Swartz 1998

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

© 1998 Nature America Inc. • http://neurosci.nature.

com

articles

Inhibition of T-type voltage-gated


calcium channels by a new
scorpion toxin
Rosalind S-I. Chuang1, Howard Jaffe2, Leanne Cribbs3, Edward Perez-Reyes3 and Kenton J. Swartz1

1 Molecular Physiology and Biophysics Unit, National Institute of Neurological Disorders and Stroke, National Institutes of Health,
Bldg. 36, Rm. 2C19, 36 Convent Drive, MSC 4066, Bethesda, Maryland 20892, USA
2 Protein/Peptide Sequencing Facility, National Institute of Neurological Disorders and Stroke, National Institutes of Health,
Bldg. 36, Rm. 3B26, 36 Convent Drive, MSC 4066, Bethesda, Maryland 20892, USA
3 Department of Physiology, Loyola University Medical Center, 2160 South First Avenue, Maywood, Illinois 60153, USA
© 1998 Nature America Inc. • http://neurosci.nature.com

Correspondence should be addressed to K.J.S. (kjswartz@codon.nih.gov)

The biophysical properties of T-type voltage-gated calcium channels are well suited to pacemaking
and to supporting calcium flux near the resting membrane potential in both excitable and non-
excitable cells. We have identified a new scorpion toxin (kurtoxin) that binds to the α1G T-type
calcium channel with high affinity and inhibits the channel by modifying voltage-dependent gating.
This toxin distinguishes between α1G T-type calcium channels and other types of voltage-gated
calcium channels, including α1A, α1B, α1C and α1E. Like the other α-scorpion toxins to which it is
related, kurtoxin also interacts with voltage-gated sodium channels and slows their inactivation.
Kurtoxin will facilitate characterization of the subunit composition of T-type calcium channels and
help determine their involvement in electrical and biochemical signaling.

The biophysical properties of T-type voltage-gated calcium Results


channels differ from those of other types of voltage-gated cal- Expression of the recently cloned α1G calcium channel in
cium channels in that they activate at voltages near the rest- Xenopus oocytes gives rise to calcium channels with bio-
ing membrane potential of many cells and inactivate (close physical characteristics resembling those of native T-type
despite continued depolarization) rapidly1–8. T-type calcium calcium channels studied in neurons, muscle and other
channels also deactivate (close) slowly and have a small sin- cell types16. To improve expression, we subcloned the α1G
gle-channel conductance2–8. Because of these unique gating cDNA into a vector containing the 5’ and 3’ untranslated
properties, T-type calcium channels are thought to be involved regions of a Xenopus β-globin gene17. Injection of oocytes
in pacemaking in neurons and cardiac myocytes1,9–11 and to with cRNA for the α1G channel flanked by these untrans-
contribute to a standing calcium current near the resting lated sequences resulted in robust expression of T-type
membrane potential in various cell types11–13. The expression channels (peak inward currents of 1 to 10 µA with 5 mM
of T-type calcium channels outside the nervous system8,12–14 barium as charge carrier) within 4 to 7 days.
suggests that they have other important functions. We screened the venom from different scorpions for
Progress in understanding the subunit composition and activity against the α 1G T-type calcium channel. The
physiological roles of T-type calcium channels has been hin- venom of Parabuthus transvaalicus contained activity that
dered by a scarcity of ligands that can distinguish between T- produced robust inhibition of the α1G calcium channel
type calcium channels and other types of voltage-gated (not shown). The scorpion venom was fractionated by
calcium channels. T-type calcium channels are not inhibited reversed-phase high-performance liquid chromatogra-
by many types of calcium channel inhibitors, and the phy (HPLC), and individual fractions were examined for
inhibitors that do act on T-type calcium channels lack speci- activity against the α1G calcium channel. After two con-
ficity. For example, mibefradil, perhaps one of the best lig- secutive separations (Fig. 1a and b), a single symmetrical
ands for T-type calcium channels, also inhibits other types of peak (Fig. 1c) had strong inhibitory activity. The pri-
high-voltage-activated calcium channels15. We began a search mary amino-acid sequence of the active material was
for a ligand that binds to T-type voltage-gated calcium chan- determined by Edman degradation. The protein, which
nels using the recently identified α 1G T-type calcium chan- we named kurtoxin, contains 63 amino-acid residues
nel 16 as a target. Venoms from different scorpions were (Fig. 1d). The predicted mass of this protein is
examined for activity against the α1G T-type calcium channel 7386.5 daltons, consistent with a mass of 7386.1 daltons
expressed in Xenopus oocytes. Here we describe the purifica- measured using mass spectrometry. Comparison of the
tion, biochemical analysis and biophysical characterization of amino-acid sequences of kurtoxin and other previously
kurtoxin, a new protein toxin from the venom of a South described protein toxins indicates that kurtoxin is close-
African scorpion (Parabuthus transvaalicus). ly related to the α-scorpion toxins, a large family of tox-

668 nature neuroscience • volume 1 no 8 • december 1998


© 1998 Nature America Inc. • http://neurosci.nature.com

articles

ins that slow inactivation of voltage- a b

Absorbance (280 nm)

