Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 271, No. 13, Issue of March 29, pp.

7522–7528, 1996
© 1996 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

A New a-Conotoxin Which Targets a3b2 Nicotinic Acetylcholine


Receptors*
(Received for publication, November 13, 1995, and in revised form, January 8, 1996)

G. Edward Cartier‡, Doju Yoshikami‡, William R. Gray‡, Siqin Luo‡,


Baldomero M. Olivera‡, and J. Michael McIntosh‡§¶
From the Departments of ‡Biology and §Psychiatry, University of Utah, Salt Lake City, Utah 84112

We have isolated a 16-amino acid peptide from the (2). However, the precise structural composition and functional
venom of the marine snail Conus magus which potently role of the different neuronal subtypes of nicotinic receptors are
blocks nicotinic acetylcholine receptors (nAChRs) com- less well understood. The development of subtype-specific li-
posed of a3b2 subunits. This peptide, named a-cono- gands will greatly aid progress in this area.
toxin MII, was identified by electrophysiologically Although a number of valuable nicotinic antagonists have
screening venom fractions against cloned nicotinic re- been described, few are highly subtype-selective, particularly
ceptors expressed in Xenopus oocytes. The peptide’s in the case of neuronal nAChRs. d-Tubocurarine, an alkaloid
structure, which has been confirmed by mass spectrom- from the Chondrodendron tomentosum bush, used for centuries
etry and total chemical synthesis, differs significantly
as an arrow poison to kill wild game, blocks both muscle and
from those of all previously isolated a-conotoxins. Disul-

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


neuronal nAChRs (3). In addition it binds to all neuronal nic-
fide bridging, however, is conserved. The toxin blocks
the response to acetylcholine in oocytes expressing a3b2 otinic receptors with more or less similar affinities (4). Like-
nAChRs with an IC50 of 0.5 nM and is 2– 4 orders of wise, dihydro-b-erythroidine, the hydrogenated derivative of
magnitude less potent on other nAChR subunit combi- erythroidine, isolated from trees and shrubs of the genus
nations. We have recently reported the isolation and Erythrina is a competitive antagonist at both muscle and neu-
characterization of a-conotoxin ImI, which selectively ronal nAChRs (4). Lophotoxin, a small cyclic diterpene, is used
targets homomeric a7 neuronal nAChRs. Yet other by the soft shell coral Lophogorgia rigida to discourage its
a-conotoxins selectively block the muscle subtype of consumption by fish (5). This toxin forms a covalent bond with
nAChR. Thus, it is increasingly apparent that a-conotox- Tyr190 of the a-subunit of Torpedo nAChRs, irreversibly block-
ins represent a significant resource for ligands with ing the binding of ACh to the receptor (6, 7). Studies with
which to probe structure-function relationships of var- nAChRs expressed in Xenopus oocytes reveal that this toxin
ious nAChR subtypes. blocks muscle nAChRs as well as a2b2, a3b2, and a4b2 neu-
ronal nAChRs (8). Neosurugatoxin, a glycoside from the gas-
tropod Babylonia japonica (9), potently but nonselectively
The muscle subtype of nicotinic acetylcholine receptor blocks axb2 nAChRs expressed in oocytes, where x is 2, 3, or 4
(nAChR)1 is one of the best understood ligand-gated channels (8). The synthetically derived small molecules trimethaphan
due in part to the availability of a large number of protein and and mecamylamine discriminate between ganglionic and neu-
small molecule ligands which serve as specific probes for this romuscular nAChRs and are used clinically as ganglionic block-
channel. The nAChR is a heteropentameric ion channel com- ing agents (3).
plex and is a member of a superfamily that includes glycine, Numerous protein toxins which act at muscle nAChRs have
GABAA, and 5-HT3 receptors (1). The mammalian nAChR has been isolated from a variety of snake venoms and proven highly
the subunit composition (a1)2b1gd in developing muscle, and useful for studying nAChRs. Two toxins from the Taiwanese
the g subunit is replaced by an e subunit in mature muscle. In banded krait, Bungarus multicinctus, have been particularly
mammalian neurons the situation is much more complex with well characterized. The major nicotinic antagonist in this
at least seven a subunits, designated a2-a7 and a9 (in chick venom, a-bungarotoxin, in addition to blocking the muscle re-
there is also an a8 subunit), and three b subunits, b2-b4. The
ceptor, potently blocks a7 subunit-containing neuronal
a2, a3, and a4 subunits can each combine with b2 or b4
nAChRs (10). Methyllycaconitine, an alkaloid toxin from the
subunits to form functional channels when expressed in Xeno-
seeds of Delphinium brownii has markedly greater affinity for
pus oocytes, e.g. a2b2, a3b2, a2b4, etc. In addition, a7 and a9
the 125I-a-bungarotoxin binding site in brain versus that in
subunits can be expressed as functional homooligomers in this
muscle demonstrating that these receptor subtypes can be
system. Studies employing either nucleotide probes or antibod-
pharmacologically distinguished (11). A minor component of
ies indicate that each of these a and b subunits have a unique
Bungarus venom, k-bungarotoxin (also known as neuronal
pattern of anatomical expression in the central nervous system
bungarotoxin, toxin F, or Bgt 3.1) preferentially blocks a3b2
receptors (8), although the presence of venom purification con-
* This work was supported by Scientist Development Award for Cli- taminants has led to inconsistent findings (8, 12). Unfortu-
nicians Grant MH00929 (to J. M. M.) and National Institutes of Health nately, due to limited availability of venom, this potent toxin is
Grant PO1 48677 (to B. M. O. and D. Y.). The costs of publication of this
commercially unavailable at the present time.
article were defrayed in part by the payment of page charges. This
article must therefore be hereby marked “advertisement” in accordance A growing number of nicotinic antagonists have been iso-
with 18 U.S.C. Section 1734 solely to indicate this fact. lated from the venom of the carnivorous marine snail Conus
¶ To whom correspondence should be addressed: University of Utah, and are known as a-conotoxins. In contrast to snake a-toxins
201 S. Biology, Salt Lake City, Utah 84112.
1 (;60 – 80 amino acids), a-conotoxins are much smaller (;12–25
The abbreviations used are: nAChR, nicotinic acetylcholine re-
ceptors; RPLC, reverse phase liquid chromatography; Fmoc, amino acids), a feature which has allowed them to be readily
N-(9-fluorenyl)methoxycarbonyl. chemically synthesized (13). a-Conotoxins, which target the

