ArticlePDF Available

The Sea Anemone Toxins BgII and BgIII Prolong the Inactivation Time Course of the Tetrodotoxin-Sensitive Sodium Current in Rat Dorsal Root Ganglion Neurons

Authors:

Abstract and Figures

We have characterized the effects of BgII and BgIII, two sea anemone peptides with almost identical sequences (they only differ by a single amino acid), on neuronal sodium currents using the whole-cell patch-clamp technique. Neurons of dorsal root ganglia of Wistar rats (P5-9) in primary culture (Leibovitz's L15 medium; 37 degrees C, 95% air/5% CO2) were used for this study (n = 154). These cells express two sodium current subtypes: tetrodotoxin-sensitive (TTX-S; K(i) = 0.3 nM) and tetrodotoxin-resistant (TTX-R; K(i) = 100 microM). Neither BgII nor BgIII had significant effects on TTX-R sodium current. Both BgII and BgIII produced a concentration-dependent slowing of the TTX-S sodium current inactivation (IC50 = 4.1 +/- 1.2 and 11.9 +/- 1.4 microM, respectively), with no significant effects on activation time course or current peak amplitude. For comparison, the concentration-dependent action of Anemonia sulcata toxin II (ATX-II), a well characterized anemone toxin, on the TTX-S current was also studied. ATX-II also produced a slowing of the TTX-S sodium current inactivation, with an IC50 value of 9.6 +/- 1.2 microM indicating that BgII was 2.3 times more potent than ATX-II and 2.9 times more potent than BgIII in decreasing the inactivation time constant (tau(h)) of the sodium current in dorsal root ganglion neurons. The action of BgIII was voltage-dependent, with significant effects at voltages below -10 mV. Our results suggest that BgII and BgIII affect voltage-gated sodium channels in a similar fashion to other sea anemone toxins and alpha-scorpion toxins.
Content may be subject to copyright.
The Sea Anemone Toxins BgII and BgIII Prolong the
Inactivation Time Course of the Tetrodotoxin-Sensitive Sodium
Current in Rat Dorsal Root Ganglion Neurons
EMILIO SALCEDA, ANOLAND GARATEIX, and ENRIQUE SOTO
Instituto de Fisiologı´a, Universidad Auto´ noma de Puebla, Me´ xico (Em.S., En.S.); and Instituto de Oceanologı´a, Ministerio de Ciencia,
Tecnologia y Medio Ambieute, La Habana, Cuba (A.G.)
Received May 23, 2002; accepted September 4, 2002
ABSTRACT
We have characterized the effects of BgII and BgIII, two sea
anemone peptides with almost identical sequences (they only
differ by a single amino acid), on neuronal sodium currents
using the whole-cell patch-clamp technique. Neurons of dorsal
root ganglia of Wistar rats (P5-9) in primary culture (Leibovitzs
L15 medium; 37°C, 95% air/5% CO
2
) were used for this study
(n154). These cells express two sodium current subtypes:
tetrodotoxin-sensitive (TTX-S; K
i
0.3 nM) and tetrodotoxin-
resistant (TTX-R; K
i
100
M). Neither BgII nor BgIII had
significant effects on TTX-R sodium current. Both BgII and BgIII
produced a concentration-dependent slowing of the TTX-S
sodium current inactivation (IC
50
4.1 1.2 and 11.9 1.4
M, respectively), with no significant effects on activation time
course or current peak amplitude. For comparison, the concen-
tration-dependent action of Anemonia sulcata toxin II (ATX-II), a
well characterized anemone toxin, on the TTX-S current was
also studied. ATX-II also produced a slowing of the TTX-S
sodium current inactivation, with an IC
50
value of 9.6 1.2
M
indicating that BgII was 2.3 times more potent than ATX-II and
2.9 times more potent than BgIII in decreasing the inactivation
time constant (
h
) of the sodium current in dorsal root ganglion
neurons. The action of BgIII was voltage-dependent, with sig-
nificant effects at voltages below 10 mV. Our results suggest
that BgII and BgIII affect voltage-gated sodium channels in a
similar fashion to other sea anemone toxins and
-scorpion
toxins.
Voltage-dependent sodium channels have receptor sites
that may be recognized by different groups of neurotoxins
(Adams and Olivera, 1994; Trainer et al., 1994). For this
reason, this kind of compound has been shown to be a useful
pharmacological tool for studying the functional and struc-
tural mapping of sodium channel proteins (Catterall, 2000).
Toxins that act on sodium channels can be classified in two
main groups according to their pharmacological effects on the
channel: blockers (e.g., tetrodotoxin and
-conotoxin) and
modulators (e.g., batrachotoxin,
- and
-scorpion toxins, and
sea anemone toxins). The latter is further divided into sev-
eral classes based upon the effects on channel activation and
inactivation kinetics (Narahashi, 1998).
A large number of neurotoxins have modulatory actions on
Na
channel function by modifying the processes linked to
channel activation and inactivation, ionic selectivity, and
other properties involved in action potential generation. At
least six neurotoxin receptor sites have been identified on the
mammalian sodium channel (Strichartz et al., 1987; Catter-
all, 1995). Anemone peptide neurotoxins and
-scorpion tox-
ins share receptor site 3 on sodium channels (Couroud et al.,
1978; Catterall, 1995, 2000; Gordon et al., 1998), which in-
volves the extracellular loops IS5-S6, IVS3-S4, and IVS5-S6
of the ionic channel (Rogers et al., 1996). Anemone toxins are
generally smaller than the structurally unrelated
-scorpion
toxins and have three rather than four disulfide bridges
(Norton, 1997); nevertheless, both groups bind to site 3, ex-
hibit similar pharmacological properties, displace one an-
other from their binding site, and their main effect is to delay
channel inactivation, resulting in a prolongation of the action
potential.
Bunodosoma granulifera is an anemone species very com-
mon at the Cuban seashores. Several active compounds with
pharmacological actions on ionic channels have been isolated
from its secretions (Aneiros et al., 1993; Loret et al., 1994;
Dauplais et al., 1997; Salinas et al., 1997; Alessandri-Haber
et al., 1999; Garateix et al., 2000). Among them, BgII and
BgIII are two peptide toxins (molecular masses: 5072 and
5073, respectively) with almost identical sequences (they
only differ by a single amino acid), causing toxicity in mice
This work was supported by CONACyT Grant E120.1869/2000
Article, publication date, and citation information can be found at
http://jpet.aspetjournals.org.
DOI: 10.1124/jpet.102.038570.
ABBREVIATIONS: DRG, dorsal root ganglion; TTX-S, tetrodotoxin-sensitive; TTX-R, tetrodotoxin-resistant; L15, Leibovitzs L15 medium;
h
,
inactivation time constant; ATX-II, Anemonia sulcata toxin II; G
Na
, peak sodium conductance.
0022-3565/02/3033-1067–1074$7.00
THE JOURNAL OF PHARMACOLOGY AND EXPERIMENTAL THERAPEUTICS Vol. 303, No. 3
Copyright © 2002 by The American Society for Pharmacology and Experimental Therapeutics 38570/1026540
JPET 303:1067–1074, 2002 Printed in U.S.A.
1067
when injected intracerebroventricularly and markedly differ-
ent binding to rat brain synaptosomes; both effects are
higher for BgII (Loret et al., 1994).
In this work, we have characterized the effects of BgII and
BgIII toxins on neuronal sodium currents. Rat dorsal root
ganglion (DRG) neurons were chosen for this study since
these cells express two sodium current subtypes: a tetrodo-
toxin-sensitive (TTX-S) sodium current, which is readily
blocked by TTX (K
I
0.3 nM), and a tetrodotoxin-resistant
(TTX-R) sodium current, which is highly resistant to TTX (K
i
100
M) (Roy and Narahashi, 1992; Novakovik et al.,
2001). To our knowledge, this is the first electrophysiological
characterization of the action of these toxins on neuronal
cells.
Materials and Methods
Animal care and procedures were carried out in accordance with
the Declaration of Helsinki. The number of animals used for this
work was kept to the minimum necessary for a meaningful interpre-
tation of the data.