Absorbance (215 nm)


gated sodium channels18.
We examined the inhibitory effect of
kurtoxin on α1G T-type calcium channels
expressed in Xenopus oocytes (Fig. 2).
The channel was activated by depolariz-
ing steps to –40 mV from a holding volt-
Time (min) Time (min)
age of –90 mV, given every 10 seconds.
External application of kurtoxin inhibit- c d

Absorbance (215 nm)


ed voltage-activated barium currents in
oocytes expressing the α1G calcium chan-
nel. After removal of the toxin from the
recording chamber, calcium channel cur-
rents recovered slowly (τoff ~250 s).
Because kurtoxin is highly related to
the α-scorpion toxins, which are known
Time (min)
gating modifiers of sodium channels, we
anticipated that kurtoxin was not Fig. 1. Purification of kurtoxin from scorpion venom. (a) Initial fractionation of Parabuthus
© 1998 Nature America Inc. • http://neurosci.nature.com

inhibiting the T-type calcium-channel transvaalicus venom by reversed-phase HPLC. The chromatogram shown resulted from the
by a pore-blocking mechanism, but by injection of the component of venom (0.5 mg) soluble in 25% aqueous acetonitrile with 0.1%
modifying channel gating. We examined trifluoroacetic acid. The asterisk indicates the peak containing kurtoxin. The solvent system
this possibility by studying the effects of used aqueous acetonitrile with trifluoroacetic acid. (see Methods) (b) The active peak in (a),
kurtoxin at different voltages (Fig. 3). At indicated by the asterisk, was fractioned a second time using reversed-phase HPLC yielding
a concentration of 350 nM, kurtoxin pure kurtoxin as indicated. The solvent system used aqueous methanol with trifluoroacetic
inhibited T-type calcium-channel cur- acid (see Methods). (c) Chromatogram of pure kurtoxin. Chromatography was the same as
rent elicited by a weak depolarization by in (b). (d) Complete primary amino-acid sequence of kurtoxin.
over 95% (Fig. 3a). However, calcium-
channel currents elicited by strong depo-
larization were less inhibited by kurtoxin
(Fig. 3a). Activation curves for the α 1G calcium channel macroscopic and single-channel currents suggested that the
(Fig. 3b) showed that 350 nM kurtoxin shifted the opening of inactivation step for T-type calcium channels is not intrinsi-
the α1G calcium channel to more positive voltages and that the cally voltage dependent, but derives its apparent voltage depen-
slope of the activation–voltage relation was more shallow in dence at negative voltages from coupling to voltage-dependent
the presence of the toxin. These effects of kurtoxin are similar activation steps8. The convergence of τinactivation at depolarized
to what has been observed for grammotoxin19, ω-AgaIVA 20 voltages (Fig. 4c) suggests that the effect of kurtoxin on inac-
and hanatoxin21, three established gating modifiers of other tivation kinetics is probably secondary to an effect of the toxin
types of voltage-gated ion channels. The α1G calcium channels on transitions in the activation pathway. These results, togeth-
opened by large depolarization in the presence of kurtoxin er with the sequence homology between kurtoxin and other
deactivated with kinetics virtually identical to those of the con- known gating-modifier toxins (for example, α-scorpion tox-
trol channels without toxin (Fig. 3c). In contrast, the activa-
tion kinetics of channels opened by strong depolarization in
the presence of kurtoxin were much slower than those of con- a
trol channels (Fig. 3c, see also Fig. 4b). The effects of kurtoxin
on inactivation were examined using longer (50 ms) depolar-
izing pulses (Fig. 4). Channels opened by moderate depolar-
ization (for example, 0 mV) in the presence of toxin inactivated
more slowly (τ = 25 ms) than control channels (τ = 6 ms) at
the same voltage (Fig. 4b). The slowed inactivation for channels 10 ms
opened in the presence of toxin became less pronounced with
stronger depolarization (Fig. 4b and c; at +100 mV, τ = 8 ms
in control and τ = 12 ms in toxin). Previous analyses of both

b
Fig. 2. Inhibition of α1G T-type calcium channels by kurtoxin.
(a) Currents elicited by 30 ms depolarizations to –40 mV in the
absence or presence (200 nM) of kurtoxin. Charge carrier was 5
IBa2+ (µA)

mM barium. Holding voltage was –90 mV, and tail voltage was –60
mV. Dashed lines indicate zero current. Linear capacitive current
and very small background conductances, isolated by blocking cal-
cium channels with 1 mM CdCl2, have been subtracted. (b) Time
course for both onset and recovery from inhibition by kurtoxin.
Peak inward barium current is plotted against time. The pulse pro-
tocol illustrated in (a) was repeated every 10 s. Time (s)