7522
a-Conotoxin MII Targets a3b2 nAChRs 7523
muscle nAChR are enjoying increasing use due to their recently grade) was from Aldrich; acetonitrile (UV grade for semipreparative
discovered ability to differentiate between the two acetylcho- and analytical RPLC, non-spectro grade for preparative RPLC) was
from Baxter.
line binding sites on the receptor. In the mouse muscle-derived
Pyridylethylation and Purification of Modified Peptide—Peptide
BC3H-1 cell line a-conotoxins MI, GI, and SIA (respectively from the final purification was stored in the RPLC buffer in which it
from Conus magus, Conus geographus, and Conus striatus) eluted. A 287-ml solution of this purified peptide (;250 pmol) was
selectively bind to the ACh binding site at the a/d interface with combined with 14.4 ml (20:1 v/v) of 0.5 M Tris base which raised the pH
more than 104-fold greater affinity than the site at the a/g to a value between 7 and 8 as measured with pH paper. Seventy-five ml
interface (14, 15). With Torpedo nAChR, the situation is re- of 50 mM dithiothreitol was added (final concentration 10 mM); the
reaction vessel was flushed with argon, and the reaction incubated at
versed. a-Conotoxins MI and GI bind at the a/g interface with
65 °C for 15 min. The solution was allowed to cool; 15 ml of 20% 4-vinyl
approximately 2 orders of magnitude greater affinity than the pyridine in ethanol was added, and the solution was reacted for a
a/d interface (15, 16). Like a-conotoxins MI, GI, and SIA, further 25 min at room temperature in the dark. The solution was
a-conotoxin EI, from Conus ermineus, prefers the a/d to the a/g diluted 3-fold with 0.1% trifluoroacetic acid, and the alkylated peptide
interface of receptors in BC3H-1 cells, but with only 30-fold was loaded on the Brownlee column. After washing the column with
difference. In contrast to these other a-conotoxins, with Tor- 20% buffer B to allow the baseline to return to 10% of the initial
reading, the peptide was eluted with the gradient described in Fig. 1,
pedo receptors, a-conotoxin EI preferentially binds the a/d ver- panel B.
sus the a/g interface by a 400-fold difference in affinity and is Sequence Analysis—Sequencing was performed with Edman chem-
the only ligand known to possess this selectivity (17). Thus, istry on an Applied Biosystems 477A Protein Sequencer at the Protein/
these a-conotoxins can serve as specific probes to investigate DNA Core Facility at the University of Utah Cancer Center. Mass
structure-function relationships of nAChRs (18). spectrometry was performed as described previously (17).
There are approximately 500 species of Conus. Each of these
predatory gastropods hunt prey from one of five different phyla, Peptide Synthesis
and all of these prey have cholinergic synapses (19). Thus, Linear Peptide—All amino acid derivatives were purchased from

