

L., von Eickstedt, V.R.D., Mota, I., 1997. Study of Loxosceles

spider venom. J. Venom. Anim. Toxins 3, 232.

Barnett, D.W., Liu, J., Misler, S., 1996. Single-cell measurements of

quantal secretion induced by alpha-latrotoxin from rat

chromaf®n cellsÐdependence on extracellular Ca

21

. P¯uÈgers

Arch. 432, 1039±1046.

Beardmore, G.L., 1991. Necrotising arachnidism in Australia. Med.

J. Aust. 155, 208.

Bento, A.C., Novello, J.C., Marangoni, S., Antunes, E., Giglio, J.R.,

Oliveira, B., de Nucci, G., 1993. Identi®cation of a new vascular

smooth muscle contracting polypeptide in Phoneutria

nigriventer spider venom. Biochem. Pharmacol. 46,

1092±1095.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

245

background image

Bento, A.C., Rego, E., Pedroso-Mariani, S.R., Mancuso, L.C.,

Giglio, J.R., Novello, J.C., Marangoni, S., Caracelli, I., Oliveira,

B., Antunes, E., de Nucci, G., 1995. Isolation of a polypeptide

from Phoneutria nigriventer spider venom responsible for the

increased vascular permeability in rabbit skin. Toxicon 33,

171±178.

Bettini, S., Brignoli, P.M., 1978. Review of the spider families. In:

Bettini, S. (Ed.). Handbook of Experimental Pharmacology,

Vol. 48, Arthropod Venoms. Springer-Verlag, New York, pp.

101±120.

Bettini, S., Maroli, M., 1978. Venoms of Theridiidae, genus

Latrodectus. In: Bettini, S. (Ed.). Handbook of Experimental

Pharmacology, Vol. 48, Arthropod Venoms. Springer-Verlag,

New York, pp. 149±184.

Bindokas, V.P., Adams, M.E., 1989. v-Aga-I: a presynaptic

calcium channel antagonist from the venom of the funnel-web

spider, Agelenopsis aperta. J. Neurobiol. 20, 171±188.

Blagbrough, I.S., Moya, E., 1994. Practical synthesis of the putative

polyamine spider toxin FTX: a proposed blocker of voltage-

sensitive calcium channels. Tetra. Letters 35, 2057±2060.

Bloomquist, J.R., Kinne, L.P., Deutsch, V., Simpson, S.F., 1996.

Mode of action of an insecticidal peptide toxin from the venom

of a weaving spider (Diguetia canities). Toxicon 34,

1072±1075.

Bormann, J., 1988. Electrophysiology of GABA

A

and GABA

B

receptor subtypes. Trends Neuro. Sci. 11, 112±116.

Bowers, C.W., Phillips, H.S., Lee, P., Jan, Y.N., Jan, L.Y., 1987.

Identi®cation and puri®cation of an irreversible presynaptic

neurotoxin from the venom of the spider Hololena curta.

Proc. Natl. Acad. Sci. USA 84, 3506±3510.

Branden, C., Tooze, J., 1991. Introduction to Protein Structure.

Garlan Publishers, New York.

Branton, W.D., Kolton, L., Jan, Y.N., Jan, L.Y., 1987. Neurotoxins

from Plectreurys spider venom are potent presynaptic blockers

in Drosophila. J. Neurosci. 7, 4195±4200.

Branton, W.D., Rudnick, M.S., Zhou, Y., 1992. A potent peptide

calcium channel blocker from Plectreurys spider venom

contains a C terminal lipid. Soc. Neurosci. Abs. 18, 10.

Braz, A., Minozzo, J., Abreu, J.C., Gubert, I.C., ChaÂvez-OloÂrtegui,

C., 1999. Development and evaluation of the neutralizing

capacity of horse antivenom against the Brazilian spider

Loxosceles intermedia. Toxicon 37, 1323±1328.

Bruno, L., 1990. Isolation of a galactose-binding lectin from the

venom of the snake Bothrops godmani. Toxicon 28, 75±81.

Bucherl, W., 1971. Spiders. In: Bucherl, W., Buckley, E.E. (Eds.).

Venomous Animals and their Benoms. Vol. III Venomous

Invertebrates. Academic Press, London, pp. 197±277.

Cabbiness, S.G., Gherke, C.W., Kuo, K.C., Chan, T.K., Hall, J.E.,

Hudiburg, S.A., Odell, G.V., 1980. Polyamines in some

tarantula venoms. Toxicon 18, 681±683.

Cavalieri, M., D'Urso, D., Lassa, A., Pierdominici, E., Robello, M.,

Grasso, A., 1987. Characterisation and some properties of the

venom gland extract of a therdiid spider (Steatoda paykulliana)

frequently mistaken for black widow spider (Latrodectus

tredecimguttatus). Toxicon 9, 965±974.

Ceccarelli, B., Grohovaz, F., Hurlbut, W.P., 1979. Freeze-fracture

studies of frog neuromuscular junctions during intense release of

neurotransmitter. I. Effects of black widow spider venom and

Ca

21

-free solutions on the structure of the active zone. J. Cell

Biol. 81, 163±177.

CeÂleÂrier, M.L., Paris, C., Lange, C., 1993. Venom of an aggressive

African theraphosidae (Scodra griseipes): Milking the venom, a

study of its toxicity and its characterization. Toxicon 31,

577±590.

Chan, T.K., Geren, C.R., Howell, D.E., Odell, G.V., 1975.

Adenosine triphosphate in tarantula spider venoms and its

synergistic effect with the venom toxin. Toxicon 13, 61±66.

Cohen, E., Quistad, G.B., 1998. Cytotoxic effects of arthropod

venoms on various cultured cells. Toxicon 36, 353±358.

Conley, E.C., Brammar, W.J., 1999. The Ion Channel Factsbook. 4,

Voltage-gated Channels. Academic Press, San Diego.

Cordeiro, M., do, N., Diniz, C.R., Valentim, A., do, C.,

von Eickstedt, V.R.D., Gilroy, J., Richardson, M., 1992. The

puri®cation and amino acid sequence of four Tx2 neurotoxins

from the venom of the Brazilian `armed' spider Phoneutria

nigriventer (Keys). FEBS Lett. 310, 153±156.

Cordeiro, M., do, N., de Figueiredo, S.G., Valentim, A., do, C.,

Diniz, C.R., von Eickstedt, V.R.D., Gilroy, J., Richardson, M.,

1993. Puri®cation and amino acid sequences of six Tx3 type

neurotoxins from the venom of the Brazilian `armed' spider

Phoneutria nigriventer (Keys). Toxicon 31, 35±42.

Croucamp, W., Veale, R., 1999. The venom of the sac spider

Cheiracanthium furculatum Karsch. Toxicon 37, 1335±1879.

Cruz, J.S., Cotta, G., Diniz, C.R., BeiraÄo, P.S.L., 1994. Partial

puri®cation and pharmacological characterisation of a

neurotoxic fraction isolated from the venom of the spider

Lycosa erythrognatha. Braz. J. Med. Biol. Res. 27, 2653±2659.

Cull-Candy, S.G., Neal, H., Usherwood, P.N.R., 1973. Action of

black widow spider venom on an aminergic synapse. Nature

241, 353±354.

D'Amour, F.E., Becker, F.E., van Riper, W., 1936. The black

widow spider. Quart. Rev. Biol. 11, 123±160.

Davletov, B.A., Krasnoperov, V., Hata, Y., Petrenko, A.G., SuÈdhof,

T.C., 1995. High af®nity binding of alpha-latrotoxin to

recombinant neurexin 1-alpha. J. Biol. Chem. 270, 23903±

23905.

Did-Hajj, S.D., Tyrrell, L., Cummins, T.R., Black, J.A., Wood,

P.M., Waxman, S.G., 1999. Two tetrodotoxin-resistant sodium

channels in human dorsal root ganglion neurons. FEBS Letters

462, 117±120.

Dingledine, R., Borges, K., Bowie, D., Traynelis, S.F., 1999. The

glutamate receptor ion channels. Pharmacol. Rev. 51, 7±61.

Diochot, S., Drici, M-D., Moinier, D., Fink, M., Lazdunski, M.,

1999. Effects of phrixotoxins on the Kv4 family of potassium

channels and implications for the role of I

to1

in cardiac

electrogenesis. Br. J. Pharmacol. 126, 251±263.

Duchen, L.W., Gomez, S., 1984. Pharmacology of spider venoms.

In: Tu, A.T. (Ed.). Insect Poisons, Allergens, and Other Inverte-

brate Venoms. Handbook of Natural Toxins. Marcel Dekker,

New York, pp. 483±512.

Duf®eld, P.H., Duf®eld, A.M., Carroll, P.R., Morgans, D., 1979.