nature neuroscience • volume 1 no 8 • december 1998 669


© 1998 Nature America Inc. • http://neurosci.nature.com

articles

a b c

Itail (µA)
10 ms
Test voltage (mV)
10 ms

Fig. 3. Kurtoxin shifts the opening of the α1G T-type calcium channel to more depolarized voltages. (a) Currents elicited by 7.5-ms depolar-
izations to various voltages in the absence (open circles) or presence (closed circles; 350 nM) of kurtoxin. Holding voltage was –90 mV, and
tail voltage was –60 mV. Dashed lines indicate zero current. Linear capacitive current and very small background conductances have been
subtracted using 1 mM CdCl2. (b) Activation curves in the absence (open circles) and presence (closed circles; 350 nM) of kurtoxin. Tail
amplitude, measured 1 ms after repolarization to –60 mV, plotted against test voltage. Holding voltage was –90 mV. Linear capacitive current
© 1998 Nature America Inc. • http://neurosci.nature.com

and very small background conductances have been subtracted using 1 mM CdCl 2. In control solution, the Itail–voltage relation at a constant
test-pulse duration of 7.5 ms does not precisely reflect the voltage dependence of steady-state activation. At negative voltages, 7.5 ms is not
quite long enough to reach steady state, and at positive voltages, many channels inactivate before the tail current is elicited (see traces in a).
(c) Scaled traces for depolarizations to +80 mV in either control or 350 nM kurtoxin. Tail voltage was –60 mV.

ins), suggest that kurtoxin inhibits the α 1G T-type calcium which the fraction of uninhibited current at negative voltages
channel by modifying the energetics of channel gating. was used to estimate the fraction of unbound channels.
We examined the concentration dependence for inhibition (Because kurtoxin does not affect deactivation kinetics, we
of the α1G calcium channel by kurtoxin. The fraction of unin- cannot use tail-current kinetics to directly distinguish between
hibited current does not faithfully approximate the fraction the opening of toxin-unbound and toxin-bound channels, as
of unbound channels at all voltages because kurtoxin shifts has been done with hanatoxin21). Current amplitudes were
the activation of channels to more depolarized voltages. We measured following various weak depolarizations in the pres-
therefore used the approach described for hanatoxin 21 , in ence of different concentrations of kurtoxin (Fig. 5a). Con-
sistent with previous observations for hanatoxin inhibition of
a voltage-gated potassium channel21, the fraction of uninhib-
a ited current reached a plateau at negative voltages and
increased with stronger depolarization. The fraction of unin-
hibited current at negative voltages, which approximates the
fractional occupancy of the channel by toxin, is plotted against
kurtoxin concentration (Fig. 5b). The data were well described
by an equation for 1:1 stoichiometry between toxin and chan-
nel with an equilibrium dissociation constant (Kd) of 15 nM.
We conclude that kurtoxin binds to the α1G T-type calcium
10 ms
channel with high affinity and that the stoichiometry between
toxin and channel is probably 1:1.
We examined the selectivity of kurtoxin for different types of
voltage-gated calcium channels. At 350 nM, kurtoxin almost
completely inhibited the α1G T-type calcium channel and inhib-
ited the α1H T-type calcium channel22 (closely related to α1G)
by about 85% (Fig. 6). This suggests that kurtoxin binds to α1H
b c with nearly as high an affinity as α 1G (Fig. 5 and legend to

Fig. 4. Kurtoxin slows activation and inactivation of α1G T-type cal-


τ inactivation (ms)

cium channels. (a) Currents elicited by 50-ms depolarizations to


various voltages in the absence (open circles) or presence (closed
circles; 350 nM) of kurtoxin. Holding voltage was –90 mV, and tail
voltage was –60 mV. Dashed lines indicate zero current. Linear
capacitive current and very small background conductances have
been subtracted using 1 mM CdCl2. (b) Scaled currents elicited by
depolarization to either 0 mV or +100 mV in both the absence and
presence of 350 nM kurtoxin. (c) τ inactivation plotted as a function
10 ms
Test voltage (mV) of test voltage in control and 350 nM kurtoxin.

670 nature neuroscience • volume 1 no 8 • december 1998


© 1998 Nature America Inc. • http://neurosci.nature.com

articles

a b effects of known gating modi-


fiers 19–21 on other voltage-gated
channels argue that kurtoxin inhibits
the T-type calcium channel by mod-
I

Io ρ0 ifying channel gating. This conclu-
sion is strengthened by the sequence
homology between kurtoxin and the
α-scorpion toxins, a group of well
studied toxins that modify the gating
of sodium channels18.
Test voltage (mV) [kurtoxin] (M) The cross-reactivity of kurtoxin
with the α1G T-type calcium channel
Fig. 5. Concentration dependence of inhibition by kurtoxin. (a) Fraction of uninhibited current
for various strength depolarizations in different concentrations of kurtoxin. Holding voltage was
and a sodium channel is remarkable
–90 mV. I is peak inward current in the specified concentration of toxin, and I0 is peak inward and suggests that the toxin recognizes
current in control. (b) Probability unbound (ρ0) plotted against kurtoxin concentration. ρ0 is a structural motif common to both
equal to the fraction of uninhibited current in the plateau phase at negative voltages, typically channels. Similar cases of toxin
33
near –50 mV. Data points are mean ± standard error for 3–6 determinations. Smooth curve is a promiscuity have been reported for
fit of ρ0 = (1 + ([kurtoxin]/Kd)) to the data.
–1 grammotoxin (a calcium channel
© 1998 Nature America Inc. • http://neurosci.nature.com