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


there is a potentially very wide diversity of nAChRs for cono- Bachem (Torrance, CA). The linear peptide chain was built on Rink
toxins to target, and it is likely therefore that there are a amide resin by Fmoc (N-(9-fluorenyl)methoxycarbonyl) procedures with
comparably wide spectrum of cholinergically active peptides in 2-(1H-benzotriole-1-yl)-1,1,3,3-tetramethyluronium tetrafluoroborate
the venom of Conus. We are seeking to exploit this situation to coupling, using an ABI model 430A peptide synthesizer. Side chain
protection of non-Cys residues was in the form of t-butyl (Glu, Ser) and
develop a bank of peptides which act on specific subtypes of trityl (Asn, His). Orthogonal protection was used on cysteines: Cys3 and
neuronal nicotinic receptors. By use of a bioassay involving Cys16 were protected as the stable Cys(S-acetamidomethyl), while Cys2
intracranial injections into mice to guide purification, we pre- and Cys8 were protected as the acid-labile Cys(S-trityl). After assembly
viously isolated a-conotoxin ImI which, unlike other a-conotox- of the resin-bound peptide, the terminal Fmoc group was removed in
ins, selectively targets a7, and to a lesser degree a9, nAChRs situ by treatment with 20% piperidine in N-methylpyrrolidone. Linear
(20, 21). In the present study we used a much more specific peptide amide was cleaved from 93 mg of resin by treatment with 1 ml
of trifluoroacetic acid/H2O/ethanedithiol/phenol/thioanisole (90/5/2.5/
screening assay to purify a novel nicotinic antagonist from C. 7.5/5 by volume) for 1.5 h at 20 °C. Released peptide was precipitated by
magus venom. Voltage-clamped Xenopus oocytes expressing filtering the reaction mixture into methyl-t-butyl ether which had been
a3b2 nAChRs were used in the assay to isolate a-conotoxin cooled to 210 °C. This procedure simultaneously cleaved peptide from
MII. We report the structural characterization and nAChR the resin and deprotected the Cys(S-trityl) and the non-Cys residue side
subtype selectivity of this peptide. chains, but not the Cys(S-acetamidomethyl) residues. The cleavage
reaction vessel was rinsed with 100% trifluoroacetic acid, and this rinse
MATERIALS AND METHODS was also filtered into the methyl-t-butyl ether solution. The precipitate
was washed two additional times with chilled methyl-t-butyl ether, and
Peptide Isolation and Sequencing
the supernatants were discarded. Pelleted peptide was dissolved by the
Venom Extraction—Crude venom from dissected ducts of C. magus addition of approximately 10 ml of 0.1% trifluoroacetic acid in 60%
was collected in the Philippines, lyophilized, and stored at 270 °C until acetonitrile, with gentle swirling (to avoid foaming). The linear peptide
used. All reagents were precooled to, and extraction procedures were solution was diluted with 190 ml of 0.1% trifluoroacetic acid and puri-
conducted at, 4 °C. Fifteen ml of 0.1% trifluoroacetic acid was added to fied by RPLC on the preparative C18 Vydac column with a 10 – 60%
500 mg of lyophilized venom, and the mixture was vortexed for 20 min. buffer B gradient over 50 min. Flow rate was 20 ml/min. This gradient
This mixture was centrifuged at 17,000 3 g for 20 min. The supernatant was also used for all subsequent preparative RPLC purifications of the
was transferred to a separate tube, and another 15 ml of 0.1% triflu- synthetic peptide.
oroacetic acid was added to the pellet which was then sonicated with a Peptide Cyclization—To form a disulfide bridge between Cys2 and
Sonifier (Branson Instruments) at setting #4, vortexed for 10 min, and Cys8 (i.e. the first and third cysteines), the major peptide fraction from
centrifuged as above. The supernatants were combined and filtered the preparative RPLC (see above) was diluted to 1 liter with H2O and
through a Whatman GF/C filter (Whatman, Ltd, Maidstone, UK), and solid Tris base was added to increase the pH to 7.6. The solution was
then placed in two Centriprep 30 microconcentrators (Amicon, Beverly, placed in a 4-liter flask and gently swirled at room temperature for 38
MA) which have a 30,000 molecular weight cut-off. The Centripreps h at which time the reaction was judged to be complete by analytical
were centrifuged at 1500 3 g until the retentate in each was reduced to RPLC. The pH of the solution was decreased to a value of 2–3 (meas-
5 ml (;45 min). The filtrate was removed and 10 ml of 0.1% trifluoro- ured with pH paper) by the addition of 4 ml of trifluoroacetic acid. The
acetic acid was added to the retentate of each Centriprep. The Cen- monocyclic peptide was then purified by RPLC and collected in a vol-
tripreps were again centrifuged at 1500 3 g for 120 min. The addition ume of 45 ml. Removal of the S-acetamidomethyl groups and closure of
of trifluoroacetic acid reduced the viscosity of the retentate and im- the second disulfide bridge (Cys3-Cys16, i.e. the second and fourth cys-
proved the recovery of filtrate. Filtrates were combined and used for teines) was carried out simultaneously by iodine oxidation. The 45 ml of
further purification of a-conotoxin MII. RPLC eluent containing the monocyclic peptide was dripped into a
RPLC Purification—All RPLC columns were from Rainin Instru- rapidly stirred 50-ml solution of 20 mM iodine in H20/trifluoroacetic
ments. Crude venom extract was fractionated on a semipreparative acid/acetonitrile/MeOH (50:20:20:10 by volume) over 5.5 min at room
Vydac C18 column (10 mm 3 25 cm, 5 mm particle size, 330 Å pore size) temperature. The reaction was allowed to proceed for another 15 min
equipped with a guard module (catalog number 83-223-65). All other and terminated by the addition of ascorbic acid. The solution was
chromatographic purifications involved an analytical Brownlee C8 col- diluted to 1 liter and the bicyclic peptide purified by RPLC.
umn (4.6 mm 3 22 cm, RP300 packing, 7 mm particle size, 300 Å pore
size) with a guard cartridge which also had RP300 packing material.
Synthetic peptide was purified on a preparative Vydac C18 column (22
Electrophysiology
mm 3 25 cm, 10 mm particle size, 330 Å pore size). For all chromato- cRNA Preparation—cDNA clones encoding nAChR subunits were
grams buffer A was 0.1% trifluoroacetic acid and buffer B was 0.1% provided by S. Heinemann and D. Johnson (Salk Institute, San Diego,
trifluoroacetic acid, 60% acetonitrile. Trifluoroacetic acid (sequencing CA). cRNA was transcribed using either RiboMAXTM large scale RNA
7524 a-Conotoxin MII Targets a3b2 nAChRs

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


FIG. 1. Purification of a-conotoxin MII by RPLC. Panel A, filtrate of venom extract (38.2 ml) was loaded onto a semipreparative Vydac C18
column with 0% buffer B and subsequently eluted using a gradient system. The gradient was 0 –15% buffer B/15 min, then 15–39% buffer B/72
min, then 39 – 65% buffer B/15 min, then 65–100% buffer B/5 min and held at 100% buffer B for 2 min. Flow rate was 5 ml/min. Panel B, fractions
indicated by the arrow in Panel A were diluted with 2 volumes of 0.1% trifluoroacetic acid and repurified on an analytical Brownlee C8 column.
The gradient was 20 –50% buffer B/60 min at a flow rate of 1 ml/min. Panel C, the right half of a center cut of the absorbance peak indicated by
the arrow in Panel B was diluted with 2 volumes of 0.1% trifluoroacetic acid and re-chromatographed as described in panel B to obtain the final
purified product. A 5-ml sample loading loop was used in all chromatography. Buffer A, 0.1% trifluoroacetic acid; and buffer B, 0.1% trifluoroacetic
acid, 60% acetonitrile. Absorbance was monitored at 220 nm.