Analysis of the venom of the Sydney funnel-web spider, Atrax

robustus using gas chromatography mass spectrometry.

Biomed. Mass. Spectrom. 6, 105±108.

Dupere, J.R.B., Moya, E., Blagbrough, I.S., Usowicz, M.M., 1996.

Differential inhibition of Ca

21

channels in mature rat cerebellar

Purkinje cells by sFTX-3.3 and FTX-3.3. Neuropharmacol. 35,

1±11.

Ertel, E.A., Waren, V.A., Adams, M.E., Grif®n, P.R., Cohen, C.J.,

Smith, M.M., 1994. Type-III v-agatoxins: A family of probes

for similar binding sites on L- and N-type calcium channels.

Biochem. 33, 5098±5108.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

246

background image

Escoubas, P., CeÂleÂrier, M.L., Romi-Lebrun, R., Nakajima, T., 1997.

Two novel peptide neurotoxins from the venom of the tarantula

Lasiodora parahybana. Toxicon 35, 805±806.

Escoubas, P., CeÂleÂrier, M.L., Nakajima, T., 1998. Composition and

neurotoxicity of tarantula venoms: biogeographic and

taxonomic variations. Toxicon 36, 1737.

Escoubas, P., Corzo, G., Whiteley, B.J., CeÂleÂrier, M.L., Nakajima,

T., 1999. HPLC and MALDI-TOF-MS study of quantitative and

qualitative variation in tarantula spider venoms. Proceedings of

the 5th Asia-Paci®c Congress on Animal, Plant and Microbial

Toxins, 37.

Escoubas, P., De Weille, J.R., Lecoq, A., Diochot, S., Waldmann,

R., Champigny, G., Moinier, D., Menez, A., Lazdunski, M.,

2000. Isolation of a tarantula toxin speci®c for a class of proton

gated Na 1 channels. J. Biol. Chem. 275, 25116±25121.

Escoubas, P., Diochot, S., Corzo, G., 2000. Structure and

pharmacology of spider venom neurotoxins. Biochimie 82,

893±907.

Feitosa, L., Gremski, W., Veiga, S.S., Elias, M.C.Q.B., Graner,

E., Mangili, O.C., Brentani, R.R., 1998. Detection and

characterization of metalloproteinases with gelatinolytic,

®brinectinolytic and ®brinogenolytic activities in brown

spider (Loxosceles intermedia) venom. Toxicon 36, 1039±

1051.

Fenton, A.W., West, P.R., Odell, G.V., Hudiburg, S.M., Ownby,

C.L., Mills, J.N., Scroggins, B.T., Shannon, S.B., 1995.

Arthropod venom citrate inhibits phospholipase A

2

. Toxicon

33, 763±770.

Ferreira, L.A.F., Henriques, O.B., 1992. Isolation of a kinin-like

peptide from the venom of Bothrops jararacussu. J. Nat. Toxins

1, 59±65.

Ferreira, L.A.F., Henriques, O.B., Lebrun, I., Batista, M.B.C.,

Prezoto, B.C., Andreoni, A.S.S., Zelnik, R., Habermehl, G.,

1992. A new bradykinin potentiating peptide (peptide P)

isolated from the venom of Bothrops jararcussu. Toxicon 30,

33±40.

Ferreira, L.A.F., Alves, E.W., Henriques, O.B., 1993. Peptide T, a

novel bradykinin potentiator isolated from Tityus serrulatus

scorpion venom. Toxicon 31, 941±947.

Ferreira, L.A.F., Mollring, T.H., Lebrun, F.L.A.S., Raida, M.,

Znottka, R., Habermehl, G.G., 1995. Structure and effects of a

kinin potentiating peptide F (AppF) isolated from Agkistrodon

piscivorus piscivorus venom. Toxicon 33, 1313±1319.

Ferreira, L.A.F., Alves, W.E., Lucas, M.S., Habermehl, G.G., 1996.

Isolation and characterization of a bradykinin potentiating

peptide (BPP-S) isolated from Scaptocosa raptoria venom.

Toxicon 34, 599±603.

Ferreira, L.A.F., Lucas, M.S., Alves, W.E., Hermann, V.V., Reichl,

A.P., Habermehl, G., Zingali, R.B., 1998. Isolation,

characterization and biological properties of two kinin-like

peptides (Peptide-S and Peptide-R) from Scaptocosa raptoria

venom. Toxicon 36, 31±39.

Ferreira, S.H., 1965. A bradykinin-potentiating factor (BPF) present

in the venom of Bothrops jararaca. Br. J. Pharmacol. 24,

164±168.

Figueiredo, S.G., de Lima, M.E., Cordeiro, M.N., Diniz, C.R.,

Patten, D., Halliwell, R.F., Gilroy, J., Richardson, M., 2001.

Puri®cation and amino acid sequence of a highly insecticidal

toxin from the venom of the Brazilian spider Phoneutria

nigriventer which inhibits NMDA-evoked currents in rat

hippocampal neurones. Toxicon 39, 309±317.

Figueiredo, S.G., Mafra, R.A., Pimenta, A.M.C., Cordeiro, M.N.,

Diniz, C.R., de Lima, M.E., 1997. Puri®cation and

pharmacological activity of two potent insecticidal neurotoxins

from Phoneutria nigriventer spider venom. J. Venom. Anim.

Toxins 3, 231.

Finkelstein, A., Rubin, L.L., Tzeng, M.C., 1976. Black widow

spider venom: Effect of puri®ed toxin on lipid bilayer

membranes. Science 193, 1009±1011.

Fischer, F.G., Bohn, H., 1957. Die giftsekrete der brasilianishen

tarantel Lycosa erythrognatha und der wanderspinne

Phoneteuria fera. Hoppe-Seyler's Z. Physiol. Chem. 306,

265±268.

Florey, E., 1967. Neurotransmitters and modulators in the animal

kingdom. Federation Proceedings 26, 1164±1178.

Forrester, L.J., Barrett, J.T., Campell, B.J., 1978. Red blood cell

lysis induced by the venom of the brown recluse spider: The role

of sphingomyelinase D. Arch. Biochem. Biophys. 187,

355±365.

Frew, R., Hamilton, M.G., Lundy, P.M., 1994. Identi®cation of

noradrenaline in venom from the funnel-web spider Hololena

curta. Toxicon 32, 511±515.

Fritz, L.C., Tzeng, M.-C., Mauro, A., 1980. Different components of

black widow spider venom mediate transmitter release at

vertebrate and lobster neuromuscular junctions. Nature 283,

486±487.

Frontali, N., Ceccarelli, B., Gorio, A., Mauro, A., Siekevitz, P.,

Tzeng, M., Hurlbut, W.P., 1976. Puri®cation from black

widow spider venom of a protein factor causing the depletion

of synaptic vesicles at neuromuscular junctions. J. Cell Biol. 68,

462±479.

Frontali, N., Granata, F., Parisi, P., 1972. Effects of black widow

spider venom on acetylcholine release from rat cerebral cortex

slices in vitro. Biochem. Pharmacol. 21, 969±974.

Futrell, J.M., 1992. Loxoscelism. Am. J. Med. Sci. 304, 261±267.

Garcia, M.L., Garcia-Calvo, M., Hidalgo, P., Lee, A., McKinnon,

R., 1994. Puri®cation and characterisation of three inhibitors of

voltage-dependent K

1

channels from Leiurus quinquestriatus

var. hebraeus venom. Biochem. 33, 6834±6839.

Garcia, M.L., Hanner, M., Knaus, H-G., Koch, R., Schmalhofer, W.,

Slaughter, R.S., Kaczorowski, G.J., 1997. Pharmacology of

potassium channels. Adv. Pharmacol. 39, 425±471.

Gates, C.A., Rees, R.S., 1990. Serum amyloid P component: Its role

in platelet activation stimulated by sphingomyelinase D puri®ed

from the venom of the brown recluse spider (Loxosceles

reclusa). Toxicon 28, 1303±1315.

Geppert, M., Khvochtev, M., Krasnoperov, V., Goda, Y., Missler,

M., Hammer, R.E., Ichtchenko, K., Petrenko, A.G., SuÈdhof,

T.C., 1998. Neurexin Ia is a major a-latrotoxin receptor that

cooperates in a-latrotoxin action. J. Biol. Chem. 273, 1705±

1710.

Geren, C.R., Odell, G.V., 1984. The biochemistry of spider venoms.

In: Insect poisons, allergens, and other invertebrate venoms. In:

Tu, A.T. (Ed.). Handbook of Natural Toxins, Vol. 2. Marcel

Dekker, New York, pp. 441±481.

Geren, C.R., Chan, T.K., Warb, B.C., Howell, D.E., Pinkston, K.,

Odell, G.V., 1973. Composition and properties of extract of

®ddleback (Loxosceles reclusa) spider venom apparatus.