inhibitor)19 and hanatoxin (a potas-


sium channel inhibitor)21, two relat-
ed toxins from tarantula venom. Each
Fig. 6). In contrast, the same concentration of kurtoxin did not of these toxins exhibits cross-reactivity with voltage-gated cal-
affect α1B (the N-type calcium channel) 23, α1A (the P/Q-type cium and potassium channels and inhibits these channels by
calcium channel)24, α1C (an L-type calcium channel)25 or α1E modifying the energetics of channel gating33. At least some of
(ref. 26), suggesting that the Kd of kurtoxin binding to these the same residues, for example, Glu 277 and Phe 273 in the
channels is probably greater than 10 µM. Thus, kurtoxin dis- S3–S4 loop of the drk1 potassium channel, contribute to form
tinguishes between α1G and the high-voltage-activated calcium receptors for both hanatoxin and grammotoxin33,34. We spec-
channels with over 600-fold selectivity.
The sequence homology between kur- p control
toxin and the other sodium channel tox- Fig. 6. Selectivity of kurtoxin for differ-
P 350 nM kurtoxin
ent voltage-gated calcium channels.
ins ranges from 30 to 60% (Fig. 7a), α1G α1H Superimposed traces were recorded in
which raised the possibility that kurtoxin
control or 350 nM kurtoxin. Holding
also interacts with voltage-gated sodium voltage was –90 mV, and depolarizations
channels. Therefore, we tested whether were to –40 mV for α1G and α1H, +20
kurtoxin influences the rat brain IIA mV for α1B, –5 mV for α1A, +10 mV for
sodium channel α subunit coexpressed
27
α1E and –10 mV for α1C. Dashed lines
with rat β1 (ref. 28). Application of kur- indicate zero current. All high-voltage-
toxin produced a pronounced slowing of activated calcium channel α subunits
both activation and inactivation kinetics were expressed with β1b and α2δ. Charge
(Fig. 7). These results indicate that kur- carrier was 5 mM barium. Linear capaci-
toxin binds to the sodium channel and tive current and very small background
modifies channel gating. The effect of the conductances (< 10 nA from –90 to +50
toxin on inactivation is similar to what α1B
mV) were isolated by blocking calcium-
α1A
has been observed for other α-scorpion (+ β1B, α2δ) (+ β1B, α2δ)
channel current with CdCl2 and have
toxins18,29–32. been subtracted. 100 µM CdCl2 was used
for blocking all high-voltage-activated
Discussion channels, whereas 1 mM CdCl2 was used
Protein toxins that bind to ion channels for blocking α1G and α1H. Mean ± stan-
have been invaluable tools for studying dard error (n = 3) values for ratio of cur-
the structural basis of channel function rent in 350 nM kurtoxin divided by
current in control (at voltages that elicit
and for investigating the physiological
peak inward current) for the high-volt-
roles of particular channel subtypes. By
age-activated channels were as follows:
screening for ligands that bind to the α1G α1B, 0.99 ± 0.036; α1A, 1.00 ± 0.015; α1E,
T-type calcium channel, we identified 0.97 ± 0.021; α1C, 1.03 ± 0.035. Mean ±
kurtoxin, a 63 amino-acid protein toxin α1E α1C
standard error (n = 5–6) for ratio of cur-
found in venom of the Parabuthus trans- (+ β1B, α2δ) (+ β1B, α2δ)
rent in 350 nM kurtoxin divided by cur-
vaalicus scorpion. Kurtoxin binds with rent in control (both elicited by
high-affinity (Kd = 15 nM) to what seems depolarization to –40 mV) for the low-
to be a single site on the α1G T-type cal- voltage-activated channels were as fol-
cium channel and inhibits the channel in lows: α1H, 0.15 ± 0.011; α1G, 0.035 ±
a voltage-dependent manner. The simi- 0.0053. A value of 0.15 for the fraction of
larities between the effects of kurtoxin on uninhibited current (at 350 nM kurtoxin )
the T-type calcium channel and the for α1H corresponds to a Kd of 61 nM.

nature neuroscience • volume 1 no 8 • december 1998 671


© 1998 Nature America Inc. • http://neurosci.nature.com

articles

Fig. 7. Kurtoxin slows inactiva-


tion of voltage-gated sodium
a
channels. (a) Amino-acid
sequence of kurtoxin (Ktx)
compared to eleven other α-
scorpion toxins. Shading of
residues in all toxins (except
kurtoxin) indicates conservation
with kurtoxin. Shading in kur-
toxin indicates the residue is
conserved in at least one toxin
below. (b) Effect of kurtoxin on
a voltage-gated sodium channel.
The rat brain IIA α subunit and
rat β1 subunit were coex- b
pressed. (left traces) Currents
elicited by depolarization to –5
mV in the presence or absence
of 500 nM kurtoxin. Holding
© 1998 Nature America Inc. • http://neurosci.nature.com

voltage was –90 mV. Multiple


traces showing currents elicited
by various strengths of depolar-
ization in control (middle) or
500 nM kurtoxin (right).
10 ms
Depolarizations were to volt-
ages between –60 mV and +10
mV in 5-mV increments. Dashed lines indicate zero current. Linear capacitive current and very small background conductances (<10
nA), isolated by blocking sodium channels with 1 µM tetrodotoxin, have been subtracted.