production systems (Promega, Madison, WI) or an RNA transcription penicillin G (Sigma), 100 mg/ml streptomycin (Sigma), and 100 mg/ml
kit (Stratagene, La Jolla, CA). Diguanosine triphosphate (Sigma) was gentamycin (Life Technologies, Inc., Grand Island, NY)). The oocytes
used for synthesis of capped cRNA transcripts according to the protocol were visually examined and only healthy appearing oocytes were trans-
of the manufacturer. Plasmid constructs of mouse and rat nAChR ferred to a second dish containing ND-96 and antibiotics. Oocytes were
subunits were as described: a1, b1, g, d (22); a2 (23); a3 (24); a4 (25); injected 1–2 days after harvesting and recordings were made 1–7 days
a72; a9 (26); b2 (27); and b4 (28). after injection.
cRNA Injection—cRNA was injected with a Drummond 10-ml micro- Voltage-clamp Recording—An injected oocyte was placed in a ;30-ml
dispenser (Drummond Scientific, Broomall, PA) essentially as described recording chamber consisting of a cylindrical well (;4 mm diameter 3
by Goldin (29). It was fitted with micropipettes pulled from glass cap- 2 mm deep) fabricated from Sylgard, and gravity-perfused with either
illaries provided for the microdispenser. The pipette tips were broken to ND96 or ND96 containing 1 mM atropine (ND96A) at a rate of ;1
an OD of 22–25 mm and back-filled with paraffin before mounting on ml/min. All toxin solutions also contained 0.1 mg/ml bovine serum
the microdispenser. cRNA was drawn into the micropipette and 50 nl,
albumin to reduce nonspecific adsorption of toxin. The perfusion me-
containing 5 ng of cRNA of each subunit, was injected into each oocyte.
dium could be switched to one containing toxin or acetylcholine (ACh)
In the case of muscle subunits, 0.5–2.5 ng of each subunit was injected.
by use of a distributor valve (SmartValve, Cavro Scientific Instruments,
Oocyte Harvesting—Oocytes were removed from Xenopus frogs, cut
Sunnyvale, CA) and a series of three-way solenoid valves (model
into clumps of 20 –50 oocytes, and placed in a 50-ml polypropylene tube
161TO31, Neptune Research, Northboro, MA). ACh-gated currents
(Sarstedt) containing 580 units/ml type 1 collagenase (Worthington) in
OR-2 (82.5 mM NaCl, 2.0 mM KCl, 1.0 mM MgCl2, and 5 mM HEPES, pH were obtained with a two-electrode voltage-clamp amplifier (model OC-
;7.3). The tube was incubated for 1–2 h on a rotary shaker rotating at 725B, Warner Instrument Corp., Hamden, CT) set for “fast” clamp and
50 rpm. Half-way through the incubation, the solution was exchanged with clamp gain at maximum (3 2000). Glass microelectrodes, pulled
with fresh collagenase solution. The oocytes were then washed with six from fiber-filled borosilicate capillaries (1 mm outer diameter 3 0.75
to eight ;50-ml volumes of OR-2, transferred to a 60 3 15-mm Petri mm inner diameter, WPI Inc., Sarasota, FL) and filled with 3 M KCl,
dish containing ND-96 (96.0 mM NaCl, 2.0 mM KCl, 1.8 mM CaCl2, 1.0 served as voltage and current electrodes. Resistances were 0.5–5
mM MgCl2, 5 mM Hepes, pH 7.1–7.5)/Pen/Strep/Gent (100 units/ml megohm for voltage, and 0.5–2 megohm for current electrodes. The
membrane potential was clamped at 270 mV, and the current signal,
recorded through virtual ground, was low-pass filtered (5 Hz cut-off)
2
J. Boulter, unpublished data. and digitized at a sampling frequency of 20 Hz. The solenoid perfusion
a-Conotoxin MII Targets a3b2 nAChRs 7525
valves were controlled with solid state relays (model ODC5 in a of a dose-response curve represents the average 6 S.E. of at least three
PB16HC digital I/O backplane, Opto 22, Temecula, CA). A/D conversion oocytes. Dose-response curves were fit, with Prism software (GraphPad
and digital control of solenoid valves were performed with a Lab-LC or Software Inc., San Diego, CA), to the equation: % response 5 100/{1 1
Lab-NB board (National Instruments, Austin, TX) in a Macintosh ([toxin]/IC50)nH}, where nH is the Hill coefficient. All recordings were
(Quadra 630 or IIcx) computer. The computer communicated with the done at room temperature (;22 °C).
distributor valve via its serial port. Data acquisition and activities of Bioassay—Intraperitoneal injections of toxin into Swiss Webster
the distributor and solenoid valves were automatically controlled by a mice and intramuscular injections into goldfish were performed as
home-made virtual instrument constructed with the graphical pro- described previously (17, 20).
gramming language LabVIEW (National Instruments, Austin, TX). To
apply a pulse of ACh to the oocyte, the perfusion fluid was switched to RESULTS
one containing ACh for 1 s. This was automatically done every 5 min.
Purification and Characterization of a-Conotoxin MII—Se-
The concentration of ACh was 1 mM for oocytes expressing a1b1gd, 1 mM
for a7, and 300 mM for all others. The ACh was diluted in ND96A for all rial dilutions of a 50 mg/ml ND96 buffer extract of crude C.
except a7, in which case the diluent was ND96. For control responses, magus venom were tested for their ability to block the ACh-
the ACh pulse was preceded by perfusion with ND96 (for a7) or ND96A induced current in Xenopus oocytes expressing a3b2 nAChRs.
(all others). No atropine was used with oocytes expressing a7, since it Dose-dependent block was observed; 82% block was produced
has been demonstrated to be an antagonist of these receptors (30). For with 0.071 mg/ml crude venom extract solution (data not
responses in toxin (test responses), the oocyte was perfused with toxin
shown). C. magus venom was purified by RPLC as described
solution until equilibrated (generally 5 min, but up to 25 min at lower
toxin concentrations) before the ACh pulse was applied. All ACh pulses under “Materials and Methods.” For the initial RPLC fraction-
contained no toxin, for it was assumed that little, if any, bound toxin ation (Fig. 1), 5-ml fractions were collected. Aliquots of 0.2% of
would have washed away in the brief time (,2 s) it takes for the each fraction were pooled in groups of three, lyophilized, and
responses to peak (see Fig. 3). The peak amplitudes of the ACh-gated 30% of each pool was tested on oocytes expressing a3b2 sub-
current responses were measured by the virtual instrument. The aver- units (see “Materials and Methods”). Individual fractions of the
age of three control responses just preceding a test response was used
active pool were then tested within the oocyte system and the
to normalize the test response to obtain “% response.” Each data point