Toxins. BgII and BgIII were isolated and purified from sea anem-
one B.granulifera, as previously described (Aneiros et al., 1993;
Loret et al., 1994). Some experiments were performed with ATX-II
obtained from the sea anemone Anemonia sulcata (a gift from Pro-
fessor L. Beress, Kiel, Germany). TTX was obtained from Sigma-
Aldrich (St. Louis, MO). Aliquots of stock solution in deionized water
were prepared and stored in a freezer (20°C). Before each experi-
ment, they were dissolved in the perfusion solution.
Cell Preparation. Young Wistar rats (P5-9) of either sex were
anesthetized with ether and subsequently decapitated. DRGs were
isolated from the vertebral column and incubated (30 min at 37°C) in
Leibovitzs L15 medium (L15) (Invitrogen, Carlsbad, CA) containing
1.125 mg/ml trypsin and 1.125 mg/ml collagenase (both from Sigma-
Aldrich). Following enzyme treatment, the ganglia were washed 3
times with sterile L15 and mechanically dissociated. The cells were
then plated onto 35-mm culture dishes (Corning, Corning, NY) con-
taining 12 10-mm glass coverslips (Corning) previously coated
with poly-D-lysine (Sigma-Aldrich). Neurons were incubated 4 to 6 h
in a humidified atmosphere (95% air/5% CO
2
at 37°C, using a CO
2
water-jacketed incubator; Nuaire, Plymouth, MN) to allow the iso-
lated cells to settle and adhere to the coverslips. The incubation
medium (pH 7.4) contained L15, 15.7 mM NaHCO
3
(Merck, Naucal-
pan, Mexico), 10% fetal bovine serum, 2.5
g/ml Fungizone (both
from Invitrogen), 100 U/ml penicillin (Lakeside, Toluca, Mexico), and
15.8 mM HEPES (Sigma-Aldrich).
Electrophysiological Recording. A coverslip with attached
neurons was transferred to a 500-
l perfusion chamber mounted on
the stage of an inverted phase-contrast microscope (Nikon Diaphot,
Tokyo, Japan). Cells were bathed with an external solution contain-
ing 20 mM NaCl, 1 mM MgCl
2
, 1.8 mM CaCl
2
, 45 mM TEA-Cl, 70
mM choline chloride, 10 mM 4-aminopyridine, and 5 mM HEPES.
The pH of this solution was adjusted to 7.4 with HCl. Osmolarity was
monitored by a vapor pressure osmometer (Wescor, Logan, UT) and
adjusted to 290 mOsm using dextrose. A gravity-driven perfusion
system maintained the external solution flowing into the chamber at
a rate of around 100
l/min. In addition to this perfusion system, a
double-barrel array built up with borosilicate glass capillaries
(TW1203; WPI, Sarasota, FL) was placed approximately 40
m
above the cell under study; each barrel was coupled to an indepen-
dent syringe driven by a Baby Bee pump (BAS, West Lafayette, IN).
Through this system, the neuron was continuously microperfused
(10
l/min) with external solution or with external solution plus
toxin. Some experiments were designed to study the effects of BgII or
BgIII on TTX-R sodium currents. To achieve this, TTX (300 nM) was
added to both bath and microperfusion solutions.
The whole-cell patch-clamp technique was used to record ionic
currents. Patch pipettes were pulled from borosilicate glass capillar-
ies (TW1203; WPI), using a Flaming-Brown electrode puller (P80/
PC; Sutter Instruments, San Rafael, CA), which had resistances of
0.9 to 1.8 Mwhen filled with internal solution. Pipette solution
contained 10 mM NaCl, 100 mM CsF, 30 mM CsCl, 10 mM TEA-Cl,
8 mM EGTA, and 5 mM HEPES. The pH of this solution was
adjusted to 7.3 with CsOH. Osmolarity was adjusted to 300 mOsm.
The internal solution was filtered on the day of use with a 0.22-
m
pore size syringe filter (Millipore, Bedford, MA).
To measure ionic currents, an Axopatch-1D amplifier (Axon In-
struments, Foster City, CA) was used. Command pulse generation
and data sampling were controlled by the PClamp 8.0 software (Axon
Instruments) using a 16-bit data acquisition system (Digidata
1320A; Axon Instruments). Signals were low-pass filtered at 5 kHz
and digitized at 20 kHz. Leakage and capacitive currents were dig-
itally subtracted with the P-P/2 method. Capacitance and series
resistance (80%) were electronically compensated. Experiments were
rejected when, at the maximum peak current, the voltage error
exceeded 5 mV after compensation of series resistance. No correc-
tions were made for smaller values.
The type of sodium current present in the cell under study was
determined before each experiment. In accordance with the criterion
used by Strachan et al. (1999), only those cells with 10% TTX-R
sodium current, as derived from a steady-state inactivation profile,
were accepted to determine the effects of BgII and BgIII on TTX-S
sodium currents. Experiments were performed at room temperature
(2325°C).
Data Analysis. Recordings were analyzed off-line using PClamp
8.0 and Origin software (Microcal Software, Northampton, MA).
Statistical differences were determined using a Studentsttest with
p0.05. Curve-fitting routines were performed using a nonlinear
least-squares method. Numerical data are presented as the mean
S.E. for at least four measurements.
Concentration-response curves were obtained by measuring the
parameters under study in sodium currents elicited by a single-step
voltage protocol, where 40 ms depolarizing test pulses to 10 mV
were applied from a holding potential of 90 mV every 8 s. Data
were then plotted as a function of toxin concentration and fit by the
following function.
yA1A2A1
110log IC50 xP
where A
1
is the yvalue at the bottom plateau, A
2
is the yvalue at the
top plateau, log IC
50
is the xvalue when the response is halfway
between A
1
and A
2
, and Pis the Hill slope.
To test whether BgII or BgIII showed a use-dependent action,
from a holding potential of 90 mV repetitive pulses to 10 mV were
applied at frequencies of 0.1, 1 and 5 Hz. Current-voltage relation-
ships and availability curves were constructed using a standard
double-pulse protocol; from a holding potential of 100 mV, a 40-ms
test pulse to 10 mV was preceded by 40-ms prepulses between
100 and 70 mV (time interval between sweeps 8 s).
The peak amplitudes of the currents were measured at the pre-
pulse and converted to sodium conductance by means of the following
equation:
GNa INa
Vtest Vrev
where G
Na
is the sodium conductance, I
Na
is the sodium current peak
amplitude, V
test
is the test potential, and V
rev
is the reversal poten-
tial for the sodium current. Normalized conductance curves were
then plotted and fit by a Boltzmann distribution with the following
function.
GNa
Gmax 1
1exp
Vtest V1/ 2
k
1068 Salceda et al.
where G
max
is the maximum sodium conductance, V
1/2
is the test
potential that activates 50% of the sodium channels, kis the slope of
the curve, and the other terms are as above.
The steady-state inactivation parameter (h
) was calculated by
dividing the current achieved following a given prepulse by the
maximum current achieved in the test pulse. The results of such
operation were plotted as a function of the prepulse potential and
were fitted by the following Boltzmann function.
h1
1exp
Vpre V1/ 2 inact
k
where V
pre
is the prepulse potential, V
1/2 inact
is the prepulse poten-
tial at which h
is 0.5, and kis the slope of the curve at this
mid-point.
The effects of BgII and BgIII on the rate at which sodium channels
recover from inactivation were investigated using a two-pulse proto-
col with a variable interpulse interval (t) as follows: from a holding
potential of 100 mV, a 40-ms conditioning prepulse to 10 mV was
used to inactivate sodium channels, after which a 2.5-ms depolariz-
ing test pulse to 10 mV was applied. The interpulse interval be-
tween the conditioning and the test pulses was varied between 1.25
and 80 ms. The peak current recorded during the test pulse was
normalized against the current amplitude during the conditioning
prepulse and plotted as a function of t. Data obtained from this
protocol were fitted by a single exponential function.