ulate that kurtoxin binds to a region of the cloned T-type cal- um channels that are similar to T-type calcium channels in
cium channels and the sodium channel that is homologous to many native cells16,22. However, other types of voltage-gated
the hanatoxin/grammotoxin receptor on voltage-gated potas- calcium-channel α subunits have been implicated as giving rise
sium and calcium channels. Glu 1613 in the S3–S4 loop of to low-voltage-activated calcium channels 35–38 . Kurtoxin
repeat IV in the rat brain IIA sodium channel is homologous to should be a useful reagent to help clarify these issues because it
Glu 277 in the potassium channel33, and mutation of Glu 1613 nicely distinguishes between α1G, α1H and the other voltage-
has a large effect on the binding affinity of an α-scorpion gated calcium-channel α subunits. In general, kurtoxin should
toxin32. Kurtoxin is an α-scorpion toxin, defined by its amino- be a valuable tool for defining both the subunit composition
acid sequence and by its ability to bind to sodium channels and of native T-type calcium channels and the involvement of T-
slow inactivation (Fig. 7). Moreover, its effect on the α1G T- type calcium channels in electrical and biochemical signaling.
type calcium channel (Figs 3 and 5) is similar to the effects of
hanatoxin and grammotoxin on both voltage-gated potassium Methods
and calcium channels19,21,33. It is interesting that only in repeat E XPRESSION OF VOLTAGE - GATED CALCIUM AND SODIUM CHANNELS IN
IV of α 1G and α 1H T-type calcium channels does the S3–S4 X ENOPUS OOCYTES . To improve expression of the rat α 1G calcium
loop have a Glu at the position equivalent to that of Glu277 in channel16, its cDNA was subcloned into the pGEM-HEA vector, thus
the drk1 potassium channel and Glu1613 in the sodium chan- adding the 5’ and 3’ untranslated regions of a Xenopus β-globin gene.
nel. Perhaps only the voltage-sensing domain in repeat IV of pGEM-HEA was initially created from pGEM-HE 17 by introducing a
T-type calcium channels contains a competent kurtoxin bind- unique AflII restriction site downstream from the 3’ untranslated
ing site. This would explain why our experiments indicate a region of the Xenopus β-globin gene using PCR, verified by DNA
sequencing. α1G was excised from pSP72 by EcoRV digestion. pGEM-
1:1 stoichiometry between toxin and channel, even though HEA was digested with SmaI and HindIII, ends filled in using Klenow
there are four homologous voltage-sensing domains in T-type and blunt-end ligated with the EcoRV fragment of α1G. Correct ori-
calcium channels. Two general implications arise from these entation was initially determined by SacI restriction mapping and
comparisons. The first is that many gating-modifier toxins then verified by DNA sequencing around restriction sites. The α1G
interact with a very similar region of the voltage-sensing -pGEM-HEA construct was linearized with AflII. The human α 1H
domains of voltage-gated ion channels. A Glu residue at the C cDNA 22 was also subcloned into pGEM-HEA by excising it from
terminal edge of S3 (in the S3–S4 loop) seems to be important pcDNA3 (Invitrogen, Carlsbad, California) by digestion with EcoRV
for the binding of hanatoxin, grammotoxin and the α-scorpi- and BspEI, then BspEI and XbaI. Both fragments were ligated into
on toxins. The second implication is that the structure of these SmaI- and XbaI-digested pGEM-HEA. The α 1H-pGEM-HEA con-
struct was linearized with BfrI. To express the rat brain α1B calcium
domains, as seen by the gating modifier toxins, has been high-
channel23 in oocytes, its cDNA was also subcloned into pGEM-HEA.
ly conserved between members of the voltage-gated ion chan- α 1B was excised from pRC/CMV (provided by T. Snutch) by ClaI
nel superfamily. digestion, ends filled in using Klenow, and blunt-end ligated with
The subunit composition of native T-type calcium chan- the Klenow-treated, SmaI/HindIII-digested pGEM-HEA described
nels remains to be determined and has become the subject of above. Correct orientation was initially determined by restriction
some controversy. The α1G and α1H subunits each form calci- mapping using StyI and then verified by DNA sequencing around