Toxicon 11, 471±479.

Geren, C.R., Chan, T.K., Howell, D.E., Odell, G.V., 1976. Isolation

and characterization of toxins from brown recluse spider venom

(Loxosceles reclusa). Arch. Biochem. Biophys. 174, 90±99.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

247

background image

Gilbo, C.M., Coles, N.W., 1964. An investigation of certain

components of the venom of the female Sydney funnel-web

spider, Atrax robustus. Cambr. Aust. J. Biol. Sci. 17,

758±763.

Goldin, A.L., 1999. Diversity of mammalian voltage-gated sodium

channels. Ann. NY Acad. Sci. 868, 38±50.

Granja, R., Fernandez-Fernandez, J.M., Izaguirre, V., Gonzales-

Garcia, C., Cena, V., 1995. v-agatoxin IVA blocks nicotinic

receptor channels in bovine chromaf®n cells. FEBS Lett. 362,

15±18.

Gray, M., 1989. A signi®cant illness that was produced by the

white-tailed spider, Lampona cylindrata. Med. J. Aust. 151,

115±116.

Gray, M.R., Sutherland, S.K., 1978. Venoms of Dipluridae. In:

Bettini, S. (Ed.). Handbook of Experimental Pharmacology,

Vol. 48, Arthropod Venoms. Springer-Verlag, New York, pp.

121±148.

Greenamyre, J.T., Porter, R.H.P., 1994. Anatomy and physiology of

glutamate in the CNS. Neurology 44 (8), S7±S13.

GreÂgio, A.M.T., Heleno, M.G., von Eickstedt, V.R., Fontana, M.D.,

1999. The neuromuscular action of Ancylometes sp. spider

venom on the rat phrenic nerve-diaphragm preparation. Toxicon

37, 545±550.

Grishin, E.V., 1998. Black widow spider toxins: the present and the

future. Toxicon 36, 1671±1693.

Grishin, E.V., Volkova, T.M., Arseniev, A.S., Reshetova, O.S.,

Onoprienko, V.V., Magaznik, L.G., Antonov, S.M., Fedorova,

I.M., 1986. Structural-functional characteristics of argiopineÐ

an ion channel blocker from the venom of the spider Argiope

lobata. Biorg. Khim. 12, 1121±1124.

Guatimosin, G., Romano-Silva, M.A., Cruz, J.S., Beirao, P.S.L.,

Kalapothakis, E., Moraes-Santos, T., Cordeiro, M.N., Diniz,

C.R., Gomez, M.V., Prado, M.A.M., 1997. A toxin from the

spider Phoneutria nigriventer that blocks calcium channels

coupled to exocytosis. Br. J. Pharmacol. 122, 591±597.

Haeberli, S., Kuhn-Nentwig, L., Schaller, J., Nentwig, W., 2000.

Characterisation of antibacterial activity of peptides isolated

from the venom of the spider Cupiennius salei (Araneae:

Ctenidae). Toxicon 38, 373±380.

Hagiwara, K., Sakai, T., Miwa, A., Kawai, N., Nakajima, T., 1990.

Complete amino acid sequence of a new type of neurotoxin

from the venom of the spider, Agelena opulenta. Biomed.

Res. 1, 181±186.

Hagiwara, K., Inui, T., Nakajima, K., Kimura, T., Kitada, C.,

Fujino, M., Sakakibara, S., Nakajima, T., 1991. Agelenin, a

spider neurotoxin-determination of the C-terminal as amide

form and investigation of the disul®de bond arrangement.

Biomed. Res. 12, 357±363.

Hall, L.M., Ren, D., Feng, G., Eberl, D.F., Dubald, M., Yang, M.,

Hannan, F., Kousky, C.T., Zheng, W., 1994. Calcium channel as

a new potential target for insecticides. In: Clark, J.M. (Ed.).

ACS Symposium Series 591: Molecular action of Insecticides

on Ion Channels. American Chemical Society, San Diego, pp.

162±172.

Harvey, A.L., 1990. Presynaptic effects of toxins. Int. Rev.

Neurobiol. 32, 201±236.

Harvey, A.L., Anderson, A.J., 1991. Dendrotoxins: snake toxins

that block potassium channels and facilitate neurotransmitter

release. In: Harvey, A.L. (Ed.). Snake Toxins. Permagon Press

Inc, New York, pp. 131±164.

Harvey, M.S., Raven, R.J., 1991. Necrotising arachnidism in

Australia: a simple case of misidenti®cation. Med. J. Aust.

154, 856.

Hayman, J.A., Smith, I.M., 1991. Necrotising arachnidism in

Australia. Med. J. Aust. 155, 351.

Heitz, J.R., Norment, B.R., 1974. Characteristics of an alkaline

phosphatase activity in brown recluse venom. Toxicon 12,

181±187.

Henkel, A.W., Betz, W.J., 1995. Visualization of black widow

spider induced exo- and endocytosis at the frog neuromus-

cular junction with FM1-43. Neuropharmacology 34, 1397±

1406.

Henkel, A.W., Sankaranarayanan, S., 1999. Mechanisms of

a-latrotoxin action. Cell Tissue Res. 296, 229±233.

Herlitze, S., Raditsch, M., Ruppersberg, J.P., Jahn, W., Moyner, H.,

Schoepfer, R., Witzeman, V., 1993. Argiotoxin detects

molecular differences in AMPA receptor channels. Neuron 10,

1131±1140.

Hisada, M., Fujita, T., Naoki, H., Itagaki, Y., Irie, H., Miyashita, M.,

Nakajima, T., 1998. Structures of spider toxins: Hydroxyindole-

3-acetylpolyamines and a new generalized structure of type-E

compounds obtained from the venom of the joro spider, Nephila

clavata. Toxicon 36, 1115±1125.

Horni, A., Weickmann, D., Hesse, M., 2001. The main products of

the low molecular mass fraction in the venom of the spider

Latrodectus menavodi. Toxicon 39, 425±428.

Hufford, D.C., Morgan, P.N., 1981. C-reactive protein as a mediator

in the lysis of human erythrocytes sensitized by brown recluse

spider venom. Proc. Soc. Exp. Biol. Med. 167, 493±497.

Ibrahim, N., Morgan, M.F., Ahmed, M.R., 1989. Arachnidism: a

serious new Australian disease. Aust. N.Z. J. Surg. 59, 507±510.

Isbister, G., Gray, M., 2000. Acute and recurrent skin ulceration

after spider bite. Med. J. Aust. 172, 303.

Isbister, G., Gray, M., 2000. Prospective study of the clinical effects

of bites by white-tail spiders in Australia (Lampona cylindrata/

murina group). Proc. Aust. Soc. Clin. Exp. Pharmacol. Toxicol.

8, 15.

Jackson, H., Parks, T.N., 1989. Spider toxins: Recent applications in

neurobiology. Ann. Rev. Neurosci. 12, 405±414.

Jackson, H., Urnes, M., Gray, W.R., Parks, T.N., 1987. Effects

of spider venoms on transmission mediated by non-N-

methyl-D-aspartate receptors in the avian cochlear nucleus. In:

Hicks, T.P., Lodge, D., McLennan, H. (Eds.). Excitatory Amino

Acid Transmission. Liss, New York, pp. 51±54.

Jan, L.Y., Jan, Y.N., 1994. Potassium channels and their evolving

gates. Nature 371, 119±122.

Jentsch, T.J., Friedrich, T., Schriever, A., Yamada, H., 1999. The

ClC chloride channel family. P¯ugers-ArchivÐEur. J. Physiol.

437, 783±795.

Jong, Y.S., Norment, B.R., Heitz, J.R., 1979. Separation and

characterization of venom components in Loxosceles reclusa-II.

Protease enzyme activity. Toxicon 17, 529±537.

Kaire, G.H., 1963. Observation on some funnel-web spiders (Atrax

species) and their venoms, with particular reference to Atrax

robustus. Med. J. Aust. 2, 307±311.

Kaiser, E., 1956. Enzymatic activity of spider venoms. In: Buckley,

E.E., Porges, N. (Eds.). Venoms. American Association for the

Advancement of Science, Washington, DC, pp. 91±93.

Kaiser, E., Raab, W., 1967. Collagenolytic activity of snake and

spider venoms. Toxicon 4, 251±255.

Kaiser, I.I., Grif®n, P.R., Aird, S.D., Hudiburg, S., Shabano-

witz, J., Francis, B., John, T.R., Hunt, D.F., Odell, G.V.,

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

248

background image

1994. Primary structures of two proteins from the venom of

the Mexican red knee tarantula (Brachypelma smithii).

Toxicon 32, 1083±1093.