Results
A total of 154 neurons (mean capacity 52.5 18 pF, S.D.)
were successfully voltage-clamped for a sufficient time to
allow the study of BgII, BgIII, and ATX-II actions. The ca-
pacitances of the DRG neurons used for this study formed a
unimodal histogram with the mean, which corresponds to a
cell diameter of about 41
m. This distribution represents
only the cells selected for recording, basically neurons with a
medium size cell body (probably A
neurons).
The percentage of change in peak amplitude and activation
and inactivation time constants were calculated for ionic
currents from both TTX-S and -R sodium currents before and
about 1 min after perfusion with toxin. For the TTX-S cur-
rent subtype, concentration-response curves were obtained
using concentrations of 0.1, 0.3, 1, 3, 10, and 30
M. For the
rest of the experiments in this study, we used a 10
M toxin
concentration.
To study the effects of BgII and BgIII on TTX-R sodium
currents, TTX (300 nM) was added to both bath and mi-
croperfusion solutions (n15). This procedure made evident
the existence of two subtypes of TTX-R current: 1) a slowly
activating (
0.46 0.2 ms) and slowly inactivating (
4.55 1.96 ms) current that activates at about 40 mV (n
7); and 2) a current that fails to inactivate, giving rise to a
large late component (n8). These two types of TTX-R
currents coincide with those previously described in DRG
neurons (Bossu and Feltz, 1984; Baker and Wood, 2001).
Neither 10
M BgII (n9) nor 10
M BgIII (n6) had
significant effects (p0.05, Studentsttest) on either TTX-R
sodium current subtypes (Fig. 1).
Both BgII (n87) and BgIII (n22) produced a concen-
tration-dependent effect on the TTX-S sodium current inac-
tivation time course, with no significant effects (p0.05,
Studentsttest) on activation time course or current peak
amplitude (Fig. 2). The inactivation time course of TTX-S
sodium currents was adjusted with an exponential function
over the following 10 ms after peak current. In the presence
of 3
M BgII, the inactivation time constant (
h
) increased
97.6 35.4%, whereas a 10
M toxin concentration produced
154.2 17.4% increase in
h
. The IC
50
value for BgII was
4.1 1.2
M, with a fixed slope value of 1.0. In contrast, the
action of BgIII was less effective; at 3
M,
h
was increased
15.2 8.5%, whereas at 10
M,
h
increased 60.9 13.0%.
The IC
50
value for BgIII experiments was 11.9 1.4
M,
with a fixed slope of 1.0.
To compare the effect of these toxins with ATX-II, a well
Fig. 1. Effects of 10
M BgII (A and C) and BgIII (B and D) on TTX-R sodium currents (n15). Recordings were made in presence of 300 nM TTX.
Currents were elicited by a single-step voltage protocol (40-ms depolarizing test pulses to 10 mV from a holding potential of 90 mV every 8 s). Two
subtypes of TTX-R current were found, partially inactivating currents with a time constant
4.55 1.96 ms (A and B; n7) and very slow
activating currents (
1.38 0.2 ms) failing to inactivate (C and D; n8). For each case, two superimposed traces are presented, under control
conditions and approximately 2 min after toxin perfusion. Neither BgII (n9) nor BgIII (n6) produced significant effects on either TTX-R sodium
current subtypes.
BgII and BgIII Anemone Toxins 1069
characterized sea anemone toxin, an experimental series was
done to study the effect of ATX-II on the DRG neurons (Fig.
2, C and D). ATX-II increased the inactivation time constant
of the Na
TTX-S current with an IC
50
value of 9.6 1.3
M,
indicating that BgII is 2.3 more potent than ATX-II and 2.9
times more potent than BgIII in decreasing the
h
of the Na
current. The maximum effect of both BgII and BgIII on the
TTX-S sodium current inactivation time course was always
achieved within the first minute after perfusion with either
toxin, and once reached, it was stable throughout the expo-
sure period to these agents (about 100 s). Washout of toxin
effects was complete for both BgII and BgIII at 10
M and
took less than a minute (see inserts in Fig. 2). The slowing in
the inactivation time course elicited by 10
M BgII (n4)
was voltage-dependent, whereas it was not for 10
M BgIII
(n4). Analysis of
h
versus voltage curves showed that the
effect of 10
M BgII on the inactivation time constant of the
TTX-S Na
current was significant (p0.05, Studentst
test) only for voltages under 10 mV. The increase in
h
induced by BgII at 10 mV was 64% smaller than that
produced at 30 mV. At voltages above 10 mV, although
there is a tendency of
h
to increase, the change was not
significant. The action of both toxins was not use-dependent
(data not shown).
From current-voltage relationships, current density versus
voltage curves were obtained by normalizing ionic current
amplitudes as a function of membrane capacity (Fig. 3). Un-
der control conditions, the maximum current density
(138 24 pA/pF) was achieved at 10 mV. Perfusion with
10
M BgII (Fig. 3A) produced a nonsignificant decrease (p
0.05, Studentsttest) in the current density (101 12
pA/pF). Perfusion with BgIII (Fig. 3B) produced a nonsignif-
icant increase in the current density. The reversal potential
remained unchanged in the presence of toxins, and it was
Fig. 2. Concentration-response curves of the effect of BgII, BgIII, and ATX-II. A and B, typical experiments showing the effects of 10
M BgII and BgIII
on TTX-S sodium currents, respectively. Traces were obtained by a single-step voltage protocol (40-ms depolarizing test pulses to 10 mV from a
holding potential of 90 mV every 8 s). Toxins produced a marked slowing of the inactivation process. The insets show the time course of toxin effects
on
h
. Bars in the inserts indicate the time interval of toxin perfusion around the cell. C and D, concentration-response curves of the effect of ATX-II
(n32; open circles), BgII (n87; triangles), and BgIII (n22; filled circles) on
h
. Data were fit by (solid lines) the dose-response function described
under Materials and Methods. The IC
50
values calculated from these experiments were 4.1 1.2
M (BgII), 9.6 1.3 (ATX-II), and 11.9 1.4
M
(BgIII). Asterisks denotes a significant effect of BgII and BgIII with respect to its controls. ⫹⫹ denotes significant effects with respect to ATX-II action
(Studentsttest; p0.05). Points represent the mean standard error of the mean.
1070 Salceda et al.
consistent with the one calculated from the Nernst equation
(18 mV).
The peak sodium conductance (G
Na
) at several potential
values was calculated as a chord conductance from the cor-
responding peak current. Normalized conductance curves
were then fitted by a Boltzmann distribution and the corre-
sponding V
1/2
and slope were calculated for each curve. Fig-
ure 4 shows the voltage dependence of G/G
max
under control
conditions and after perfusion with 10
M of either BgII or
BgIII. The mean value of V
1/2
for control experiments was
22 0.8 mV. No significant differences (p0.05, Students
ttest) in the G
Na
were produced when BgII or BgIII were
applied.
Measurements of the voltage dependence of h
(steady-
state inactivation parameter) were made using a two-pulse
protocol. It was found that 10
M of either BgII or BgIII
caused a significant (p0.05, Studentsttest) hyperpolar-
izing shift in the voltage at which half of the channels are
inactivated (V
1/2 inact
), from a control value of 61 0.8 mV
to 73 0.9 mV after BgII and 69 1.4 mV after BgIII
application (Fig. 5). The calculated slopes were 7.7 0.6 mV
(control), 8.8 0.8 mV (BgII), and 10 1.6 mV (BgIII). The
slope of the curve for BgIII indicates that steady-state inac-
tivation became significantly less voltage-dependent (p
0.05, Studentsttest). The hyperpolarization shift of the Na
current implies that at 70 mV the current availability in
the presence of BgIII is 82% of the control and that in the
presence of BgII it is 70% of the control value.