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


active fraction purified to homogeneity via RPLC.
The purified peptide was reduced, alkylated, and sequenced
as described under “Materials and Methods.” The sequence is:
GCCSNPVCHLEHSNLC. Liquid secondary ion mass spec-
trometry indicated that Cys residues are present as disulfides
and that the COOH-terminal a-carboxyl group is amidated
(monoisotopic MH1: calculated 1710.65, observed 1710.6). The
sequence was further verified by total chemical synthesis (see
below). The sequence resembles previously isolated a-conotox-
ins in its spacing of Cys residues, yet differs substantially in
other amino acids. As will be shown below, the peptide potently
targets the nAChR, and we have therefore named the peptide
a-conotoxin MII in accordance with the nomenclature previ-
ously proposed for conotoxins (31).
Chemical Synthesis—Solid phase chemical synthesis of
a-conotoxin MII was undertaken to provide an abundant sup-
ply of peptide. It was assumed that the disulfide bridging of
FIG. 2. Comparison of natural and synthetic a-conotoxin MII a-conotoxin MII would be analogous to all previously charac-
by RPLC. Native and synthetic peptide have similar elution times terized a-conotoxins, i.e. 1st Cys-3rd Cys, 2nd Cys-4th Cys. Cys
when chromatographed separately and comigrate when coinjected on groups were protected in pairs to direct disulfide formation.
RPLC. RPLC conditions were as described in the legend to Fig. 1, panel The acid-labile trityl protecting groups were removed from
B, except that a 1-ml sample loading loop was used. Absorbance was
monitored at 220 nM. Maximum OD readings were 0.0031, 0.16, and
Cys2 and Cys8 (i.e. the first and third cysteines) during the
0.012 absorbance units in the first, second, and third panels, cleavage reaction which released linear peptide from the resin.
respectively. Closure of the disulfide bridge was accomplished by air oxida-

FIG. 3. a-Conotoxin MII blocks ACh responses in oocytes expressing a3b2 nicotinic acetylcholine receptors. Oocytes expressing a3b2
nAChRs were voltage-clamped and the response to a 1-s pulse of ACh was measured (see “Materials and Methods”). Panel A, 0.5 nM a-conotoxin-
MII blocks 45% of the ACh-induced response. Panel B, in a different oocyte, 20 nM toxin blocks 98% of the ACh-induced response.
7526 a-Conotoxin MII Targets a3b2 nAChRs

FIG. 4. nAChR block by a-conotoxin MII is slowly reversible. 20


nM toxin was applied to a3b2-expressing oocytes for 5 min. The oocytes
were then continuously perfused with buffer without toxin. Responses
to ACh from a single experiment are shown. The experiment was
repeated twice with similar results.

tion, and the monocyclic peptide was purified by RPLC. The


acid-stable acetamidomethyl group was next removed from FIG. 5. a-Conotoxin MII selectively blocks a3b2 nAChRs. Oo-
cytes expressing various nAChR subunit combinations were voltage
Cys3 and Cys16 (i.e. the second and fourth cysteines), and the clamped and the response to ACh measured. The IC50 for a-conotoxin