Kalapothakis, E., Penaforte, C.L., LeaÄo, R.M., Cruz, J.S., Prado,

V.F., Cordeiro, M.N., Diniz, C.R., Romano-Silva, M.A.,

Prado, M.A.M., Gomez, M.V., BeiraÄo, P.S.L., 1998. Cloning,

cDNA sequence analysis and patch clamp studies of a toxin

from the venom of the armed spider (Phoneutria nigriventer).

Toxicon 36, 1971±1980.

Kalapothakis, E., Penaforte, C.L., BeiraÄo, P.S.L., Romano-Silva,

M.A., Cruz, J.S., Prado, M.A.M., GuimaraÄes, P.E.M., Gomez,

M.V., Prado, V.F., 1998. Cloning of cDNAs encoding

neurotoxic peptides from the spider Phoneutria nigriventer.

Toxicon 36, 1843±1850.

Kashnikova, O.V., 1979. Poisonousness of the scorpion Buthus

eupeus. Izvestiya Akademii Nauk Kazakhskoi Ssr Seriya

Biologicheskaya 17, 30±35.

Kawai, N., Mauro, A., Grundfest, H., 1972. Effect of black widow

spider venom on the lobster neuromuscular junction. J. Gen.

Physiol. 60, 650±664.

Kawai, N., Miwa, A., Abe, T., 1982. Spider venom contains speci®c

receptor blockers of glutaminergic synapses. Brain Res. 247,

169±171.

Kawai, N., Miwa, A., Abe, T., 1982. Effect of a spider venom toxin

on glutaminergic synapses in mammalian brain. Biomed Res. 3,

353±355.

Kawai, N., Miwa, A., Abe, T., 1983. Speci®c antagonism of the

glutamate receptor by an extract from the venom of the spider

Araneus ventricosus. Toxicon 21, 438±440.

Kawai, N., Miwa, A., Shimazaki, K., Sahara, Y., Robinson, H.P.,

Nakajima, T., 1991. Spider toxin and the glutamate receptors.

Comp. Biochem. Physiol. 98C, 87±95.

Kemp, D.R., 1990. Inappropriate diagnosis of necrotizing

arachnidism. Watch out Miss MuffetÐbut don't get paranoid.

Med. J. Aust. 152, 671±699.

Kent, C.G., Tu, A.T., Geren, C.R., 1984. Isotachophoretic and

immunological analysis of venoms from sea snakes (Laticauda

semifasciata) and brown recluse spiders (Loxosceles reclusa) of

different morphology, locality, sex and developmental stages.

Comp. Biochem. Physiol. 77B, 303±311.

Kobayashi, K., Fischer, J.B., Knapp, A.G., Margolin, L., Daly, D.,

Reddy, N.L., Roach, B., McCormick, K., Meinwald, J., Goldin,

S.M., 1992. A novel polyamine calcium channel antagonist

isolated from venom of the spider Dolomedes okefenokiensis.

Soc. Neurosci. Abs. 18, 10.

Korszniak, N., Story, D.F., 1993. a- and b-adrenoceptor agonist

activity in the venom of the Australian huntsman spiders,

Delena cancerides and Isopeda montana. Clin. Exp. Pharmacol.

Physiol. 20, 127±134.

Krasnoperov, V.G., Beavis, R., Chenpurny, O.G., Plotnikov, A.N.,

Petrenko, A.G., 1996. The calcium-independent receptor of

a-latrotoxin is not a neurexin. Biochem. Biophys. Res.

Comm. 227, 868±875.

Kreil, G., 1994. Peptides containing a D-amino acid from frogs and

molluscs. J. Biol. Chem. 269, 10967±10970.

Kreil, G., 1995. HyaluronidasesÐA group of neglected enzymes.

Prot. Sci. 4, 1666±1669.

Kuhn-Nentwig, L., Nentwig, W., 1997. Venom of the hunting

spider Cupiennius salei (Ctenidae). Toxicon 35, 813.

Kuhn-Nentwig, L., Bucheler, A., Studer, W., 1998. Taurine and

histamine: Low molecular weight compounds in prey

hemolymph increase the killing power of spider venom.

Naturwissenschaften 85, 136±138.

Kuhn-Nentwig, L., Schaller, J., Nentwig, W., 1994. Puri®cation

of toxic peptides and the amino acid sequence of CSTX-1

from the multicomponent venom of Cupiennius salei (Araneae:

Ctenidae). Toxicon 32, 287±302.

Kurpiewski, G., Forrester, L.J., Barrett, J.T., Campell, B.T., 1981.

Platelet aggregation and sphingomyelinase D activity of a

puri®ed toxin from the venom of Loxosceles reclusa. Biochim.

et Biophys. Acta 678, 467±476.

Kushmerick, C., Kalapothakis, E., BeiraÄo, P.S.L., Penaforte, C.L.,

Prado, V.F., Cruz, J.S., Diniz, C.R., Cordeiro, M.N., Gomez,

M.V., Romano-Silva, M.A., Prado, M.A.M., 1999. Phoneutria

nigriventer toxin Tx3-1 blocks A-type K

1

currents controlling

Ca

21

oscillation frequency in GH

3

cells. J. Neurochem. 72,

1472±1481.

Kuwada, M., Teramoto, T., Kumagaye, K.Y., Watanabe, T., Kawai,

T., Kawakami, Y., Niidome, T., Sawada, K., Nishizawa, Y.,

Katayama, K., 1994. v-Agatoxin-TK containing D-serine at

position 46, but not synthetic v-[L-Ser46]agatoxin-TK, exerts

blockade of P-type calcium channels in cerebellar Purkinje

neurons. Mol. Pharmacol. 46, 587±593.

Lampe, R.A., Defeo, P.A., Davidson, M.D., Young, J., Herman, J.L.,

Spreen, R.C., Horn, M.B., Mangano, T.J., Keith, R.A., 1993.

Isolation and pharmacological characterisation of v-grammotoxin

SIA, a novel peptide inhibitor of neuronal voltage-sensitive

calcium channel responses. Mol. Pharmacol. 44, 451±460.

LeaÄo, R.M., Cruz, J.S., Diniz, C.R., Cordeiro, M.N., BeiraÄo, P.S.L.,

1997. Calcium channel blocking toxins in the venom of

Phoneutria nigriventer. J. Venom. Anim. Toxins 3, 223.

Lee, C.K., Chan, T.K., Ward, B.C., Howell, D.E., Odell, G.V.,

1974. The puri®cation and characterization of a necrotoxin

from tarantula, Dugesiella hentzi, venom. Arch. Biochem.

Biophys. 164, 341±350.

Lee, J.K., John, S.A., Weiss, J.N., 1999. Novel gating mechanism of

polyamine block in the strong inward recti®er K

1

channel

Kir2.1. J. Gen. Physiol. 113, 555±563.

Lee, R.Y.N., Lobel, L., Hengartner, M., Horvitz, H.R., Avery, L.,

1997. Mutations in the a1 subunit of an L-type voltage-

activated Ca

21

channel cause myotonia in Caenorhabditis

elegans. EMBO J. 16, 6066±6076.

Lelianova, V.G., Davletov, B.A., Sterling, A., Rahman, M.A.,

Grishin, E.V., Totty, N.F., Ushkaryov, Y.A., 1997. Alpha-

latrotoxin receptor, latrophilin, is a novel member of the secretin

family of G protein-coupled receptors. J. Biol. Chem. 72, 21504±

21508.

Leung, H.T., Branton, W.D., Phillips, H.S., Jan, L., Byerly, L.,

1989. Spider toxins selectively block calcium currents in

Drosophila. Neuron 6, 767±772.

Liang, S.P., Pan, X., 1995. A lectin-like peptide isolated from the

venom of the Chinese bird spider Selenocosmia huwena.

Toxicon 33, 875±882.

Liang, S.P., Chen, X.D., Shu, Q., Zhang, Y., Peng, K., 2000. The

presynaptic activity of huwentoxin-I, a neurotoxin from the

venom of the Chinese bird spider Selenocosmia huwena.

Toxicon 38, 1237±1246.

Liang, S.P., Zhang, D.Y., Pan, X., Chen, Q., Zhou, P.A., 1993.

Properties and amino acid sequence of huwentoxin-I, a

neurotoxin puri®ed from the venom of the Chinese bird spider

Selenocosmia huwena. Toxicon 31, 969±978.

Lin, J.W., Rudy, B., Llinas, R., 1990. Funnel-web spider venom and

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

249

background image

a toxin fraction block calcium current expressed from rat brain

mRNA in Xenopus oocytes. Proc. Natl. Acad. Sci. USA 87,

4538±4542.

Li-Smerin, Y.Y., Swartz, K.J., 1998. Gating modi®er toxins reveal a

conserved structural motif in voltage-gated Ca

21

and K

1

channels. Proc. Natl. Acad. Sci. USA 95, 8585±8589.