It has been shown that the recovery rate of the TTX-S
current could be described by the sum of two exponential
functions: a fast component (
f
) with a time constant of a few
milliseconds and a slow component (
s
) with a time constant
in the order of several hundred milliseconds. Elliot and Elliot
(1993) pointed out the inherent difficulty in the characteriza-
tion of the slow component, which would require the use of
interpulse intervals lasting several seconds. For this reason,
we defined
hrec
as the time required to reach 63% of reacti-
vated channels. Under control conditions,
hrec
was 15.5 2
Fig. 3. Effects of BgII and BgIII on the current density versus voltage
relationships. Curves were obtained by normalizing TTX-S current am-
plitudes to membrane capacity. Ionic currents were obtained by a voltage
protocol in which 40-ms pulses between 100 and 70 mV were applied
from a holding potential of 100 mV. Under control conditions (n4 for
each case), the maximum current density was achieved at 10 mV. BgII
(10
M) (A; n4) produced a nonsignificant decrease in the current
density at this voltage (101 12 pA/pF). When 10
M BgIII (B; n4)
was perfused, the maximum current density was larger (173 44 pA/pF),
although not significantly when compared with control values (p0.05;
Studentsttest).
Fig. 4. Effects of BgII and BgIII on the voltage dependence of sodium
conductance. A, typical experiments from which activation curves were
obtained. B, the voltage dependence of G
Na
under control conditions and
after perfusion with BgII (n4) or BgIII (n4), each at a 10
M
concentration. Data were fit by a Boltzmann function (solid lines). The
mean value of V
1/2
for control experiments was 22 0.8 mV. After 10
M BgII application, the V
1/2
was 24 0.9 mV. After perfusion with 10
M BgIII, the V
1/2
was 25 1 mV. The slope for the three curves were
also very similar, around 6.5 mV.
BgII and BgIII Anemone Toxins 1071
ms. No significant effects (p0.05, Studentsttest) were
observed when 10
M BgII (n5) or 10
M BgIII (n4)
were applied. The values of
hrec
were 17.5 1.2 and 16.4
1.8 ms, respectively (data not shown).
Discussion
In the present work, we have made a comparative study of
the effects of BgII, BgIII, and ATX-II on neuronal sodium
currents. Our experiments show that the main effect exerted
by both BgII and BgIII is a concentration-dependent slowing
of the inactivation process of TTX-S sodium current, with no
significant effects either on activation kinetics or current
peak amplitude. No significant effects were observed on
TTX-R sodium currents.
In our experimental conditions, BgII was about 3 times
more potent than BgIII, a finding that is consistent with a
previous study in mice by Loret et al. (1994), showing that
BgII is more toxic than BgIII when injected intracerebroven-
tricularly. These authors showed that the higher toxicity of
BgII is correlated to a higher binding competition with the
-scorpion toxin AaH-II (from Androctonus australis Hector)
in rat brain synaptosomes despite their lack of sequence
homology. An interesting fact worth mentioning is that BgII
and BgIII amino acid sequences are almost identical, differ-
ing only by a single amino acid; in BgIII, at position 16, an
aspartic acid replaces the asparagine of BgII. BgII and BgIII
exhibit a higher similarity with type 1 sea anemone toxins
like ATX-II, ApA, or ApB (both from Anthopleura xantho-
grammica) than with type 2 toxins like ShI (from Stichodac-
tyla helianthus). The asparagine in position 16 is a very
conservative residue among type 1 toxins. Our results rein-
force the idea that this amino acid residue plays a central role
in the action of these toxins (Loret et al., 1994; Goudet et al.,
2001).
The electrophysiological effects of ATX-II and
-scorpion
toxins on macroscopic TTX-S sodium currents include the
inhibition of the Na
channel inactivation (Pelhate et al.,
1984; Neumcke et al., 1985), the presence of sodium currents
after prolonged depolarization (Warashina and Fujita, 1983),
and voltage-dependent action (Strichartz and Wang, 1986).
The effects of both BgII and BgIII are very similar. Never-
theless, only BgII showed a voltage-dependent action. As it
was described for ATX-II, it seems to show more affinity for
its binding site at hyperpolarized than at depolarized poten-
tials (El-Sherif et al., 1992). In contrast, BgIII action was not
voltage-dependent. According to Rogers et al. (1996), sea
anemone toxin binding is less voltage-dependent than
-scor-
pion toxins, suggesting that they have fewer binding contacts
outside of the IVS3-S4 loop, so it is subjected to less steric or
torsional distortion when the channel is depolarized. The
absence of voltage-dependent action observed for BgIII could
be explained by a modification of the electrostatic interac-
tions with the sodium channel due to the presence of an
additional negative charge at the toxin molecule. In the lit-
erature, there are reports indicating a voltage-dependent
action of ATX-II (Lawrence and Catterall, 1981; Strichartz
and Wang, 1986) and
-PMTX (pompilidotoxin, a site 3 neu-
rotoxin from wasp venom) (Sahara et al., 2000) and also a
voltage-independent action (Vincent et al., 1980; Isenberg
and Ravens, 1984). At the moment, there is not a clear
explanation for these discrepancies, but it is possible that
isoform or species-specific differences in gating among so-
dium channels could be playing a role in this respect. The fact
that BgII and BgIII actions were not use-dependent suggests
that these toxins show no preference for the open state of the
sodium channel.
BgIII caused a decrease in the voltage dependence of so-
dium channel inactivation (i.e., significantly increased the
slopes of steady-state inactivation). This result is in agree-
ment with other studies involving site 3 neurotoxins (Gal-
lagher and Bluementhal, 1994; Cahine et al., 1996; Chen et
al., 2000).
In contrast with the lack of effect of BgII and BgIII on the
TTX-R currents in DRG neurons, BgII produced a significant
slowing of the inactivation and an increase in the slope of the
inactivation curve in ventricular cardiomyocytes Na
cur-
rents that are TTX-R (Goudet et al., 2001). Moreover, accord-
ing to our experiments, in the presence of BgII or BgIII,
steady-state inactivation curves of DRG neurons showed a
shift to the left (i.e., to more hyperpolarized potentials) in the
voltage at which half of the channels are inactivated. This
Fig. 5. Effects of BgII and BgIII on steady-state sodium current inacti-
vation. A, typical records from which inactivation curves were obtained.
B, steady-state inactivation profiles under control conditions (n8) and
during perfusion of 10
M BgII (n4) or 10
M BgIII (n4). The
steady-state inactivation (h
) was determined using the two-pulse proto-
col shown in the inset. Data were plotted as a function of the prepulse
potential and fit by a Boltzmann function. A 10
M concentration of
either BgII or BgIII caused a significant hyperpolarizing shift in the
V
1/2 inact
, from 61 0.8 mV (control) to 73 0.9 mV (BgII) and 69
1.4 mV (BgIII).
1072 Salceda et al.
effect is shared by several site 3 toxins (Gordon et al., 1996).
BgII and BgIII applied on cloned hH1 sodium channels pro-
duced a depolarizing shift in the steady-state inactivation
curves but no shift at all when tested on rat ventricular
cardiomyocytes (Goudet et al., 2001). Neither BgII nor BgIII
showed a significant effect on sodium current recovery from
inactivation. These data are in agreement with the results
reported by Goudet et al. (2001) in rat ventricular cardiomy-
ocytes.
Several studies have revealed that sea anemone toxins
bind with higher affinity to cardiac sodium channels than to
neuronal ones (El-Sherif et al., 1992; Roden et al., 2002). The
discrepancies between the actions of BgII and BgIII on car-
diac and neuronal cells could be the result of tissue- or
species-specific differences. TTX-R currents in DRG neurons
are mainly due to type 1.8 and 1.9 Na
channels, whereas in
the heart, the 1.5 subunit is mainly expressed (Novakovik et
al., 2001; Dib-Hajj et al., 2002).
Site 3 sodium channel toxins have been suggested to slow
the open state to inactivated state transition rate
(Warashina and Fujita, 1983; Strichartz and Wang, 1986;
Schreibmayer et al., 1987; Kirsch et al., 1989), but this phe-
nomenon is difficult to study with macroscopic currents be-
cause current decay at a wide range of voltages represents a
combination of delayed channel openings and channel inac-
tivation (El-Sherif et al., 1992). Nevertheless, both the slow-
ing of inactivation and the reduction in voltage dependence of
steady-state inactivation observed in our experiments sug-
gest that the general explanation proposed by Rogers et al.