672 nature neuroscience • volume 1 no 8 • december 1998


© 1998 Nature America Inc. • http://neurosci.nature.com

articles

restriction sites. The α1B -pGEM-HEA construct was linearized with amino-acid sequencing, the inhibitor was reduced and alkylated.
AflII. Human α1E (pHBE239)26 was linearized with HindIII. Rabbit Approximately 1 nmole of the inhibitor was reduced in 15 mM dithio-
cardiac α1C in the pAGA2 vector containing an N-terminal deletion threitol in 0.4 M ammonium bicarbonate/8 M urea for 30 min at 50ºC.
(DN60)25 was linearized with HindIII. cDNAs encoding rabbit brain The sample was allowed to cool to room temperature and alkylated in
α1A (BI-1; ref. 24), rat brain β1b (ref. 39) and rabbit skeletal muscle 15 mM iodoacetamide for 30 min. The reduced and alkylated inhibitor
α2δ40 were provided by L. Birnbaumer in the pAGA2 vector41. α1A was purified by reversed-phase HPLC and subjected to automated
and α2δ were linearized with XhoI; β1b was linearized with HindIII. Edman degradation on a described system43 resulting in the unequiv-
The rat brain IIA sodium channel α subunit27 and the rat β1 sub- ocal determination of the identity of the first 62 residues, except for
unit28, both in bluescript vectors (provided by D. Hanck and J. Kyle), some uncertainty in the assignment of residues 51 and 57. The reduced
were linearized with XhoI and BamHI, respectively. All transcripts and alkylated inhibitor was digested with chymotrypsin and endo-
of calcium and sodium channel subunits were synthesized using T7 proteinase Asp-N essentially according to the method of ref. 44 and
RNA polymerase. α1G (~1–10 ng per oocyte) and α1H (~100–500 pg the digests separated by reversed-HPLC as described43. Automated
per oocyte) cRNAs were injected alone into oocytes 4 to 9 days before Edman degradation of the overlapping chymotryptic peptide,
recording. α1A, α1B, α1C and α1E cRNAs (~100 pg to 1 ng per oocyte) CQGLPDNAR (residues 46–54) and the overlapping C-terminal endo-
were injected with approximately equal quantities of β 1b and α 2δ proteinase Asp-N peptide DNARIKRSGRCRA (residues 51–63)
cRNAs 24 to 48 h before recording. The rat brain IIA α subunit and resolved the uncertainty in residues 51 and 57 and identified the C-
β1 cRNAs were co-injected at equal concentrations (~100 pg per terminus of the inhibitor. The calculated average molecular weight of
oocyte) 24–48 h before recording. intact inhibitor, 7386.5 daltons, agrees with the experimentally deter-
ELECTROPHYSIOLOGY. Oocytes from Xenopus laevis frogs were removed mined molecular weight of 7386.1. An extinction coefficient (280 nm)
surgically and incubated with agitation for 1.5 h in a solution con- of 2.3 x 104 per M per cm was determined by quantitative amino-acid
© 1998 Nature America Inc. • http://neurosci.nature.com