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


disulfide bridge closed by rapid iodine oxidation and the bicy- MII block of the a3b2 receptor is 0.5 nM (nH 5 0.8). Oocytes expressing
clic peptide purified by RPLC. Peptide yield was 41.7 nmol/mg other nAChR subunit combinations were perfused with 200 nM a-cono-
peptide resin. toxin MII (400 3 the IC50 on a3b2 receptors). The mean 6 S.E. response
to ACh was as follows: å, a4b4, 96 6 1; ç, a2b4, 96 6 6; ●, a1b1gd, 89
The order of RPLC elution of the synthetic peptides is of note:
6 3; M, a3b4, 85 6 5; D, a2b2, 80 6 3; É, a4b2, 70 6 1; L, a7, 44 6 8.
linear first, followed by monocyclic and bicyclic last. With the Data represent the mean for 3–5 oocytes.
formation of each disulfide bridge, a-conotoxin MII becomes
increasingly hydrophobic. This is exactly the opposite behavior TABLE I
Potencies of a-conotoxins on nAChRs expressed in Xenopus oocytes
from a-conotoxin EI, where the formation of each disulfide
bond results in decreased retention time on RPLC (17), and a-Conotoxin nAChR IC50 Ref.
may indicate that the disulfide bridges in a-conotoxin MII force nM
hydrophobic residues to face outward. Synthetic peptide comi- MI a1b1gd 12.0 21
grated with native on RPLC (Fig. 2). Liquid secondary ioniza- GI a1b1gd 20.0 21
tion mass spectrometry of synthetic a-conotoxin MII was con- ImI a7 220.0 21
sistent with the amidated sequence (monoisotopic MH1: a9 1800.0 21
MII a3b2 0.5 This work
calculated 1710.65, observed 1710.8). Lyophilization or bath
application of small (fmol 2 pmol) amounts of a-conotoxin MII
resulted in apparent loss of peptide (data not shown). The paralysis (n 5 3). This is in contrast to a-conotoxin MI, 0.67
hydrophobic nature of the peptide may account for this problem nmol of which kills a 20-g mouse in 20 min (32). Intramuscular
which was minimized by the use of carrier protein (bovine injection of 5 nmol of a-conotoxin MII into fish did not result in
serum albumin) in all solutions and continuous perfusion any signs of paralysis (n 5 3). This is in contrast to a-conotoxin
rather than bath application of peptide. MI where 0.5 nmol is paralytic.
Electrophysiology—a-Conotoxin MII purification by RPLC
was guided by an assay which used Xenopus oocytes expressing DISCUSSION
a3b2 receptors. RPLC fractions were tested for their ability to nAChR Selectivity—Xenopus oocytes expressing mammalian
block the ACh-induced response in this assay. We also exam- neuronal a3b2 nAChRs were used in an assay which success-
ined the effect of the toxin on other nAChR subunit combina- fully guided the isolation of the novel 16-residue peptide,
tions expressed in oocytes. Both native and synthetic toxin a-conotoxin MII. This is significant in that it is the first a-cono-
blocked the a3b2 nAChRs with equal potency (data not shown). toxin known to target a3b2 receptors. Most previously reported
Due to very limited availability of native toxin, synthetic toxin a-conotoxins target the muscle nAChR. Exceptions are a-cono-
was used for all subsequent experiments. Synthetic a-cono- toxin ImI, which selectively blocks homomeric a7 and a9 re-
toxin MII showed dose-dependent block of a3b2 receptors at ceptors (20, 21), and a-conotoxins PnIA/B, which block mollus-
low nanomolar concentrations (Fig. 3). This block slowly re- can neuronal nAChRs (33). We have shown elsewhere that
versed with washing (Fig. 4). a-Conotoxin MII blocks a3b2 a-conotoxins MI and GI potently target muscle nAChRs ex-
nAChRs with an IC50 of 0.5 nM, with an apparent Hill coeffi- pressed in Xenopus oocytes, but are inactive at all neuronal
cient, nH, of 0.8 (Fig. 5). a-Conotoxin MII was also tested on nAChRs tested, including a3b2 receptors (21). As demon-
other nAChR subunit combinations. Results are shown in Fig. strated in this report, a3b2 receptors are blocked by a-cono-
5 and indicate that a-conotoxin MII is 2– 4 orders of magnitude toxin MII with an IC50 of 0.5 nM. The effectiveness of the toxin
more potent at a3b2 nAChRs than at other nAChR subtypes. on the other nAChR subunit combinations tested is 2– 4 orders
We have previously shown that a-conotoxin MI (also from C. of magnitude less. Thus, a-conotoxin MII has an entirely
magus venom), a-conotoxin GI, and a-conotoxin ImI have no unique activity profile among the a-conotoxins (Table I), and
effect on a3b2 receptors at up to 5 mM concentration (21). Thus, represents a potent and selective new probe for studying
a-conotoxin MII is the only conotoxin known to potently block nAChRs. Notably, the small size of a-conotoxin MII has al-
this neuronal receptor subtype. lowed it to be chemically synthesized and thus readily
In Vivo Activity—Intraperitoneal injections of 10 nmol of available.
a-conotoxin MII into 8 –10-g mice did not result in any signs of Structural Relationships among a-Conotoxins—Reported
a-Conotoxin MII Targets a3b2 nAChRs 7527
TABLE II paralysis in this assay. It has recently been shown that only as
Sequences of a-conotoxins few as three amino acid substitutions in the g or d subunit of
The Conus species from which each toxin was isolated and the prey of mouse nAChR can result in a 104-fold change in the affinity of
that species are indicated. Disulfide bond pattern has not been deter-
mined for PnIA/B, GII, SIA, or SII. this receptor for a-conotoxin MI (18). It has also been shown
that a single amino acid substitution in a-conotoxin SI in-
creases its affinity for mouse muscle nAChRs by 2 orders of
magnitude (35). It is possible, therefore, that a-conotoxin MII
does potently target the muscle nAChR of its natural tropical
fish prey, and that a few amino acid substitutions in the gold-
fish nAChR used in our assay may be responsible for the
observed substantial differences in toxin potency. Poor disper-
sal of the more hydrophobic a-conotoxin MII might also lead to
an apparent lack of activity in our assay. Since C. magus
already has a toxin which potently blocks muscle nAChRs in
the form of a-conotoxin MI, another possibility is that C. magus
uses a-conotoxin MII to selectively target ganglionic or adrenal
nAChRs in fish to lessen the sympathetically mediated fight or
flight response. In frog, a-conotoxin MII blocks ganglionic neu-
rotransmission.4 A related example may be neosurugatoxin.
This glycoside, isolated from mid-gut gland of the Japanese
a-conotoxins can be classified into two main groups based on ivory shell, appears to be responsible for human poisonings
the spacing of the cysteine residues. One group has a “3,5 following ingestion of this carnivorous gastropod (9). Poisoning