Liu, M., Nakazawa, K., Inoue, K., Ohno, Y., 1997. Potent and

voltage dependent block by philanthotoxin-343 of neuronal

nicotinic receptor/channels in PC12 cells. Br. J. Pharmacol.

122, 379±385.

Llinas, R., Sugimori, M., Lin, J.W., Cherskey, B., 1989. Blocking

and isolation of a calcium channel from neurons in mammals

and cephalopods utilizing a toxin fraction (FTX) from funnel-

web spider poison. Proc. Natl. Acad. Sci. USA 86, 1689±1693.

Llinas, R., Sugimori, M., Hillman, D.E., Cherskey, B., 1992.

Distribution and functional signi®cance of the P-type,

voltage-dependent Ca

21

channels in the mammalian central

nervous system. Trends Neuro. Sci. 15, 351±355.

Longnecker Jr, H.E., Hurlbut, W.P., Mauro, A., Clark, W.A., 1970.

Effects of black widow spider venom on the frog neuromuscular

junction. Nature 225, 701±705.

Macmillan, D.L., 1989. Envenomation by a window spider. Med.

J. Aust., 150.

McCormick, K.D., Meinwald, J., 1993. Neurotoxic acylpolyamines

from spider venoms. J. Ecol. Chem. 19, 2411±2451.

McDonough, S.I., Lampe, R.A., Keith, R.A., Bean, B.P., 1997.

Voltage-dependent inhibition of N- and P-type calcium

channels by the peptide toxin v-grammotoxin-SIA. Mol.

Pharmacol. 52, 1095±1104.

Macchiavello, A., 1947. Cutaneous arachnidism or gangrenous spot

of Chile. Pub. Health Trop. Med. 22, 425±466.

MacPherson, M.R., Pollock, V.P., Broderick, K.E., Kean, L.,

O'Connell, F.C., Dow, J.A.T., Davies, S.A., 2001. Model

organisms: New insights into ion channel and transporter

function. L-type calcium channels regulate epithelial ¯uid

transport in Drosophila melanogaster. Am. J. Physiol. Cell

Physiol. 280, C394±C407.

Mafra, R.A., Figueiredo, S.G., Diniz, C.R., Cordeiro, M.N., Cruz,

J.d.S., de Lima, M.E., 1999. PhTx4, a new class of toxins from

Phoneutria nigriventer spider venom, inhibits the glutamate

uptake in rat brain synaptosomes. Brain Res. 831, 297±300.

Majeski, J.A., Stinnett, J.D., Alexander, J.W., Durst Sr, G.G., 1977.

Action of venom from the brown recluse spider (Loxosceles

reclusa) on human neutrophils. Toxicon 15, 423±427.

Manzoli-Palma, M.F., Gobbi, N., Palma, M.S., 2000. The use of

insects as biological models to assay the toxicity of venoms

from spiders and scorpions. Toxicon 38, 578.

Marangoni, R.A., Antunes, E., Brain, S.D., de Nucci, G., 1993.

Activation by Phoneutria nigriventer (armed spider) venom of

tissue kallikrein-kininogen-kinin system in rabbit skin in vivo.

Br. J. Pharmacol. 109, 539±543.

Marangoni, S., Borges, N.C.C., Marangoni, R.A., Antunes, E.,

Vieira, C.A., Novello, J.C., Domont, G.B., Giglio, J.R.,

Oliveira, B., de Nucci, G., 1993. Biochemical characterization

of a vascular smooth muscle contracting polypeptide puri®ed

from Phoneutria nigriventer (armed spider) venom. Toxicon 31,

377±384.

Marban, E., Yamagishi, T., Tomaselli, G.F., 1998. Structure and

function of voltage-gated sodium channels. J. Physiol. 508,

647±657.

Maretic, Z., 1978. Venoms of Theridiidae, genus Latrodectus. In:

Bettini, S. (Ed.). Handbook of Experimental Pharmacology,

Vol. 48, Arthropod Venoms. Springer-Verlag, New York, pp.

185±212.

Maretic, Z., Levi, H.W., Levi, L.R., 1964. The theridiid spider

Steatoda paykulliana, poisonous to mammals. Toxicon 2,

149±154.

Marvin, L., De, E., Cosette, P., Gagnon, J., Molle, G., Lange, C.,

1999. Isolation, amino acid sequence and functional assays of

SGTx1: The ®rst toxin puri®ed from the venom of the spider

Scodra griseipes. Eur. J. Biochem. 265, 572±579.

Mattiello-Sverzut, A.C., Fontana, M.D., Diniz, C.R., Cruz-Hoȯing,

M.A., 1998. Pathological changes induced by PhTx1 from

Phoneutria nigriventer spider venom in mouse skeletal muscle

in vitro. Toxicon 36, 1349±1361.

Mattiello-Sverzuta, A.C., Cruz-Hoȯing, M.A., 2000. Toxin 2

(PhTx2), a neurotoxic fraction from Phoneutria nigriventer

spider venom, causes acute morphological changes in mouse

skeletal muscle. Toxicon 38, 793±812.

Matuszek, M.A., Hodgson, W.C., Sutherland, S.K., King, R.G.,

1992. Pharmacological studies of jumper ant (Myrmecia

pilosula) venom: evidence for the presence of histamine, and

haemolytic and eicosanoid-releasing factors. Toxicon 30, 1081±

1091.

Mebs, D., 1972. Proteolytic activity of a spider venom. In: DeVries,

A., Kochova, E. (Eds.). Toxins of Animal and Plant Origin.

Gordon and Breach, New York, pp. 493±497.

Minton, S.A., 1974. Venom Diseases. Charles C. Thomas

Publishing, Illinois.

Mintz, I.M., Venema, V.J., Adams, M.E., Bean, B.P., 1991.

Inhibition of N- and L-type Ca

21

channels by the spider

venom toxin v-aga-IIIA. Proc. Natl. Acad. Sci. USA 88,

6628±6631.

Mintz, I.M., Adams, M.E., Bean, B.P., 1992. Use of spider toxins to

discriminate between neuronal calcium channels. Soc.

Neurosci. Abs. 18, 9.

Mintz, I.M., Venema, V.J., Swiderek, K.M., Lee, T.D., Bean, B.P.,

Adams, M.E., 1992. P-type calcium channels blocked by the

spider toxin omega-aga IVA. Nature 355, 827±829.

Miranda, D.M., Romano-Silva, M.A., Kalapothakis, E., Diniz, C.R.,

Cordeiro, M.N., Santos, T.M., Prado, M.A., Gomez, M.V.,

1998. Phoneutria nigriventer toxins block tityustoxin-induced

calcium in¯ux in synaptosomes. Neuroreport 9, 1371±1373.

Morgan, P.N., 1969. Preliminary studies on venom from the brown

recluse spider Loxosceles reclusa. Toxicon 6, 161±165.

Morgans, D., Carroll, P.R., 1976. A direct acting adrenergic

component of the venom of the Sydney funnel-web spider,

Atrax robustus. Toxicon 14, 185±189.

Mota, I., Barbaro, K.C., 1995. Biological and biochemical

properties of venoms from medically important Loxosceles

(Araneae) species in Brazil. J. Toxicol.ÐToxin Rev. 14,

402±421.

Moya, E., Blaghbrough, I.S., 1994. Rapid, practical synthesis of the

arginyl polyamine sFTX3.3: a blocker of voltage-sensitive

calcium channels. Tetra. Letters 35, 2061±2062.

Nemeth, E.F., Mueller, A.L., Artman, L.D., Jackson, H., 1992.

Arylamimes derived from spider venoms are potent and

selective NMDA receptor antagonists in mammalian CNS.

In: Kawai, N., Nakajima, T., Barnard, E. (Eds.). Neurore-

ceptors, Ion Channels and the Brain. Elsevier, Amsterdam,

pp. 21±28.

Newcomb, R., Palma, A., Fox, J., Gaur, S., Lau, K., Chung, D.,

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

250

background image

Cong, R., Bell, J.R., Horne, B., Nadasdi, L., Ramachandran, J.,

1995. SNX-325, a novel calcium antagonist from the spider

Segestria ¯orentina. Biochem. 34, 8341±8347.

Newcomb, R., Szoke, B., Palma, A., Wang, G., Chen, X.,

Hopkins, W., Cong, R., Miller, J., Urge, L., Tarczy-

Hornoch, K., Loo, J.A., Dooley, D.J., Nadasdi, L.,

Tsien, R.W., Lemos, J., Miljanich, G., 1998. Selective peptide

antagonist of the class E calcium channel from the venom of the

tarantula Hysterocrates gigas. Biochem. 37, 15353±15362.

Newlands, G., 1982. Preliminary report on the medical importance

of Sicarius (Araneae: Sicariidae) and the action of its venom.

Mem. Inst. Butantan 46, 293±304.

