ArticlePDF Available

Analyses of a novel SCN5A mutation (C1850S): Conduction vs. repolarization disorder hypotheses in the Brugada syndrome

Authors:

Abstract and Figures

Brugada syndrome (BrS) is characterized by arrhythmias leading to sudden cardiac death. BrS is caused, in part, by mutations in the SCN5A gene, which encodes the sodium channel alpha-subunit Na(v)1.5. Here, we aimed to characterize the biophysical properties and consequences of a novel BrS SCN5A mutation. SCN5A was screened for mutations in a male patient with type-1 BrS pattern ECG. Wild-type (WT) and mutant Na(v)1.5 channels were expressed in HEK293 cells. Sodium currents (I(Na)) were analysed using the whole-cell patch-clamp technique at 37 degrees C. The electrophysiological effects of the mutation were simulated using the Luo-Rudy model, into which the transient outward current (I(to)) was incorporated. A new mutation (C1850S) was identified in the Na(v)1.5 C-terminal domain. In HEK293 cells, mutant I(Na) density was decreased by 62% at -20 mV. Inactivation of mutant I(Na) was accelerated in a voltage-dependent manner and the steady-state inactivation curve was shifted by 11.6 mV towards negative potentials. No change was observed regarding activation characteristics. Altogether, these biophysical alterations decreased the availability of I(Na). In the simulations, the I(to) density necessary to precipitate repolarization differed minimally between the two genotypes. In contrast, the mutation greatly affected conduction across a structural heterogeneity and precipitated conduction block. Our data confirm that mutations of the C-terminal domain of Na(v)1.5 alter the inactivation of the channel and support the notion that conduction alterations may play a significant role in the pathogenesis of BrS.
Content may be subject to copyright.
Analyses of a novel SCN5A mutation (C1850S):
conduction vs. repolarization disorder hypotheses
in the Brugada syndrome
Se
´
verine Petitprez
1
, Thomas Jespersen
1
, Etienne Pruvot
2
, Dagmar I. Keller
3
, Cora Corbaz
2
,
Ju
¨
rg Schla
¨
pfer
2
, Hugues Abriel
1,2
*
, and Jan P. Kucera
4
*
1
Department of Pharmacology and Toxicology, University of Lausanne, 27, Bugnon, 1005 Lausanne, Vaud, Switzerland;
2
Service of Cardiology, CHUV, Lausanne, Switzerland;
3
Department of Cardiology, University Hospital, Basel, Switzerland;
and
4
Department of Physiology, University of Bern, Bu
¨
hlplatz 5, 3012 Bern, Switzerland
Received 28 March 2007; revised 28 January 2008; accepted 31 January 2008; online publish-ahead-of-print 5 February 2008
Time for primary review: 29 days
Aims Brugada syndrome (BrS) is characterized by arrhythmias leading to sudden cardiac death. BrS is
caused, in part, by mutations in the SCN5A gene, which encodes the sodium channel alpha-subunit
Na
v
1.5. Here, we aimed to characterize the biophysical properties and consequences of a novel BrS
SCN5A mutation.
Methods and results SCN5A was screened for mutations in a male patient with type-1 BrS pattern ECG.
Wild-type (WT) and mutant Na
v
1.5 channels were expressed in HEK293 cells. Sodium currents (I
Na
) were
analysed using the whole-cell patch-clamp technique at 378C. The electrophysiological effects of the
mutation were simulated using the Luo-Rudy model, into which the transient outward current (I
to
)
was incorporated. A new mutation (C1850S) was identified in the Na
v
1.5 C-terminal domain. In
HEK293 cells, mutant I
Na
density was decreased by 62% at 220 mV. Inactivation of mutant I
Na
was accel-
erated in a voltage-dependent manner and the steady-state inactivation curve was shifted by 11.6 mV
towards negative potentials. No change was observed regarding activation characteristics. Altogether,
these biophysical alterations decreased the availability of I
Na
. In the simulations, the I
to
density necess-
ary to precipitate repolarization differed minimally between the two genotypes. In contrast, the
mutation greatly affected conduction across a structural heterogeneity and precipitated conduction
block.
Conclusion Our data confirm that mutations of the C-terminal domain of Na
v
1.5 alter the inactivation of
the channel and support the notion that conduction alterations may play a significant role in the patho-
genesis of BrS.
KEYWORDS
Brugada syndrome;
Sodium channel;
Genetics;
Electrophysiology;
Computational analysis
1. Introduction
Brugada syndrome (BrS) is a disorder characterized by ST
segment elevation in the right precordial leads V1
V3 on
the surface ECG, with atypical right bundle branch block
pattern, and an increased risk for sudden cardiac death
(SCD).
1
The BrS is a primary electrical cardiac disorder,
which is inherited as an autosomal dominant trait. Many
mutations in SCN5A, encoding the a-subunit of the voltage-
gated sodium channel Na
v
1.5, have been identified,
2
though
only 10 to 30% of clinically affected individuals carry a
mutation in this gene.
3
Among the 90 mutations reported
thus far,
2
only a few of them have been characterized at the
molecular and biophysical level using cellular expression
systems.