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


spacing” where the numerals indicate the number of amino symptoms are consistent with blockade of autonomic ganglia.
acids between the second and third Cys and the third and Like a-conotoxin MII, neosurugatoxin preferentially targets
fourth Cys, respectively (see Table II). The other group, which neuronal versus muscle nAChRs. In contrast, however, neosu-
includes a-conotoxin MII, has a “4,7 spacing.” Individual toxins rugatoxin is non-selective among axb2 nAChRs (4, 8).
from both groups can potently block the muscle nAChR, sug- The discovery of a-conotoxin MII, which differs substantially
gesting that it is not the Cys spacing which is responsible for in both structure and function from other a-conotoxins, pro-
a-conotoxin MII’s selectivity. Aside from the Cys residues, the vides further evidence of the enormous diversity of nAChR-
only completely conserved residue in all reported a-conotoxins targeted toxins present in Conus. This report demonstrates the
is a proline between the second and third Cys. The other feasibility of using specific nAChR subunit combinations ex-
non-Cys residues in a-conotoxin MII are strongly divergent pressed in oocytes as a functional screen to initially detect and
from all other a-conotoxins. Of the non-Cys residues in the 4,7 ultimately guide the purification of these peptides.
group, a-conotoxin MII from C. magus shares only 4 of 12 Acknowledgments—Mass spectrometry was performed by A. Gray
residues with a-conotoxin PnIB from C. pennaceus and only 1 of Craig of the Salk Institute. We thank Stephen F. Heinemann and David
12 residues with a-conotoxin EI from C. ermineus. Further- S. Johnson for providing nAChR subunit clones and John Syphers for
more, except for the proline, a-conotoxin MII shares little if any technical assistance.
homology with a-conotoxin MI (which has a 3,5 spacing) al- REFERENCES
though both are from the same Conus species. However, de- 1. Changeux, J.-P. (1993) Sci. Am. 269, 58 – 62
spite the difference in Cys spacing and strong divergence in 2. Arneric, S. P., Sullivan, J. P., and Williams, M. (1995) in Psychopharmacology:
The Fourth Generation of Progress (Bloom, F. E., and Kupfer, D. J., eds)
other amino acids, the disulfide bridges are exactly analogous pp. 95–109, Raven Press, New York
between the two toxins, 1st Cys to 3rd Cys, 2nd Cys to 4th Cys, 3. Taylor, P. (1990) in The Pharmacological Basis of Therapeutics, Eighth Edition
and this arrangement is also conserved in all other a-conotox- (Gilman, A. G., Rall, T. W., Nies, A. S., and Taylor, P., eds) pp. 166 –186,
Pergamon Press, New York
ins where the disulfide linkages have been studied (Table II). 4. Chiappinelli, V. A. (1993) in Natural and Synthetic Neurotoxins (Harvey, A.,
Recently, a new family of Conus peptides targeted to nAChRs eds) pp. 66 –128, Academic Press, Harcourt Brace Jovanovich, New York
5. Fenical, W., Okuda, R. K., Bandurraga, M. M., Culver, P., and Jacobs, R. S.
was reported. These aA-conotoxins have a distinctly different (1991) Science 212, 1512–1514
structure including three disulfide bridges instead of two (34). 6. Abramson, S. N., Culver, P., Kline, T., Li, Y., Guest, P., Gutman, L., and
Multiple nAChR-targeted toxins have previously been iso- Taylor, P. (1988) J. Biol. Chem. 263, 18568 –18573
7. Abramson, S. N., Li, Y., Culver, P., and Taylor, P. (1989) J. Biol. Chem. 264,
lated from two other Conus species, C. geographus and C. 12666 –12672
striatus (see Table II). In both of these cases, however, the 8. Luetje, C. W., Wada, K., Rogers, S., Abramson, S. N., Tsuji, K., Heinemann, S.,
and Patrick, J. (1990) J. Neurochem. 55, 632– 640
a-conotoxins have considerable structural homology and all 9. Kosuge, T., Tsuji, K., and Koichi, H. (1981) Tetrahedron Lett. 22, 3417–3420
target the muscle nAChR in contrast to the MI and MII pep- 10. Séguéla, P., Wadiche, J., Dineley-Miller, K., Dani, J. A., and Patrick, J. W.
tides isolated from C. magus that differ substantially in both (1993) J. Neurosci. 13, 596 – 604
11. Ward, J. M., Cockcroft, V. B., Lunt, G. G., Smillie, F. S., and Wonnacott, S.
structure and function. It will be of interest to determine which (1990) FEBS Lett. 270, 45– 48
residues in MII confer a3b2 nAChR selectivity and which in MI 12. Fiordalisi, J. J., Al-Rabiee, R., Chiappinelli, V. A., and Grant, G. A. (1994)
Biochemistry 33, 12962–12967
confer a1b1gd selectivity. 13. Myers, R. A., Cruz, L. J., Rivier, J., and Olivera, B. M. (1993) Chem. Rev. 93,
Biological Role—Injection of venom by Conus snails results 1923–1936