Nicholson, G.M., Willow, M., Howden, M.E.H., Narahashi, T.,

1994. Modi®cation of sodium channel gating and kinetics by

versutoxin from the Australian funnel-web spider Hadronyche

versuta. P¯uÈgers ArchivÐEur. J. Physiol. 428, 400±409.

Nicholson, G.M., Little, M.J., Tyler, M., Narahashi, T., 1996.

Selective alteration of sodium channel gating by Australian

funnel-web spider toxins. Toxicon 34, 1443±1453.

Nicholson, G.M., Walsh, R., Little, M.J., Tyler, M.I., 1998.

Characterisation of the effects of robustoxin, the lethal

neurotoxin from the Sydney funnel-web spider Atrax robustus,

on sodium channel activation and inactivation. P¯uÈgers

ArchivÐEur. J. Physiol. 436, 117±126.

Norment, B.R., Foil, L.D., 1979. Histopathology and physiological

action of venom from the brown recluse spider Loxosceles

reclusa. Toxicon 17 (1), 131.

Norment, B.R., Jong, Y.S., Heitz, J.R., 1979. Separation and

characterization of venom components in Loxosceles

reclusa-III. Hydrolytic enzyme activity. Toxicon 17, 539±548.

Ogilvie, M.L., Dockter, M.E., Wenz, L., Gartner, T.K., 1986.

Isolation and characterization of lactose-binding lectins from

the venoms of the snake Lachesis muta and Dendroaspis

jamesonii. J. Biochem. 100, 1425±1431.

Olivera, B.M., Miljanich, G.P., Ramachandran, J., Adams, M.E.,

1994. Calcium channel diversity and neurotransmitter release.

Annu. Rev. Biochem. 63, 823±867.

de Oliveira, K.C., Goncalves de Andrade, R.M., Guisti, A.L., Dias

da Silva, W., Tambourgi, D.V., 1999. Sex-linked variation of

Loxoceles intermedia spider venoms. Toxicon 37, 217±221.

Omecinsky, D.O., Holub, K.E., Adams, M.E., Reily, M.D., 1996.

Three-dimensional structure analysis of m-agatoxins: further

evidence for common motifs among neurotoxins with diverse

ion channel speci®cities. Biochemistry 35, 2836±2844.

Opie, L.H., Kowolik, H., 1995. The discovery of captopril: from

large animals to small molecules. Cardiovasc. Res. 30, 18±25.

Ornberg, R.L., Smyth, T., Benton, A.W., 1976. Isolation of a

neurotoxin with a presynaptic action from the venom of the

black widow spider (Latrodectus mactans, Fab.). Toxicon 14,

329±333.

Osborne, R.H., 1996. Insect neurotransmission: Neurotransmitters

and their receptors. Pharmacol. Ther. 69, 117±142.

Pansa, M.C., Migliori Natalizi, G., Bettini, S., 1972. 5-Hydr-

oxytryptamine content of Latrodectus mactans tredecimguttatus

venom from gland extracts. Toxicon 10, 85±86.

Parks, T.N., Mueller, A.L., Artman, L.D., Albemsi, B.C., Nemeth,

E.F., Jackson, H., Jasys, V.J., Saccomano, N.A., Volkmann,

R.A., 1991. Arylamine toxins from funnel-web spider

(Agelenopsis aperta) venom antagonize N-methyl-D-aspartate

receptor function in mammalian brain. J. Biol. Chem. 266,

21523±21529.

Peng, K., Chen, X-D., Liang, S-P., 2001. The effect of huwentoxin-I

on Ca

21

channels in differentiated NG108-15 cells, a patch-

clamp study. Toxicon 39, 491±498.

Perret, B.A., 1977. Proteolytic activity of tarantula venoms due to

contamination with saliva. Toxicon 15, 505±510.

Petrenko, A.G., Kovalenko, V.A., Shamotienko, O.G., Surkova,

I.N., Tarasyuk, T.A., Ushkaryov, Y.A., Grishin, E.V., 1990.

Isolation and properties of the alpha-latrotoxin receptor.

EMBO J. 9, 2023±2037.

Petrenko, A.G., Perin, M.S., Davletov, B.A., Ushkaryov, Y.A.,

Geppert, M., SuÈdhof, T.C., 1991. Binding of synaptotagmin to

the a-latrotoxin receptor implicates both in synaptic exocytosis.

Nature 353, 65±68.

Pfeiffer-Linn, C., Glantz, R.M., 1991. An arthropod NMDA

receptor. Synapse 9, 35±42.

Pincus, S.J., Winkel, K.D., Hawdon, G.M., Sutherland, S.K., 1999.

Acute and recurrent skin ulceration after spider bite. Med.

J. Aust. 171, 99±102.

Pinto, J.E.B., Rothlin, R.P., Dagrosa, E.E., 1974. Noradrenaline

release by Latrodectus mactans venom in guinea-pig atria.

Toxicon 12, 385±393.

Piser, T.M., Lampe, R.A., Keith, R.A., Thayer, S.A., 1995.

v-Grammotoxin SIA blocks multiple, voltage-gated, Ca

21

channel subtypes in cultured rat hippocampal neurons. Mol.

Pharmacol. 48, 131±139.

Platnick, N.I., 1993. Advances in Spider Taxonomy, 1988±1991.

The New York Entomological Society, New York.

Platnick, N.I., 2000. A relimitation and revision of the Australasian

ground spider family Lamponidae (Araneae: Gnaphosoidea).

Bull. Am. Mus. Nat. Hist. 245, 1±330.

Priestly, T., Woodruff, G.N., Kemp, T.A., 1989. Antagonism of

responses to excitatory amino acids on rat cortical neurones

by the spider toxin, argiotoxin

636

. Br. J. Pharmacol. 97,

1315±1323.

Qu, Y.X., Liang, S.P., Ding, J.Z., Ma, L.B., Zhang, R.J., Gu, X.C.,

1995. Proton nuclear magnetic resonance studies on

Huwentoxin-I from the venom of the spider S. huwena. 1.

Sequence-speci®c H-NMR assignment. J. Prot. Chem. 14,

549±557.

Qu, Y.X., Liang, S.P., Ding, J.Z., Liu, X.C., Zhang, R.J., Gu, X.C.,

1997. Proton nuclear magnetic resonance studies on

Huwentoxin-I from the venom of the spider S. huwena. 2.

Three-dimensional structure. J. Prot. Chem. 16, 565±574.

Quistad, G.B., Dennis, P.A., Skinner, W.S., 1992. Insecticidal

activity of spider (Araneae), centipede (Chilopoda), scorpion

(Scorpionida), and snake (Serpentes) venoms. J. Econ.

Entomology 85, 33±39.

Quistad, G.B., Reuter, C.C., Skinner, W.S., Dennis, P.A.,

Suwanrumpha, S., Fu, E.W., 1991. Paralytic and insecticidal

toxins from the funnel web spider Hololena curta. Toxicon

29, 329±336.

Rash, L.D., King, R.G., Hodgson, W.C., 1998. Evidence that

histamine in the principal pharmacological component of

venom from an Australian wolf spider (L. godeffroyi). Toxicon

36, 367±375.

Rash, L.D., Birinyi-Strachan, L.C., Nicholson, G.M., Hodgson,

W.C., 2000. Neurotoxic activity of venom from the Austra-

lian Eastern mouse spider (Missulena bradleyi) involves

modulation of sodium channel gating. Br. J. Pharmacol.

130, 1817±1824.

Rash, L.D., King, R.G., Hodgson, W.C., 2000. Sex differences in

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

251

background image

the pharmacological activity of venom from the white-tailed

spider (Lampona cylindrata). Toxicon 38, 1111±1127.

Raven, R.J., Harvey, M.S., 1991. Necrotising arachnidism in

Australia. Med. J. Aust. 155, 208.

Rees, R.S., 1989. Platelet activation stimulated by the toxin of the

brown recluse spider requires serum amyloid P component, not

C-reactive peptide. Toxicon 27, 953±954.

Rees, R.S., Gates, C., Timmons, S., Des Prez, R.M., King Jr, L.E.,

1988. Plasma components are required for platelet activation by

the toxin of Loxosceles reclusa. Toxicon 26, 1035±1045.

Rego, E., Bento, A.C., Lopes-Martins, R.A.B., Antunes, E.,

Novello, J.C., Marangoni, S., Giglio, J.R., Oliveira, B.,

de Nucci, G., 1996. Isolation and partial characterization of a

polypeptide from Phoneutria nigriventer spider venom that

relaxes rabbit corpus cavernosum in vitro. Toxicon 34,

1141±1147.

Rekow, M.A., Civello, D.J., Geren, C.R., 1983. Enzymatic and

hemolytic properties of brown recluse spider (Loxosceles

reclusa) toxin and extracts of venom apparatus, cephalothorax

and abdomen. Toxicon 21, 441±444.