(1996) regarding the putative mechanism of action of site 3
toxins is also applicable to BgII and BgIII: 1) the toxin re-
ceptor site undergoes a conformational change that is re-
quired for fast inactivation; 2) bound toxin slows this confor-
mational change and, as a consequence, slows the
inactivation process; and 3) since site 3 neurotoxins bound
across the IVS3-S4 extracellular loop of the sodium channel
and that translocation of IVS4 segment may be required for
the inactivation gate to close, anemone toxins could be slow-
ing or blocking such translocation and thus hindering inac-
tivation. Slowing of the inactivation, however, would in-
crease current density, an effect that has not been observed
in our experiments; thus we cannot exclude that BgII and
BgIII may also produce some degree of channel occlusion.
The existence of several peptides from different species
that bind site 3 of the Na
channel is relevant from an
evolutionary point of view, indicating that this site is con-
served among species, thus constituting a target for poison-
ous toxins to act. Additionally, the analysis of peptide se-
quences of toxins acting on this site may allow the
identification of the elements of the sequence that are essen-
tial for binding to site 3 of the Na
channel. This could be of
particular relevance in the case of BgII and BgIII, toxins that
only differ by a single amino acid.
Acknowledgments
We are profoundly in debt to Professor Abel Aneiros for the kind
gift of BgII and BgIII toxins and with Professor L. Beress for kindly
supplying ATX-II. We are also grateful to M. Sa´nchez-Alvarez for
proofreading the English version.
References
Adams ME and Olivera BM (1994) Neurotoxins: overview of an emerging research
technology. Trends Neurosci 17:151155.
Alessandri-Haber N, Lecoq A, Gasparini S, Grangier-Macmath G, Jacquet G, Harvey
AL, de Medeiros C, Rowan EG, Gola M, Me´nez A, and Crest M (1999) Mapping the
functional anatomy of BgK on Kv1.1, Kv1.2 and Kv1.3. Clues to design analogs
with enhanced selectivity. J Biol Chem 274:3565335661.
Aneiros A, Garcı´a I, Martı´nez JR, Harvey AL, Anderson AJ, Marshall DL, Engstro¨m
Å, Hellman U, and Karlsson E (1993) A potassium channel toxin from the secretion
of the sea anemone Bunodosoma granulifera.Biochim Biophys Acta 1157:86–92.
Baker MD and Wood JN (2001) Involvement of Na
channels in pain pathways.
Trends Pharmacol Sci 22:27–31.
Bossu JL and Feltz A (1984) Patch-clamp study of the tetrodotoxin-resistant sodium
current in group C sensory neurons. Neurosci Lett 51:241–246.
Cahine M, Plante E, and Kallen RJ (1996) Sea anemone toxin ATX-II modulation of
heart and skeletal muscle sodium channel
-subunits expressed in tsA201 cells. J
Membr Biol 152:39–48.
Catterall WA (1995) Structure and function of voltage-gated ion channels. Annu Rev
Biochem 64:493–531.
Catterall WA (2000) From ionic currents to molecular mechanisms: the structure
and function of voltage-gated sodium channels. Neuron 26:13–25.
Couroud F, Rochat H, and Lissitzky S (1978) Binding of scorpion and sea anemone
toxins to a common site related to the action potential Na
ionophore in neuro-
blastoma cells. Biochem Biophys Res Commun 83:1525–1530.
Chen H, Gordon D, and Heinemann SH (2000) Modulation of cloned skeletal muscle
sodium channels by the scorpion toxins Lqh II, Lqh III and Lqh
IT. Pfluegers Arch
439:423–432.
Dauplais M, Lecoq A, Song J, Cotton J, Jamin N, Gilquin B, Roumestand CH, Vita
C, de Medeiros CLC, Rowan EG, et al. (1997) On the convergent evolution of
animal toxins. Conservation of a diad of functional residues in potassium channel-
blocking toxins with unrelated structures. J Biol Chem 272:4302–4309.
Dib-Hajj S, Black JA, Cummins TR, and Waxman SG (2002) NaN/Na v 1.9: a sodium
channel with unique properties. Trends Neurosci 25:253–259.
El-Sherif N, Fozzard HA, and Hanck DA (1992) Dose-dependent modulation of the
cardiac sodium channel by sea anemone toxin ATXII. Circ Res 70:285–301.
Elliot AA and Elliot JR (1993) Characterization of TTX-Sensitive and TTX-Resistant
sodium currents in small cells from adult rat dorsal root ganglia. J Physiol
(London)463:39–56.
Gallagher MJ and Bluementhal KM (1994) Importance of the unique cationic resi-
dues arginine 12 and lysine 49 in the activity of the cardiotonic polypeptide
anthopleurin B. J Biol Chem 269:254–259.
Garateix A, Vega R, Salceda E, Cebada J, Aneiros A, and Soto E (2000) BgK anemone
toxin inhibits outward K
currents in snail neurons. Brain Res 864:312–314.
Gordon D, Martin-Eauclaire MF, Caste` le S, Kopeyan C, Carlier E, Khalifa RB,
Pelhate M, and Rochat H (1996) Scorpion toxins affecting sodium current inacti-
vation bind to distinct homologous receptor sites on rat brain and insect sodium
channels. J Biol Chem 271:8034–8045.
Gordon D, Savarin Ph, Gurevitz M, and Zinn-Justin S (1998) Functional anatomy of
scorpion toxins affecting sodium channels. J Toxicol-Toxin Rev 17:131–159.
Goudet C, Ferrer T, Galan L, Artiles A, Batista CFV, Possani LD, Alvarez J, Aneiros
A, and Tytgat J (2001) Characterization of two Bunodosoma granulifera toxins
active on cardiac sodium channels. Br J Pharmacol 134:1195–1206.
Isenberg G and Ravens U (1984) The effects of Anemonia sulcata toxin (ATXII) on
membrane currents of isolated mammalian myocytes. J Physiol (London)357:127–
149.
Kirsch GE, Skattebol A, Possani LD, and Brown AM (1989) Modification of Na
channel gating by an alpha scorpion toxin from Tityus serrulatus.J Gen Physiol
93:67–83.
Lawrence JC and Catterall WA (1981) Tetrodotoxin-insensitive sodium channels.
Binding of polypeptide neurotoxins in primary cultures of rat muscle cells. Bio-
chim Biophys Acta 901:273–282.
Loret EP, Menendez Soto del Valle R, Mansuelle P, Sampieri F, and Rochat H (1994)
Positively charged amino acid residues located similarly in sea anemone and
scorpion toxins. J Biol Chem 269:16785–16788.
Narahashi T (1998) Chem modulation of sodium channels, in Ion Channel Pharma-
cology (Soria B and Cen˜ a V eds) pp 23–73, Oxford University Press, New York.
Neumcke B, Schwarz W, and Stampfli R (1985) Comparison of the Anemonia toxin
II on sodium and gating currents in frog myelinated nerve. Biochim Biophys Acta
814:111–119.
Norton RS (1997) Anemone toxins (type II), in Guidebook to Protein Toxins and Their
Use in Cell Biology (Rappuoli R and Montecucco C eds) pp 134–137, Oxford
University Press, New York.
Novakovik SD, Eglen RM, and Hunter JC (2001) Regulation of Na
channel distri-
bution in the nervous system. Trends Neurosci 24:473–478.
Pelhate M, Laufer J, Pichon Y, and Zlotkin E (1984) Effects of several sea anemone
and scorpion toxins on excitability and ionic currents in the giant axon of the
cockroach. J Physiol (Paris) 79:309–317.
Roden DM, Balser JR, George AL Jr, and Anderson ME (2002) Cardiac Ion Channels.
Annu Rev Physiol 64:431–475.
Rogers JC, Qu YS, Tanada TN, Scheuer T, and Catterall WA (1996) Molecular
determinants of high affinity binding of
-scorpion toxin and sea anemone toxin in
the S3–S4 extracellular loop in domain IV of the Na
channel subunit. J Biol
Chem 271:15950–15962.
Roy ML and Narahashi T (1992) Differential properties of tetrodotoxin-sensitive and
tetrodotoxin-resistant sodium channels in rat dorsal root ganglion neurones.
J Neurosci 12:2104–2111.