in prey immobilization and capture. Nevertheless, the majority 14. Kreienkamp, H.-J., Sine, S. M., Maeda, R. K., and Taylor, P. (1994) J. Biol.
Chem. 269, 8108 – 8114
of the 100 –200 peptides present in the venom of fish-hunting 15. Groebe, D. R., Dumm, J. M., Levitan, E. S., and Abramson, S. N. (1995) Mol.
Conus do not induce paralysis when injected into fish.3 The Pharmacol. 48, 105–111
16. Hann, R. M., Pagán, O. R., and Eterovic, V. A. (1994) Biochemistry 33,
functional targets and roles of these non-paralytic peptides are 14058 –14063
under investigation. However, all previously reported a-cono- 17. Martinez, J. S., Olivera, B. M., Gray, W. R., Craig, A. G., Groebe, D. R.,
toxins from fish-hunting Conus do cause rapid paralysis when Abramson, S. N., and McIntosh, J. M. (1995) Biochemistry 34, 14519 –14526
18. Sine, S. M., Kreienkamp, H.-J., Bren, N., Masada, R., and Taylor, P. (1995)
injected into fish. a-Conotoxin MII is unique in not causing
4
S. Tavazoie, M. Tavazoie, J. M. McIntosh, B. M. Olivera, and D.
3
B. M. Olivera and J. M. McIntosh, unpublished observations. Yoshikami, manuscript in preparation.
7528 a-Conotoxin MII Targets a3b2 nAChRs
Neuron 15, 205–211 30. Gerzanich, V., Anand, R., and Lindstrom, J. (1994) Mol. Pharmacol. 45,
19. Olivera, B. M., Rivier, J., Clark, C., Ramilo, C. A., Corpuz, G. P., Abogadie, 212–220
F. C., Mena, E. E., Woodward, S. R., Hillyard, D. R., and Cruz, L. J. (1990) 31. Cruz, L. J., Gray, W. R., Yoshikami, D., and Olivera, B. M. (1985) J. Toxicol.
Science 249, 257–263 Toxin Rev. 4, 107–132
20. McIntosh, J. M., Yoshikami, D., Mahe, E., Nielsen, D. B., Rivier, J. E., Gray, 32. Gray, W. R., Rivier, J. E., Galyean, R., Cruz, L. J., and Olivera, B. M. (1983)
W. R., and Olivera, B. M. (1994) J. Biol. Chem. 269, 16733–16739 J. Biol. Chem. 258, 12247–12251
21. Johnson, D. S., Martinez, J., Elgoyhen, A. B., Heinemann, S. S., and McIntosh, 33. Fainzilber, M., Hasson, A., Oren, R., Burlingame, A. L., Gordon, D., Spira,
J. M. (1995) Mol. Pharmacol. 48, 194 –199 M. E., and Zlotkin, E. (1994) Biochemistry 33, 9523–9529
22. Patrick, J., Boulter, J., Goldman, D., Gardner, P., and Heinemann, S. (1988) 34. Hopkins, C., Grilley, M., Miller, C., Shon, K.-J., Cruz, L. J., Gray, W. R.,
Ann. N. Y. Acad. Sci. 505, 194 –207 Dykert, J., Rivier, J., Yoshikami, D., and Olivera, B. M. (1995) J. Biol.
23. Wada, K., Ballivet, M., Boulter, J., Connolly, J., Wada, E., Deneris, E. S., Chem. 270, 22361–22367
Swanson, L. W., Heinemann, S., and Patrick, J. (1988) Science 240, 35. Gray, W. R., Palmieri, S. J., Christensen, D. J., Groebe, D. R., and Abramson,
330 –334 S. N. (1995) J. Neurosci. Abstr. 21, 10
24. Boulter, J., Evans, K., Goldman, D., Martin, G., Treco, D., Heinemann, S., and 36. Gray, W. R., Luque, A., Olivera, B. M., Barrett, J., and Cruz, L. J. (1981)
Patrick, J. (1986) Nature 319, 368 –374 J. Biol. Chem. 256, 4734 – 4740
25. Goldman, D., Deneris, E., Luyten, W., Kochhar, A., Patrick, J., and 37. McIntosh, J. M., Cruz, L. J., Hunkapiller, M. W., Gray, W. R., and Olivera,
Heinemann, S. (1987) Cell 48, 965–973 B. M. (1982) Arch. Biochem. Biophys. 218, 329 –334
26. Elgoyhen, A. B., Johnson, D. S., Boulter, J., Vetter, D. E., and Heinemann, S. 38. Zafaralla, G. C., Ramilo, C., Gray, W. R., Karlstrom, R., Olivera, B. M., and
(1994) Cell 79, 705–715 Cruz, L. J. (1988) Biochemistry 27, 7102–7105
27. Deneris, E. S., Connolly, J., Boulter, J., Wada, E., Wada, K., Swanson, L. W., 39. Myers, R. A., Zafaralla, G. C., Gray, W. R., Abbott, J., Cruz, L. J., and Olivera,
Patrick, J., and Heinemann, S. (1988) Neuron 1, 45–54 B. M. (1991) Biochemistry 30, 9370 –9377
28. Duvoisin, R. M., Deneris, E. S., Patrick, J., and Heinemann, S. (1989) Neuron 40. Ramilo, C. A., Zafaralla, G. C., Nadasdi, L., Hammerland, L. G., Yoshikami,
3, 487– 496 D., Gray, W. R., Ramachandran, J., Miljanich, G., Olivera, B. M., and Cruz,
29. Goldin, A. L. (1992) Methods Enzymol. 207, 226 –279 L. J. (1992) Biochemistry 31, 9919 –9926

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


A New -Conotoxin Which Targets 32 Nicotinic Acetylcholine Receptors
G. Edward Cartier, Doju Yoshikami, William R. Gray, Siqin Luo, Baldomero M. Olivera
and J. Michael McIntosh
J. Biol. Chem. 1996, 271:7522-7528.
doi: 10.1074/jbc.271.13.7522

Access the most updated version of this article at http://www.jbc.org/content/271/13/7522

Alerts:
• When this article is cited
• When a correction for this article is posted

Click here to choose from all of JBC's e-mail alerts

Downloaded from http://www.jbc.org/ by guest on October 22, 2020


This article cites 37 references, 13 of which can be accessed free at
http://www.jbc.org/content/271/13/7522.full.html#ref-list-1

You might also like