1
Most of SCN5A mutations lead to a ‘loss-
of-function’ by reducing the sodium current (I
Na
) available
during the phases 0 (upstroke) and 1 (early repolarization)
of the cardiac action potential (AP).
1
Despite more than 10
years of research in this field, the molecular and cellular
mechanisms leading to the BrS are not yet completely
understood.
1
Interestingly, in some patients, BrS is con-
cealed on the surface ECG and can be unmasked using
sodium channel blockers
4
or during fever episodes.
5
Here, we screened SCN5A in a patient with typical clinical
manifestations of BrS. A novel mutation, C1850S, located
in the C-terminal domain of Na
v
1.5 was identified.
*
Corresponding author. Tel: þ41 21 692 5364; fax: þ41 21 692 5355 (H.A);
Tel: þ41 31 631 87 59; fax: þ41 31 631 4611 (J.P.K.).
E-mail addresses: hugues.abri el@unil.ch (H.A.); kucera@pyl.unibe.ch
(J.P.K.)
These authors contributed equally to this study.
These authors share the last authorship of this article.
Published on behalf of the European Society of Cardiology. All rights reserved. & The Author 2008.
For permissions please email: journals.permissions@oxfordjournals.org.
Cardiovascular Research (2008) 78, 494
504
doi:10.1093/cvr/cvn023
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
When mutant channels were expressed in HEK293 cells, the
peak I
Na
density was decreased by 62% (at 220 mV) and fast
I
Na
inactivation was accelerated.
Computer simulations were conducted using the Luo-Rudy
(LRd) model to examine the interaction of I
Na
and the tran-
sient outward current (I
to
), proposed to determine early
repolarization in the right epicardium.
6
The I
to
density
necessary to precipitate repolarization was similar in the
presence of wild-type (WT) and mutant I
Na
; however, our
computational analyses suggest that conduction in structu-
rally discontinuous tissue is prone to block and much more
sensitive to I
to
in the presence of the mutation at the het-
erozygous state. Moreover, these manifestations are modu-
lated by the extracellular concentration of K
þ
. Our data
confirm that mutations of the C-terminal domain of Na
v
1.5
alter the inactivation properties of the channel and
support the notion that conduction disturbances may be
involved in the pathogenesis of BrS.
2. Methods
2.1 Molecular screening
Genomic DNA was extracted from peripheral lymphocytes, and all
coding exons of SCN5A were amplified by polymerase chain reaction,
using primers designed in intronic flanking sequences according to
the gene sequences.
7
Denaturating high performance liquid chrom-
atography (DHPLC) was performed on DNA-amplification products on
at least one temperature condition. Abnormal DHPLC profiles were
analysed by sequence reaction in both strands of the exon, using
Big Dye termination mix, and analysed by cycle sequencing on an
automated laser fluorescent DNA sequencer (ABI prism 3100,
Applied Biosystems). The novel mutation C1850S was absent in
200 normal alleles. The investigation conformed with the principles
outlined in the Declaration of Helsinki. The index patient gave
written informed consent to participate in the study.
2.2 Electrophysiology
SCN5A mutation was engineered into WT cDNA (clone hH1a received
from Dr R. Kass) cloned in pcDNA3.1 (Invitrogen), using the Quick-
Change Kit (Stratagene) and verified by sequencing. For electro-
physiology studies, HEK293 cells were transiently transfected with
0.6 mg of WT or mutant Na
v
1.5 construct cDNAs or 0.3 mg of each.
All transfections included 2.0 mg pIRES-hb1-CD8 cDNA encoding
hb1 subunit and CD8 antigen as a reporter gene. Cells were trans-
fected using calcium phosphate or Lipofectamine
w
and I
Na
were
measured after 48 h. Anti-CD8 beads (Dynal) were used to identify
transfected cells. Patch-clamp recordings were conducted using
an internal solution containing (mmol/L) CsCl 60; CsAspartate 70;
EGTA 11; MgCl
2
1; CaCl
2
1; HEPES 10; and Na
2
-ATP 5, pH 7.2 with
CsOH; external solution NaCl 130; CaCl
2
2; MgCl
2
1.2; CsCl 5;
HEPES 10; and glucose 5, pH 7.4 with CsOH. Using these solutions,
5 min after rupturing the membrane, we observed no significant
alteration of the availability curve and the peak current. Peak cur-
rents were measured during an inactivation protocol and I
Na
den-
sities (pA/pF) were obtained by dividing the peak I
Na
by the cell
capacitance obtained from the pClamp function (pClamp suite,
v.8, Axon Instruments, CA, USA). For the activation and steady-state
inactivation curves, data from individual cells were fitted with
Boltzmann relationship, y(V
m
) = 1/{1 þ exp [(V
m
2V
1/2
)/K]}, in
which y is the normalized current or conductance, V
m
is the mem-
brane potential, V
1/2
is the voltage at which half of the channels
are activated or inactivated, and K is the slope factor. Recovery
from inactivation curves were fitted individually with a mono-
exponential relationship. I
Na
were measured using a VE-2 amplifier
(Alembic Instruments, Montre
´
al, QC) allowing a full compensation
of the series resistance, and were analysed using the pClamp soft-
ware suite. No correction of liquid junction potentials was per-
formed. All measurements were carried out at 37 + 18C using a
control system (Cell MicroControls) heating the perfusion solution.