Rezende Jr., L., Cordeiro, M.N., Oliveira, E.B., Diniz, C.R., 1991.

Isolation of neurotoxic peptides from the venom of the `armed'

spider Phoneutria nigriventer. Toxicon 29, 1225±1233.

Robertson, B., 1997. The real life of voltage-gated K

1

channels:

more than model behaviour. Trends Pharmacol. Sci. 18,

474±483.

Russell, F.E., 1966. Phosphodiesterase of some snake and arthropod

venoms. Toxicon 4, 153±154.

Russell, F.E., Gertsch, W.J., 1983. Letter to the editor. Toxicon 21,

337±339.

Sanguinetti, M.C., Johnson, J.H., Hammerland, L.G., Kelbaugh,

P.R., Volkmann, R.A., Saccomano, N.A., Mueller, A.L.,

1997. Heteropodatoxins: Peptides isolated from spider

venom that block Kv4.2 potassium channels. Mol. Pharma-

col. 51, 491±498.

Savel-Neimann, A., 1989. Tarantula (Eurypelma californicum)

venom, a multicomponent system. Hoppe-Seyler Biol. Chem.

370, 485±498.

Schachter, M., Thain, E.M., 1954. Chemical and pharmacological

properties of the potent, slow contracting substance (kinin) in

wasp venom. Br. J. Pharmacol. 9, 352±359.

Schanbacher, F.L., Lee, C.K., Hall, J.E., Wilson, I.B., Howell, D.E.,

Odell, G.V., 1973. Composition and properties of tarantula

Dugesiella hentzi (Girard) venom. Toxicon 11, 21±29.

Schanbacher, F.L., Lee, C.K., Wilson, I.B., Howell, D.E., Odell,

G.V., 1973. Puri®cation and characterisation of tarantula,

Dugesiella hentzi (Girard) venom hyaluronidase. Comp.

Biochem. Physiol. 44B, 389±396.

Schenone, H., Suarez, G., 1978. Venoms of Scytodidae. Genus

Loxosceles. In: Bettini, S. (Ed.). Handbook of Experimental

Pharmacology, Vol. 48, Arthropod Venoms. Springer-Verlag,

New York, pp. 247±275.

Schmaus, L.F., 1929. Case of arachnidism (spider bite). JAMA 92,

1265±1266.

Schulz, S., 1997. The chemistry of spider toxins and silk. Angew.

Chem. Int. Ed. Engl. 36, 314±326.

Schweitz, H., Bruhn, T., Guillemare, E., Moinier, D., Lancelin, J-M.,

Beress, L., Lazdunski, M., 1995. Kalicludines, Kaliseptine: two

different classes of sea anemone toxins for voltage-sensitive K

1

channels. J. Biol. Chem. 270, 25121±25126.

Scott, R.H., Dolphin, A.C., Bindokas, V.P., Adams, M.E., 1990.

Inhibition of neuronal Ca

21

channel currents by the funnel-web

spider toxin v-Aga-IA. Mol. Pharmacol. 38, 711±718.

Scott, R.H., Thatcher, N., Ayar, A., Mitchell, S.J., Pollock, J.,

Gibson, M.T., Duce, I.R., Moya, E., Blagbrough, I.S., 1998.

Extracellular or intracellular application of argiotoxin-636 has

inhibitory actions on membrane excitability and voltage-

activated currents in cultured rat sensory neurones.

Neuropharmacology 37, 1563±1578.

Sheumack, D.D., Baldo, B.A., Carroll, P.R., Hampson, F., Howden,

M.E.H., Skorulis, A., 1984. A comparative study of properties

and toxic constituents of funnel web spider (Atrax) venoms.

Comp. Biochem. Physiol. 78C, 55±68.

Shikata, Y., Watanabe, T., Teramoto, T., Inoue, A., Kawakami, Y.,

Nishizawa, Y., Katayama, K., Kuwada, M., 1995. Isolation and

characterization of a peptide isomerase from funnel-web spider

venom. J. Biol. Chem. 270, 16719±16723.

Shu, Q., Liang, S.P., 1999. Puri®cation and characterization of

huwentoxin-II, a neurotoxic peptide from the venom of the

Chinese bird spider Selenocosmia huwena. J. Peptide Res. 53,

486±491.

von Sicard, N.A.E., Candy, D.J., Anderson, M., 1989. The

biochemical composition of venom from the pavement ant

(Tetramorium caespitum L.). Toxicon 27, 1127±1133.

Skinner, M.W., Butler, C.S., 1995. Necrotising arachnidism treated

with hyperbaric oxygen. Med. J. Aust. 162, 372±373.

Skinner, W.S., Dennis, P.A., Li, J.P., Quistad, G.B., 1992.

Identi®cation of insecticidal peptides from venom of the

trap-door spider, Aptostichus schlingeri (Ctenizidae). Toxicon

30, 1043±1050.

Skinner, W.S., Dennis, P.A., Lui, A., Carney, R., Quistad, G.B.,

1990. Chemical characterisation of acylpolyamine toxins from

venom of a trap-door and two tarantulas. Toxicon 28, 541±546.

Smith, C.W., Micks, D.W., 1970. The role of polymorphonuclear

leucocytes in the lesion caused by the venom of the brown

recluse spider, Loxosceles reclusa. Lab. Invest. 22, 90±93.

Sosnina, N.A.A., Golubenco, Z., Akhunov, A.A., Kugaevskaya,

E.V., Yu, E., Orekhovich, V.N., 1990. Bradykinin potentiating

peptides of the spider Latrodectus tredecimguttatus, inhibitors

of carboxicathepsin and a preparation of karakurt spider venom

kininase. Dokl. Akad. Nauk. SSSR 315, 281±284.

Spring, W.J., 1987. A probable case of necrotizing arachnidism.

Med. J. Aust. 147, 605±607.

Stapleton, A., Blankenship, D.T., Ackermann, B.L., Chen, T.M.,

Gorder, G.W., Manley, G.D., Palfreyman, M.G., Coutant, J.E.,

Cardin, A.D., 1990. Curtatoxins: Neurotoxic insecticidal

polypeptides isolated from the venom of the funnel-web spider

Hololena curta. J. Biol. Chem. 265, 2054±2059.

Suarez, G., Schenone, H., Socias, T., 1971. Loxosceles laeta

venom-partial puri®cation. Toxicon 17, 291.

Sugita, S., Khvochtev, M., SuÈdhof, T.C., 1999. Neurexins are

functional a-latrotoxin receptors. Neuron 22, 489±496.

Sutherland, S.K., 1978. Venoms of Dipluridae. In: Bettini, S. (Ed.).

Handbook of Experimental Pharmacology, Vol. 48, Arthropod

Venoms. Springer-Verlag, New York, pp. 121±148.

Sutherland, S.K., 1983. Spiders. Australian Animal Toxins: the

creatures, their venom and care of the poisoned patient Oxford

University Press, Melbourne, pp. 222±241.

Sutherland, S.K., 1983. Spider bites in Australia: There are still

some mysteries. Med. J. Aust. 2, 597.

Sutherland, S.K., 1987. Watch out Miss Muffet!. Med. J. Aust. 147,

531.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

252

background image

Sutherland, S.K., 1990. Inappropriate diagnosis of necrotising

arachnidism. Med. J. Aust. 153, 499.

Sutherland, S.K., 1991. Necrotising arachnidism in Australia. Med.

J. Aust. 155, 136.

Sutko, J.L., Airey, J.A., Welch, W., Ruest, L., 1997. The

pharmacology of ryanodine related compounds. Pharmacol.

Rev. 49, 53±98.

Sutton, K.G., Siok, C., Stea, A., Zamponi, G.W., Heck, S.D.,

Volkmann, R.A., Ahlijanian, K., Snutch, T.P., 1998. Inhibition

of neuronal calcium channels by a novel peptide spider toxin,

DW13.3. Mol. Pharmacol. 54, 407±418.

Sutton, K.G., Stapleton, S.R., Scott, R.H., 1998. Inhibitory actions

of synthesised polyamine spider toxins and their analogues on

Ca

21

-activated Cl

2

currents recorded from cultured DRG

neurones from neonatal rats. Neurosci. Letters 251, 117±120.

Swartz, K.J., MacKinnon, R., 1995. An inhibitor of the Kv2.1

potassium channel isolated from the venom of a Chilean

tarantula. Neuron 15, 941±949.

Tambourgi, D.V., Magnoli, F.C., Petricevich, V.L., Sucupira, M.,

Benedetti,

Z.C.,

Assaf,

S.L.M.R.,

Alves,

E.W.,

von Eickstedt, V.R.D., Jancar, S., Dias da Silva, W., 1997.