Sahara Y, Gotoh M, Konno K, Miwa A, Tsubokawa H, Robinson HPC, and Kawai N
(2000) A new class of neurotoxin from wasp venom slows inactivation of sodium
current. Eur J Neurosci 12:1961–1970.
Salinas EM, Cebada J, Valdes A, Garateix A, Aneiros A, and Alvarez JL (1997)
Effects of a toxin from the mucus of the Caribbean sea anemone (Bunodosoma
granulifera) on the ionic currents of single ventricular mammalian cardiomyo-
cytes. Toxicon 35:1699–1709.
BgII and BgIII Anemone Toxins 1073
Schreibmayer W, Kazerani H, and Tritthart H (1987) A mechanistic interpretation
of the action of toxin II from Anemonia sulcata on the cardiac sodium channel.
Biochim Biophys Acta 901:273282.
Strachan LC, Lewis RJ, and Nicholson GM (1999) Differential actions of Pacific
ciguatoxin-1 on sodium channel subtypes in mammalian sensory neurons. J Phar-
macol Exp Ther 288:379388.
Strichartz G, Rando T, and Wang GK (1987) An integrated view of the molecular
toxinology of sodium channel gating in excitable cells. Ann Rev Neurosci 10:237
267.
Strichartz GR and Wang GK (1986) Rapid voltage-dependent dissociation of scorpion
alpha-toxins coupled to Na channel inactivation in amphibian myelinated nerves.
J Gen Physiol 88:413435.
Trainer VL, Baden DG, and Catterall WA (1994) Identification of peptide compo-
nents of the brevotoxin receptor site of rat brain sodium channels. J Biol Chem
269:1990419909.
Vincent JP, Balerna M, Barhanin J, Fosset M, and Lazdunski M (1980) Differential
actions of Pacific ciguatoxin-1 on sodium channel subtypes in mammalian sensory
neurons. J Pharmacol Exp Ther 288:379388.
Warashina A and Fujita S (1983) Effects of sea anemone toxins on the sodium
inactivation process in crayfish axons. J Gen Physiol 81:305323.
Address correspondence to: Dr. Emilio Salceda, Instituto de Fisiologı´a,
Universidad Auto´noma de Puebla, Apartado Postal 406, Puebla, Pue., CP
72001, Me´xico. E-mail: esalceda@siu.buap.mx
1074 Salceda et al.
... All efforts were made to minimize animal suffering and to reduce the number of animals used. DRG neurons were isolated and maintained in primary culture according to the methodology described previously [41]. The dissection and cell culture were performed within a level I biosafety laminar flow hood (Nuaire, Plymouth, MN, USA). ...
... ASIC currents were activated by micro-perfusion of the cell under recording with an acid solution (pH 6.1) through a square tube using a rapid perfusion exchange system (SF-77B, Warner Inst., Hamden, CT, USA). The pH-gated current was activated using a pH of 6.1 which coincides with the pH 50 previously demonstrated for ASIC currents in DRG neurons [41]. Capsazepine 10 µM was added to the extracellular solution at pH 6.1 in order to limit the activation of TRPV1 receptors, which are also sensitive to acid and are expressed in DRG neurons [42]. ...
... Capsazepine 10 µM was added to the extracellular solution at pH 6.1 in order to limit the activation of TRPV1 receptors, which are also sensitive to acid and are expressed in DRG neurons [42]. To study the effect of Sa12b and Sh5b peptides on ASIC currents of DRG neurons two application protocols were used [41]. Peptides were applied by sustained application (preincubation) and by co-application. ...
Article
Full-text available
In this work, we evaluate the effect of two peptides Sa12b (EDVDHVFLRF) and Sh5b (DVDHVFLRF-NH2) on Acid-Sensing Ion Channels (ASIC). These peptides were purified from the venom of solitary wasps Sphex argentatus argentatus and Isodontia harmandi, respectively. Voltage clamp recordings of ASIC currents were performed in whole cell configuration in primary culture of dorsal root ganglion (DRG) neurons from (P7-P10) CII Long-Evans rats. The peptides were applied by preincubation for 25 s (20 s in pH 7.4 solution and 5 s in pH 6.1 solution) or by co-application (5 s in pH 6.1 solution). Sa12b inhibits ASIC current with an IC50 of 81 nM, in a concentration-dependent manner when preincubation application was used. While Sh5b did not show consistent results having both excitatory and inhibitory effects on the maximum ASIC currents, its complex effect suggests that it presents a selective action on some ASIC subunits. Despite the similarity in their sequences, the action of these peptides differs significantly. Sa12b is the first discovered wasp peptide with a significant ASIC inhibitory effect.
... Cell culture. The DRG neurons were isolated from the vertebral column of Wistar rats at postnatal ages P5 to P9 without sex distinction, and cultured according to the procedure described by Salceda and coworkers [71]. In brief, the neurons were incubated (30 min at 37˚C) in Leibovitz L15 medium (Invitrogen, USA) containing 1.25 mg/ml trypsin and 1.25 mg/ml collagenase (both from Sigma-Aldrich). ...
Article
Full-text available
We previously reported a multigene family of monodomain Kunitz proteins from Echinococcus granulosus (EgKU-1-EgKU-8), and provided evidence that some EgKUs are secreted by larval worms to the host interface. In addition, functional studies and homology modeling suggested that, similar to monodomain Kunitz families present in animal venoms, the E. granulosus family could include peptidase inhibitors as well as channel blockers. Using enzyme kinetics and whole-cell patch-clamp, we now demonstrate that the EgKUs are indeed functionally diverse. In fact, most of them behaved as high affinity inhibitors of either chymotrypsin (EgKU-2-EgKU-3) or trypsin (EgKU-5-EgKU-8). In contrast, the close paralogs EgKU-1 and EgKU-4 blocked voltage-dependent potassium channels (Kv); and also pH-dependent sodium channels (ASICs), while showing null (EgKU-1) or marginal (EgKU-4) peptidase inhibitory activity. We also confirmed the presence of EgKUs in secretions from other parasite stages, notably from adult worms and metacestodes. Interestingly, data from genome projects reveal that at least eight additional monodomain Kunitz proteins are encoded in the genome; that particular EgKUs are up-regulated in various stages; and that analogous Kunitz families exist in other medically important cestodes, but not in trematodes. Members of this expanded family of secreted cestode proteins thus have the potential to block, through high affinity interactions, the function of host counterparts (either peptidases or cation channels) and contribute to the establishment and persistence of infection. From a more general perspective, our results confirm that multigene families of Kunitz inhibitors from parasite secretions and animal venoms display a similar functional diversity and thus, that host-parasite co-evolution may also drive the emergence of a new function associated with the Kunitz scaffold.
... En particular, las toxinas con acción sobre canales iónicos, han tenido importante contribución como reactivos biológicos en la dilucidación de los mecanismos moleculares vinculados con el funcionamiento del sistema cardiovascular y nervioso. A partir de anémonas marinas han sido aisladas nuevas estructuras moleculares de naturaleza proteica con acción sobre canales iónicos presentes en células nerviosas y cardíacas (Aneiros et al., 1993, Castañeda et al., 1995, Lanio et al., 2001Goudet et al., 2001;Garateix et al., 1992, 1996, 2000, Salceda et al., 2002, Álvarez et al., 2003, Oliveira et al., 2006. Entre los principales resultados están el descubrimiento de la toxina κπ-PMTX-PCr1a en la anémona Phymanthus crucifer, reconocido como el primer inhibidor de canales activados por protones (por su sigla en inglés ASICs). ...
... It has been shown that application of veratridine, which prolongs the open state of the sodium channel, increases K Na -channel activity (Hage and Salkoff 2012). The use of this and other toxins that prolong Na -channel inactivation such as BgII and BgIII (the toxins isolated from the sea anemone Bunodosoma granulifera; Bosmans et al. 2002; Salceda et al. 2002) and CgNa (the toxin isolated from the sea anemone Condylactis gigantea; Billen et al. 2010; Ständker et al. 2006) would further support the functional coupling between K Na channels and voltage-dependent persistent Na current in cricket Kenyon cells. Hage and Salkoff (2012) suggested that I Naf is not responsible for the activation of K Na channels because I KNa persists even when I Naf is eliminated or reduced by a depolarizing holding potential of 50 mV in tufted/mitral cells. ...