taining NaCl (82.5 mM), KCl (2.5 mM), MgCl 2 (1 mM), HEPES analysis and used to measure toxin concentration.
(5 mM) and collagenase (2 mg/ml; Worthington Biochemical Corp.,
Freehold, New Jersey), pH 7.6 with NaOH. Defolliculated oocytes Acknowledgements
were injected with cRNA and incubated at 17°C in a solution con- We thank Zhe Lu, Chinfei Chen and Yingying Li-Smerin for discussions.
taining NaCl (96 mM), KCl (2 mM), MgCl2 (1 mM), CaCl2 (1.8 mM),
HEPES (5 mM) and gentamicin (50 µg/ml; Gibco BRL, Gaithersburg, RECEIVED 28 JULY: ACCEPTED 16 O CTOBER 1998
Maryland), pH 7.6 with NaOH. Oocyte membrane voltage was con-
trolled using an OC-725C oocyte clamp (Warner Instruments). Data
were filtered at 1–2 kHz (8-pole Bessel) and digitized at 10-20 kHz. 1. Llinas, R. & Yarom, Y. Properties and distribution of ionic conductances
Microelectrode resistances were 0.1 to 0.8 MΩ when filled with 3 M generating electroresponsiveness of mammalian inferior olivary neurons
KCl. Oocytes were studied in a 200 µl recording chamber perfused in vitro. J. Physiol. (Lond.) 315, 569–584 (1981).
2. Carbone, E. & Lux, H. D. A low voltage activated, fully inactivating Ca 2+
with one of two different extracellular solutions. The extracellular channel in vertebrate sensory neurones. Nature 310, 501–511 (1984).
solution for recording of calcium channel currents contained BaOH 2 3. Bossu, J. L., Feltz, A. & Thomann, J. M. Depolarization elicits two distinct
(5 mM), NaCH3SO 3 (100 mM) and HEPES (10 mM), pH 7.6 with calcium currents in vertebrate sensory neurones. Pflügers Arch. 403,
NaOH. Contamination of calcium-activated chloride current was 360–368 (1985).
minimized by recording in chloride-free solution and recording bar- 4. Bean, B. P. Two kinds of calcium channels in canine atrial cells. J. Gen.
Physiol. 86, 1–30 (1985).
ium currents usually less than 2 µA (mostly ~ 1 µA). Sodium chan- 5. Armstrong, C. M. & Matteson, D. R. Two distinct populations of calcium
nel currents were recorded in an extracellular solution containing channels in a clonal line of pituitary cells. Science 227, 65–67 (1985).
NaCl (100 mM), MgCl 2 (1 mM), CaCl 2 (0.3 mM) and HEPES 6. Nilius, B., Hess, P., Lansman, J. B. & Tsien, R. W. A novel type of cardiac
(10 mM), pH 7.6 with NaOH. Cytochrome C (1 mg/ml) was includ- calcium channel in ventricular cells. Nature 316, 443–446 (1985).
ed in the extracellular recording solution for experiments examining 7. Kostyuk, P. G., Shuba, Y. M. & Savchenko, A. N. Three types of calcium
channels in the membranes of mouse sensory neurons. Pflügers Arch. 411,
the concentration dependence of toxin inhibition (Fig. 5). Agar salt 611–669 (1988).
bridges containing 1 M NaCl were used to connect the ground elec- 8. Chen, C. & Hess, P. Mechanism of gating of T-type calcium channels. J.
trode pools and the recording chamber. All experiments were done Gen. Physiol. 96, 603–630 (1990).
at room temperature (~22–25°C). 9. Hagiwara, N., Irisawa, H. & Kameyama, M. Contribution of two types of
calcium currents to the pacemaker potentials of rabbit sino-atrial node
TOXIN PURIFICATION AND BIOCHEMICAL CHARACTERIZATION. Lyophilized cells. J. Physiol. (Lond.) 395, 233–253 (1988).
Parabuthus transvaalicus venom (Latoxan, France) was dissolved in 10. Jahnsen, H. & Llinas, R. J. Ionic basis for the electro-responsiveness and
25% aqueous acetonitrile with 0.1% trifluoroacetic acid (TFA) at a oscillatory properties of guinea-pig thalamic neurones in vitro. J. Physiol.
(Lond.) 349, 227–247 (1984).
concentration of 5 mg/ml. The resulting solution was filtered (0.2 µm, 11. Huguenard, J. R. Low-threshold calcium currents in central nervous
Gelman, Acrodisc), concentrated under vacuum and fractionated in system neurons. Annu. Rev. Physiol. 58, 329–348 (1996).
two steps using a Beckman reversed-phase HPLC. The first separation 12. Cohen, C. J., McCarthy, R. T., Barrett, P. Q. & Rasmussen, H. Ca2+
employed a 1 × 25cm C-18 (5 µ, 90 Å) ultrasphere column (Beckman channels in adrenal glomerulosa cells: K+ and angiotensin II increase T-
type Ca2+ current. Proc. Natl. Acad. Sci. USA 85, 2412–2416 (1988).
Inst., Columbia, Maryland) eluted with a linear gradient from 25% 13. Chen, C., Corbley, M., Roberts, T. & Hess, P. Voltage-sensitive calcium
mobile phase B to 100% mobile phase B over 60 min at a flow rate of channels in normal and transformed 3T3 fibroblasts. Science 239,
3 ml/min, where mobile phase A was 0.1% TFA in H20 and mobile 1024–1026 (1988).
phase B was 0.08% TFA in acetonitrile. The fraction with highest activ- 14. Santi, C. M., Darszon, A. & Hernandez-Cruz, A. A dihydropyridine-
ity was collected at about 21 min (Fig. 1a). This active fraction was sensitive T-type Ca2+ current is the main Ca2+ current carrier in mouse
primary spermatocytes. Am. J. Physiol. 271, C1583–C1593 (1996).
subsequently fractionated using a 0.46 × 25cm C-18 (5 µ, 90 Å) ultra- 15. Bezprozvanny, I. & Tsien R. W. Voltage-dependent blockade of diverse
sphere column (Beckman Inst., Columbia, Maryland) eluted with a types of voltage-gated Ca2+ channels expressed in Xenopus oocytes by the
linear gradient from 40% mobile phase B to 100% mobile phase B over Ca2+ channel antagonist mibefradil (Ro 40–5967). Mol. Pharmacol. 48,
60 min at a flow rate of 0.9 ml/min, where mobile phase A was 0.1% 540–549 (1995).
16. Perez-Reyes, E. et al. Molecular characterization of a neuronal low-
TFA in H 20 and mobile phase B was 0.08% TFA in methanol. The voltage-activated T-type calcium channel. Nature 391, 896–900 (1998).
inhibitory activity eluting at about 22 min (Fig. 1b) seemed to coin- 17. Liman, E. R., Tytgat, J. & Hess, P. Subunit stoichiometry of a mammalian
cide with a symmetrical peak after re-chromatography on the same K+ channel determined by construction of mutimeric cDNAs. Neuron 9,
column (Fig. 1c) and was therefore chemically analyzed further. The 861–871 (1992).
mass of the inhibitor was determined on a described42 microbore LC- 18. Meves, H., Simard, J. M. & Watt, D. D. Interactions of scorpion toxins
with the sodium channel. Ann. NY Acad. Sci. 479,113–132 (1986).
electrospray ionization mass spectrometry system to be 7386.1 dal- 19. McDonough, S. I., Lampe, R. A., Keith, R. A. & Bean, B. P Voltage-
tons (average of [M+4H]4+ ion at m/z 1847.3, [M+5H]5+ ion at m/z dependent inhibition of N- and P-type calcium channels by the peptide
1478.3, most intense ion, and [M+6H] 6+ ion at m/z 1232.1). Before toxin w-grammotoxin-SIA. Mol. Pharmacol. 52, 1095–1104 (1997).