Molecular mechanisms involved in the in¯ammation induced

by Loxosceles spider venoms. J. Venom. Anim. Toxins 3, 115.

Tzeng, M.C., Siekevitz, P., 1978. The effect of the puri®ed protein

factor (a-latrotoxin) of black widow spider venom on the

release of acetylcholine and norepinephrine from mouse

cerebral cortex slices. Brain Res. 139, 190±196.

Uchitel, O.D., 1997. Toxins affecting calcium channels in neurons.

Toxicon 35, 1161±1191.

Usherwood, P.N.R., Blagbrough, I.S., 1991. Spider toxins affecting

glutamate receptors: polyamines in therapeutic neurochemistry.

Pharmacol. Ther. 52, 245±268.

Usherwood, P.N.R., Blagbrough, I.S., 1994. Electrophysiology of

polyamines and polyamine amides. In: Carter, C. (Ed.).

Neuropharmacology of Polyamines. Academic Press, London,

pp. 185±204.

Usherwood, P.N.R., Blagbrough, I.S., Brackley, P.T.H., Kerry, C.J.,

Sudan, A.L., Nakanishi, K., 1992. Polyamines and polyamine

containing toxins, modulators and antagonists of excitatory

amino acid receptors. In: Kawai, N., Nakajima, T., Barnard,

E. (Eds.). Neuroreceptors, Ion Channels and the Brain. Elsevier,

Amsterdam, pp. 11±20.

Usherwood, P.N.R., Duce, I.R., Boden, P., 1984. Slowly-reversible

block of glutamate receptor channels by venoms of the spiders

Argiope trifasciata and Araneus gemma. J. Physiol. 79,

241±245.

Ushkaryov, Y.A., Petrenko, A.G., Geppert, M., SuÈdhof, T.C., 1992.

Neurexins: synaptic cell surface proteins related to the

alpha-latrotoxin receptor and laminin. Science 257, 50±56.

Usmanov, P.B., Nuritova, F.A., 1994. The anti coagulant action of a

phospholipase A from Ereus niger spider venom. Toxicon 32,

625±628.

Usmanov, P.B., Kalikulov, D., Shadyeva, N.G., Tashmukhamedov,

B.A., 1983. Effects of Argiope lobata spider venom on the

glutamatergic and cholinergic synapses. Dokl. Akad. Nuak

SSSR 273, 1017.

Usmanov, P.B., Kalikulov, D., Shadyeva, N.G., Nenilin, A.B.,

Tashmukhamedov, B.A., 1985. Postsynaptic blocking of

glutamatergic and cholinergic synapses as a common property

of Araneidae spider venoms. Toxicon 23, 528±531.

Van Aswegen, G., Van Rooyen, J.M., Van der Nest, D.G., Veldman,

F.J., de Villiers, T.H., Oberholzer, G., 1997. Venom of a

six-eyed crab spider, Sicarius testaceus (Purcell, 1908), causes

necrotic and haemorrhagic lesions in rabbits. Toxicon 35,

1149±1152.

Veiga, S.S., da Silveira, R.R., Dreyfuss, J.L., Haoach, J., Pereira,

A.M., Mangili, O.C., Gremski, W., 2000. Identi®cation of high

molecular weight serine-proteases in Loxosceles intermedia

(brown spider) venom. Toxicon 38, 825±839.

Venema, V.J., Swiderek, K.M., Lee, T.D., Hathaway, G.M., Adams,

M.E., 1992. Antagonism of synaptosomal calcium channels by

subtypes of v-agatoxins. J. Biol. Chem. 267, 2610±2615.

Vest, D.K., 1987. Envenomation by Tegenaria agrestis

(Walckenaer) spider in rabbits. Toxicon 25, 221±224.

Vest, D.K., 1987. Necrotic arachnidism in the north-west United

States and its probable relationship to Tegenaria agrestis

(Walckenaer) spiders. Toxicon 25, 175±184.

Wang, X., Smith, R., Fletcher, J.I., Wilson, H., Wood, C.J.,

Howden, M.E.H., King, G.F., 1999. Structure-function studies

of v-atracotoxin, a potent antagonist of insect voltage-gated

calcium channels. Eur. J. Biochem. 264, 488±494.

Ward, P.A., Cochrane, C.G., 1965. Bound complement and

immunologic injury of blood vessels. J. Exp. Med. 121,

215±234.

Wiener, S., 1961. Observations on the venom of the Sydney

funnel-web spider (A. robustus). Venom. In: Keegan, H.L.,

MacFarlane, W.V. (Eds.). Venomous and Poisonous Animals

and Noxious Plants of the Paci®c Region. Permagon Press,

New York, pp. 141±151.

Welsh, J.H., Batty, C.S., 1963. 5-Hydroxytryptamine content of

some arthropod venoms and venom-containing parts. Toxicon

1, 165±173.

White, J., 1987. Review of clinical and pathological aspects of

spider-bite in Australia. In: Gopalakrishnakone, P., Tan, C.K.

(Eds.). Progress in Venom and Toxin Research. Faculty of

Medicine, National University of Singapore, Singapore, pp.

531±541.

White, J., Cardoso, J.L., Fan, H.W., 1995. Clinical toxicology of

spider bite. In: Meier, J., White, J. (Eds.). Handbook of Clinical

Toxicology of Animal Venoms and Poisons. CRC Press, Boca

Raton. pp. 259±329.

White, J., 1999. Necrotising arachnidism: Does the white-tailed

spider deserve its bad name? Med. J. Aust. 171, 98.

White, J., Hirst, D., Hender, E., 1989. 36 cases of bites by spiders,

including the white-tailed spider, Lampona cylindrata. Med.

J. Aust. 150, 401±403.

Woo, K.M.L., Smart, D.R., 1999. The clinical effects of spider bites

in southern Tasmania. Emerg. Med. 11, 145±149.

Wright, R.P., Elgert, K.D., Campell, B.J., Barrett, J.T., 1973.

Hyaluronidase and esterase activities of the venom of the

poisonous brown recluse spider. Arch. Biochem. Biophys.

159, 415±426.

Xu, K., Ji, Y., Qu, X., 1989. Puri®cation and characterisation of an

antibacterial peptide from venom of Lycosa singoriensis. Acta

Zoologica Sinica 35, 300±305.

Yan, L., Adams, M.E., 1998. Lycotoxins, antimicrobial peptides

from venom of the wolf spider Lycosa carolinensis. J. Biol.

Chem. 273, 2059±2066.

Yasahura, T., Hira, M., Nakajima, T., Yanaihira, N., Yanaihira, C.,

Hashimoto, T., Sakura, N., Tachibana, S., Araki, K., Bessho,

M., Yamanaka, M., 1973. Active peptides on smooth muscle in

the skin of Bombina orientalis Boulenger and characterization

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

253

background image

of a new bradykinin analogue. Chemical and Pharmaceutical

Bulletin 21, 1388±1394.

Young, A.R., Pincus, S.J., 2001. Comparison of enzymatic activity

from three species of necrotising arachnids in Australia:

Loxosceles rufescens, Badumna insignis and Lampona

cylindrata. Toxicon 39, 391±400.

Zhang, D.Y., Liang, S.P., 1993. Assignment of the disulphide

bridges of huwentoxin-I, a neurotoxin from the spider

Selenocosmia huwena. J. Prot. Chem. 12, 735±740.

Zhou, P.A., Xie, X.-J., Li, M., Yang, D.M., Xie, Z.P., Zong, X.,

Liang, S.P., 1997. Blockade of neuromuscular transmission by

huwentoxin-I, puri®ed from the venom of the Chinese bird

spider Selenocosmia huwena. Toxicon 35, 39±45.

Zheng, W., Feng, G., Ren, D., Eberl, D.F., Hannan, F., Dubald, M.,

Hall, L.M., 1995. Cloning and characterization of a calcium

channel a

1

subunit from Drosophila melanogaster with

similarity to the rat brain type D isoform. J. Neurosci. 15,

1132±1143.

Zlotkin, E., 1999. The insect voltage-gated sodium channel as target

of insecticides. Annu. Rev. Entomol. 44, 429±455.

L.D. Rash, W.C. Hodgson / Toxicon 40 (2002) 225±254

254


Wyszukiwarka

Podobne podstrony:
Historia książki 4
Krótka historia szatana
Historia Papieru
modul I historia strategii2002
Historia turystyki na Swiecie i w Polsce cz 4
Historia elektroniki
Historia książki
historia administracji absolutyzm oświecony
Psychologia ogólna Historia psychologii Sotwin wykład 7 Historia myśli psychologicznej w Polsce
Historia hotelarstwa wukład
1Wstep i historia 2id 19223 ppt
Historia europejskiej integracji
Historia Prawa Publicznego (1)
materialy na zajecia historia sejmu staropolskiego

więcej podobnych podstron