... It has been shown that application of veratridine, which prolongs the open state of the sodium channel, increases K Na -channel activity (Hage and Salkoff 2012). The use of this and other toxins that prolong Na -channel inactivation such as BgII and BgIII (the toxins isolated from the sea anemone Bunodosoma granulifera; Bosmans et al. 2002; Salceda et al. 2002) and CgNa (the toxin isolated from the sea anemone Condylactis gigantea; Billen et al. 2010; Ständker et al. 2006) would further support the functional coupling between K Na channels and voltage-dependent persistent Na current in cricket Kenyon cells. Hage and Salkoff (2012) suggested that I Naf is not responsible for the activation of K Na channels because I KNa persists even when I Naf is eliminated or reduced by a depolarizing holding potential of 50 mV in tufted/mitral cells. ...
... It has been shown that application of veratridine, which prolongs the open state of the sodium channel, increases K Na -channel activity (Hage and Salkoff 2012). The use of this and other toxins that prolong Na -channel inactivation such as BgII and BgIII (the toxins isolated from the sea anemone Bunodosoma granulifera; Bosmans et al. 2002; Salceda et al. 2002) and CgNa (the toxin isolated from the sea anemone Condylactis gigantea; Billen et al. 2010; Ständker et al. 2006) would further support the functional coupling between K Na channels and voltage-dependent persistent Na current in cricket Kenyon cells. Hage and Salkoff (2012) suggested that I Naf is not responsible for the activation of K Na channels because I KNa persists even when I Naf is eliminated or reduced by a depolarizing holding potential of 50 mV in tufted/mitral cells. ...
... It has been shown that application of veratridine, which prolongs the open state of the sodium channel, increases K Na -channel activity (Hage and Salkoff 2012). The use of this and other toxins that prolong Na -channel inactivation such as BgII and BgIII (the toxins isolated from the sea anemone Bunodosoma granulifera; Bosmans et al. 2002; Salceda et al. 2002) and CgNa (the toxin isolated from the sea anemone Condylactis gigantea; Billen et al. 2010; Ständker et al. 2006) would further support the functional coupling between K Na channels and voltage-dependent persistent Na current in cricket Kenyon cells. Hage and Salkoff (2012) suggested that I Naf is not responsible for the activation of K Na channels because I KNa persists even when I Naf is eliminated or reduced by a depolarizing holding potential of 50 mV in tufted/mitral cells. ...
... It has been shown that application of veratridine, which prolongs the open state of the sodium channel, increases K Na -channel activity (Hage and Salkoff 2012). The use of this and other toxins that prolong Na ϩ -channel inactivation such as BgII and BgIII (the toxins isolated from the sea anemone Bunodosoma granulifera; Bosmans et al. 2002;Salceda et al. 2002) and CgNa (the toxin isolated from the sea anemone Condylactis gigantea; Billen et al. 2010;Ständker et al. 2006) would further support the functional coupling between K Na channels and voltage-dependent persistent Na ϩ current in cricket Kenyon cells. Hage and Salkoff (2012) suggested that I Naf is not responsible for the activation of K Na channels because I KNa persists even when I Naf is eliminated or reduced by a depolarizing holding potential of Ϫ50 mV in tufted/mitral cells. ...
Article
Full-text available
In this study, we examined the functional coupling between Na(+)-activated potassium (KNa) channels and Na(+)-influx through voltage-dependent Na(+) channels in Kenyon cells isolated from the mushroom body of the cricket. Single channel activity of KNa channels was recorded with the cell-attached patch configuration. The open probability (Po) of KNa channels increased with increasing Na(+) concentration in a bath solution, whereas it decreased by the substitution of Na(+) with an equimolar concentration of Li(+). The PO of K(Na) channels was also found to be reduced by bath application of a high concentration of tetrodotoxin (TTX; 1 µM), which inhibits both fast (INaf) and persistent Na(+) (INaP) currents, whereas it was unaffected by a low concentration of TTX (10 nM), which selectively blocks INaf. Bath application of Cd(2+) at a low concentration (50 µM), as an inhibitor of INaP, also decreased the Po of KNa channels. Conversely, bath application of the inorganic Ca(2+) channel blockers Co(2+) and Ni(2+) at high conentrations (500 µM) had little effect on the Po of KNa channels, although Cd(2+) (500 µM) reduced the Po of KNa channels. Perforated whole-cell clamp analysis further indicated the presence of sustained outward currents whose amplitude was dependent on the amount of Na(+) influx. Taken together, these results indicate that KNa channels could be activated by Na(+) influx passing through voltage-dependent persistent Na(+) channels. The functional significance of this coupling mechanism was discussed in relation to the membrane excitability of Kenyon cells and its possible role in the formation of long-term memory. Copyright © 2015, Journal of Neurophysiology.
Article
There are two kinds of toxins in sea anemones: neurotoxins and pore forming toxins. As a representative of the sodium channel toxin, the neurotoxin ATX II in neurotoxin mainly affects the process of action potential and the release of transmitter to affect the inactivation of the sodium channel. As the representatives of potassium channel toxins, BgK and ShK mainly affect the potassium channel current. EqTx and Sticholysins are representative of pore forming toxins, which can form specific ion channels in cell membranes and change the concentration of internal and external ions, eventually causing hemolytic effects. Based on the above mechanism, toxins such as ATX II can also cause toxic effects in tissues and organs such as heart, lung and muscle. As an applied aspect it was shown that sea anemone toxins often have strong toxic effects on tumor cells, induce cancer cells to enter the pathway of apoptosis, and can also bind to monoclonal antibodies or directly inhibit relevant channels for the treatment of autoimmune diseases.
Article
The marine environment is a vast resource of natural products of unique characteristics. Among these are the ones that are peptidic in nature, more specifically, compounds that are ribosomally produced as larger proteins and are subsequently posttranslationally modified to produce the desired peptidic marine natural product that possesses small-molecule characteristics. Venomous predatory marine animals such as cone snails and sea anemones produce a plethora of peptidic scaffolds that target specific ion channels and receptors as part of their neurochemical and biochemical strategy to capture their prey. In this chapter, we discuss in detail the chemistry and biology of the peptidic components produced by these animals. Components found in the venom of cone snails are typically smaller and heavily modified when compared to sea anemone toxins. Nevertheless, both cone snails and sea anemones produce a myriad of peptides of which a large group acts on voltage-gated ion channels. In comparison to other organisms, their venoms and toxins are not well studied; however, the first compound of marine origin to receive official approval as a drug in the United States is Prialt, a 25-residue conotoxin from the venom of Conus magus, which targets selectively Cav2.2, a voltage-gated calcium channel implicated in certain types of chronic pain. The pharmacological applications of peptidic marine natural products and their use as neuronal probes are also discussed.