nature neuroscience • volume 1 no 8 • december 1998 673


© 1998 Nature America Inc. • http://neurosci.nature.com

articles

20. McDonough, S. I., Mintz, I. M. & Bean, B. P. Alteration of P-type calcium anemone toxin in the S3-S4 extracellular loop in domain IV of the Na +
channel gating by the spider toxin w-Aga-IVA. Biophys. J. 72, 2117–2128 channel α subunit. J. Biol. Chem. 271, 15950–15962 (1996).
(1997). 33. Li-Smerin Y. & Swartz, K. J. Gating modifier toxins recognize a conserved
21. Swartz, K. J. & MacKinnon, R. Hanatoxin modifies the gating of a voltage- structural motif in voltage-gated Ca 2+ and K+ channels. Proc. Natl. Acad.
dependent K+ channel through multiple binding sites. Neuron 18, Sci. USA 95, 8585–8589 (1998).
665–673. (1997). 34. Swartz, K. J. & MacKinnon, R. Mapping the receptor site for hanatoxin, a
22. Cribbs, L. L. et al. Cloning and characterization of α1H from human heart, gating modifier of voltage-dependent K+ channels. Neuron 18, 675–682
a member of the T-type Ca 2+ channel gene family. Circ. Res. 83, 103–109 (1997).
(1998). 35. Soong, T. W. et al. Structure and functional expression of a member of the
23. Dubel, S. J. et al. Molecular cloning of the α-1 subunit of an ω-conotoxin- low voltage-activated calcium channel family. Science 260, 1133–1136 (1993).
sensitive calcium channel. Proc. Natl. Acad. Sci. USA 89, 5058–5062 36. Bourinet, E. et al. The α1E calcium channel exhibits permeation
(1992). properties similar to low-voltage-activated calcium channels. J. Neurosci.
24. Mori, Y. et al. Primary structure and functional expression from 16, 4983–4993 (1996).
complementary DNA of a brain calcium channel. Nature 350, 398–402 37. Meir, A. & Dolphin, A. C. Known calcium channel α1 subunits can form
(1991). low threshold small conductance channels with similarities to native T-
25. Wei, X. et al. Increase in Ca2+ channel expression by deletions at the type channels. Neuron 20, 341–351.
amino terminus of the cardiac α1C subunit. Receptors Channels 4, 38. Bean, B. P. & McDonough, S. I. Two for T. Neuron 20, 825–828 (1998).
205–215 (1996). 39. Pragnell, M., Sakamoto, J., Jay, S. D. & Campbell, K. P. Cloning and
26. Schneider, T. et al. Molecular analysis and functional expression of the tissue-specific expression of the brain calcium channel β-subunit. FEBS
human type E neuronal Ca2+ channel α1 subunit. Receptors Channels 2, Lett. 291, 253–258 (1991).
255–270 (1994). 40. Ellis, S. B. et al. Sequence and expression of mRNAs encoding the α1 and α2
27. Auld, V. J. et al. A rat brain Na+ channel α subunit with novel gating subunits of a DHP sensitive calcium channel. Science 241, 1661–1664 (1988).
properties. Neuron 1, 449–461 (1988). 41. Wei, X. et al. Heterologous regulation of the cardiac Ca2+ channel α1
28. Isom, L. L. et al. Primary structure and functional expression of the β1 subunit by skeletal muscle β and γ subunits. Implications for the structure
© 1998 Nature America Inc. • http://neurosci.nature.com

subunit of the rat brain sodium channel. Science 256, 839–842 (1992). of cardiac L-type Ca 2+ channels. J. Biol. Chem. 266, 21943–21947 (1991).
29. Catterall, W. A. Binding of scorpion toxin to receptor site associated with 42. Jaffe, H., Veeranna Shetty, K. T. & Pant, H. C. Characterization of the
sodium channels in frog muscle: correlation of voltage-dependent phosphorylation sites of human high molecular weight neurofilament
binding with activation. J. Gen. Physiol. 74, 357–391 (1979). protein by electrospray ionization tandem mass spectrometry and
30. Wang, G. K. & Strichartz, G. Kinetic analysis of the action of Leiurus database searching. Biochemistry 37, 3931–3940 (1998).
scorpion toxin on ionic currents in myelinated nerve. J. Gen. Physiol. 86, 43. Seok, Y-J et al. High affinity binding and allosteric regulation of
739–762 (1985). Escherichia coli glycogen phosphorylase by the histidine phosphocarrier
31. Gonoi, T. & Hille, B. Gating of Na+ channels: inactivation modifiers protein, HPr. J. Biol. Chem. 272, 26511–26521 (1997).
discriminate among models. J. Gen. Physiol. 89, 253–274 (1987). 44. Stone, K. L. & Williams, K. R. in A Practical Guide to Protein and Peptide
32. Rogers, J. C., Qu, Y., Tanada, T. N., Scheuer, T. & Catterall, W. Molecular Purification for Microsequencing (ed. Matsudaira, P.) 43–69 (Academic,
determinants of high affinity binding of α-scorpion toxin and sea San Diego, CA, 1993).

674 nature neuroscience • volume 1 no 8 • december 1998

You might also like