Article
Full-text available
TTX-sensitive (TTX-S) and TTX-resistant (TTX-R) sodium channel currents were analyzed in acutely dissociated dorsal root ganglion (DRG) neurons isolated from 3-12-d-old and adult rats. Currents were recorded using the whole-cell patch-clamp technique. TTX-R current was more likely to be present in younger animals (3-7 d), whereas TTX-S current was more common in older animals (7-10 d), although TTX-R current was recorded from adult rat DRG neurons. The TTX-R and TTX-S currents differed in their steady-state inactivation, with 50% inactivation voltage at -40 +/- 5 mV (n = 10) for TTX-R currents and -70 +/- 4 mV (n = 10) for TTX-S currents. These current types also differed in their activation kinetics, with 50% activation values of -15 +/- 5 mV (n = 5) for TTX-R currents and -26 +/- 6 mV (n = 5) for TTX-S currents. The interactions of TTX-R and TTX-S channels with various pharmacological agents and divalent cations were studied. The Kd values for TTX-S and TTX-R currents were estimated to be 0.3 nM and 100 microM for TTX, 0.5 nM and 10 microM for saxitoxin, and 50 microM and 200 microM for lidocaine, respectively. TTX-S channels did not exhibit a marked use-dependent block by lidocaine, whereas lidocaine significantly decreased TTX-R current in a use-dependent manner at frequencies ranging from 1 to 33.3 Hz. Several external divalent cations exerted different effects on these current types.(ABSTRACT TRUNCATED AT 250 WORDS)
Article
Full-text available
Sodium channels posses receptor sites for many neurotoxins, of which several groups were shown to inhibit sodium current inactivation. Receptor sites that bind α- and α-like scorpion toxins are of particular interest since neurotoxin binding at these extracellular regions can affect the inactivation process at intramembranal segments of the channel. We examined, for the first time, the interaction of different scorpion neurotoxins, all affecting sodium current inactivation and toxic to mammals, with α-scorpion toxin receptor sites on both mammalian and insect sodium channels. As specific probes for rat and insect sodium channels, we used the radiolabeled α-scorpion toxins AaH II and LqhαIT, the most active α-toxins on mammals and insect, respectively. We demonstrate that the different scorpion toxins may be classified to several groups, according to their in vivo and in vitro activity on mammalian and insect sodium channels. Analysis of competitive binding interaction reveal that each group may occupy a distinct receptor site on sodium channels. The α-mammal scorpion toxins and the anti-insect LqhαIT bind to homologous but not identical receptor sites on both rat brain and insect sodium channels. Sea anemone toxin ATX II, previously considered to share receptor site 3 with α-scorpion toxins, is suggested to bind to a partially overlapping receptor site with both AaH II and LqhαIT. Competitive binding interactions with other scorpion toxins suggest the presence of a putative additional receptor site on sodium channels, which may bind a unique group of these scorpion toxins (Bom III and IV), active on both mammals and insects. We suggest the presence of a cluster of receptor sites for scorpion toxins that inhibit sodium current inactivation, which is very similar on insect and rat brain sodium channels, in spite of the structural and pharmacological differences between them. The sea anemone toxin ATX II is also suggested to bind within this cluster.
Article
Full-text available
Specific groups of sea anemone and scorpion toxins compete on the same pharmacological site, on the voltage-gated sodium channel of mammal excitable membranes. However, these scorpion and sea anemone toxins are two distinct protein families. Here we purified and sequenced a new sea anemone toxin, Bg II, highly toxic to mammals and also a less toxic mutant, Bg III. Two Bg II models were determined from sequence homologies with two sea anemone toxin two-dimensional NMR structures. Only one model conformed to circular dichroism data obtained from Bg II and was compared with an x-ray structure of a scorpion toxin. The comparison of the two structures shows that 5 amino acid residues are located similarly in the sea anemone toxin and the scorpion toxin. From these 5 residues, 4 are basic residues, constituting two distinct positively charged poles on the surface of these toxins. In the sea anemone mutant isolated, a negative charge beside one of the positive poles decreases the toxicity. These results show that positively charged amino acid residues could be essential for the activity of these toxins and outline the role of electrostatic bonds in the interaction of sea anemone and scorpion toxins with their receptor.
Article
Full-text available
Long chain scorpion toxins (made of 60 to 70 amino acids) acting on voltage-gated sodium channels in excitable cells are responsible for human envenomation, and comprise a-toxins that inhibit sodium current inactivation and p-toxins, that modify the activation process. These toxins may be further divided according to their pharmacological activities. Thus, a-toxins highly active on mammals or insects, as well as α like toxins may be distinguished within the α-toxin class. The β-toxin class includes toxins active on mammals and, as a separate group, the excitatory and depressant toxins active exclusively on insects. All these toxins possess 4 disulfide bridges and share 15 similar non cystine residues. Accordingly, their 3D structure is highly conserved, comprising an α-helix and a triple stranded β-sheet. The most solvent exposed turns of this structure are prone to insertions or deletions, and accordingly correspond to the most structurally variable regions of the toxins. They have been tested by chemical modification (on several toxins) and site-directed mutagenesis analysis (of LqhαIT) for their possible involvement in the interactions with sodium channels.
Article
Full-text available
We studied the effects of BgK toxin on outward K+ currents in isolated neurons of the snail Helix aspersa, using the whole cell patch clamp technique. BgK partially and reversibly blocked K+ currents in the 1 pM to 100 nM concentration range (n=53). The dose–response curve for BgK current inhibition had a maximum blocking effect at 100 nM. Our results indicate that BgK is a potent, apparently non-selective, K+ channel blocker in molluscan neurons.
Chapter
Many prokaryotes, animals, and plants produce toxins that exert their toxic, sometimes lethal, activity toward other living organisms. In several cases, the same organisms produce more than one toxin at the same time. Toxins can be considered a sort of biological warfare that somehow ‘helps’ the toxin-producing species in their struggle for life. Toxins are believed to increase the chance of survival and/or proliferation and/or spreading of a particular organism in its particular environment. While it is simple to understand the advantages conferred by certain toxins contained in the snake or spider venoms, this is not always the case for plants or microorganisms. However, this is to be attributed to our present lack of knowledge of the ecology and evolutionary history of the toxin-producing species, rather than to a lack of ‘significance’ or ‘usefulness’ of the toxin.
Article
The normal electrophysiologic behavior of the heart is determined by ordered propagation of excitatory stimuli that result in rapid depolarization and slow repolarization, thereby generating action potentials in individual myocytes. Abnormalities of impulse generation, propagation, or the duration and configuration of individual cardiac action potentials form the basis of disorders of cardiac rhythm, a continuing major public health problem for which available drugs are incompletetly effective and often dangerous. The integrated activity of specific ionic currents generates action potentials, and the genes whose expression results in the molecular components underlying individual ion currents in heart have been cloned. This review discusses these new tools and how their application to the problem of arrhythmias is generating new mechanistic insights to identify patients at risk for this condition and developing improved antiarrhythmic therapies.
Article
The effects of α-pompilidotoxin (α-PMTX), a new neurotoxin isolated from the venom of a solitary wasp, were studied on the neuromuscular synapses in lobster walking leg and the rat trigeminal ganglion (TG) neurons. Paired intracellular recordings from the presynaptic axon terminals and the innervating lobster leg muscles revealed that α-PMTX induced long bursts of action potentials in the presynaptic axon, which resulted in facilitated excitatory and inhibitory synaptic transmission. The action of α-PMTX was distinct from that of other known facilitatory presynaptic toxins, including sea anemone toxins and α-scorpion toxins, which modify the fast inactivation of Na+ current. We further characterized the action of α-PMTX on Na+ channels by whole-cell recordings from rat trigeminal neurons. We found that α-PMTX slowed the Na+ channels inactivation process without changing the peak current–voltage relationship or the activation time course of tetrodotoxin (TTX)-sensitive Na+ currents, and that α-PMTX had voltage-dependent effects on the rate of recovery from Na+ current inactivation and deactivating tail currents. The results suggest that α-PMTX slows or blocks conformational changes required for fast inactivation of the Na+ channels on the extracellular surface. The simple structure of α-PMTX, consisting of 13 amino acids, would be advantageous for understanding the functional architecture of Na+ channel protein.
Article
The protein neurotoxin II from the venom of the scorpion Hector was labeled with 125I by the lactoperoxidase method to a specific radioactivity of about 100 μCi/μg without loss of biological activity. The labeled neurotoxin binds specifically to a single class of non intereacting binding sites of high affinity (KD = 0.3 – 0.6 nM) and low capacity (4000 – 8000 sites/cell) to electrically excitable neuroblastoma cells. Relation of these sites to the action potential Na+ channel is derived from identical concentration dependence of scorpion toxin binding and increase in duration and amplitude of action potential. The protein neurotoxin II from the sea anemone also affects the closing of the action potential Na+ ionophore in nerve axons. The unlabelled sea anemone toxin modifies 125I-labeled scorpion toxin II binding to neuroblastoma cells by increasing the apparent KD for labeled scorpion toxin without modification of the number of binding sites. It is concluded that both scorpion toxin II and sea anemone toxin II interact competitively with a regulatory component of the action potential Na+ channel.