2.3 Biochemistry
Western blotting conditions have been described previously.
8
The anti-Na
v
1.5 affinity purified antibody recognizes residues 493
511 of rat Na
v
1.5 (ASC-005, Alomone). These residues are
identical in the human sequence. The anti-actin antibody was
from Sigma.
2.4 Statistics
Data are presented as means + SEM. The two-tailed Student t-test
or the ANOVA test with the Bonferroni correction were used to
compare means; P , 0.05 was considered significant.
2.5 Mathematical modelling of I
Na
, the action
potential, and conduction
The LRd model of the ventricular epicardial cell
9
was used to recon-
struct I
Na
under voltage clamp conditions as well as the AP of the
epicardial myocyte. The model incorporated the updates published
by Faber and Rudy
10
as well as the ATP-sensitive K
þ
current as
described by Shaw and Rudy.
11
The transient outward K
þ
current
I
to
was introduced into the model according to the formulation of
Dumaine et al.
12
To account for the important heterogeneity of I
to
expression between the left and right ventricles and across the myo-
cardial wall, simulations were run over a broad range of I
to
maximal
conductance (g
to
, from 0 to 4 mS/mF).
In the LRd model, I
Na
is formulated as I
Na
= g
Na
.
m
3
hj(V
m
2E
Na
),
where g
Na
is the maximal conductance of I
Na
(16 mS/mF), m, h,
and j are the activation, fast inactivation, and slow inactivation
gating variables, respectively, and E
Na
is the Na
þ
Nernst potential.
The effects of the C1850S mutation (CS) were simulated by
modifying
b
h
, (the closing rate of the fast inactivation gate h)ina
voltage-dependent manner consistent with our experimental
findings to account for (i) the decrease of peak I
Na
by 62% at V
m
= 220 mV; (ii) the shift of the V
1/2
of steady-state inactivation
by 11.6 mV towards more negative potentials; and (iii) the voltage-
dependent decrease of the time constant of fast I
Na
inactivation for
V
m
240 mV.
The WT and CS formulations of
b
h
are as follows:
WT (Original formulation of the LRd model):
ForV
m
, 40mV :
b
h;WT
¼ 3:56 expð0:079V
m
Þþ3:1 10
5
expð0:35V
m
Þ:
For V
m
40 mV :
b
h;WT
¼ 1=ð0:13 ð1 þ expððV
m
þ 10:66Þ=ð11:1ÞÞÞÞ:
CS:
For V
m
, 40 mV :
b
h;CS
¼ 2:8606 ðexpð0:0672 ðV
m
þ 40ÞÞÞ:
For V
m
40 mV :
b
h;CS
¼
b
h;WT
ð1:5 þ 4:1 expððV
m
þ 40Þ=16:8ÞÞ:
The opening rate constant
a
h
of gate h as well as the rates of the
m and j gates were not modified. To simulate heterozygote (HZ)
cells (or I
Na
in cells transfected with equal amounts of WT
and mutant cDNA), g
Na
was split into two equal components of
8 mS/mF each corresponding to the two SCN5A alleles (I
Na
=
I
Na,WT
þI
Na,CS
).
Further details concerning the modelling approach are provided in
the Supplementary material online.
Conduction vs. repolarization disorder in Brugada syndrome 495
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
3. Results
3.1 Index case
The index case was a 55-year old Caucasian male seen at our
cardiology clinic (Lausanne, Switzerland) for investigation of
a typical type-1 BrS ECG pattern recorded during cystitis
with high fever. However, the ECG remained abnormal
after fever (Figure 1A). The patient never had syncope or
palpitations but his mother died suddenly at age 35. The
echocardiography was normal. An electrophysiological
study was conducted using two quadripolar leads. Premature
ventricular stimulation was performed at cycle length of 600
and 400 ms followed by one and then two premature beats.
Ventricular fibrillation was reproducibly inducible at a drive
train of 400 ms followed by two premature beats at a coup-
ling interval of 240 and 200 ms, respectively. Sinus rhythm
was restored by an external biphasic shock of 120 J. No
ajmaline test was performed. An internal cardioverter-
defibrillator was implanted.
3.2 Identification and character ization of the
SCN5A mutation
A heterozygous G to C substitution at position 5549 of SCN5A
resulted in a C1850S mutation in the C-terminal tail of
Na
v
1.5 (Figure 1B). In HEK293 cells transiently transfected
with WT and mutant cDNAs, Western blot analyses showed
that the level of protein expression of C1850S mutant chan-
nels was comparable with that of WT (Figure 1C). In con-
trast, using the patch-clamp technique in the whole-cell
configuration, at 378C, C1850S channels generated I
Na
with
biophysical properties that were clearly different compared
with WT currents. First, the peak I
Na
density mediated by
C1850S channels was reduced by 62 + 5% at 220 mV
(Figure 1D). When cells were co-transfected with equal
amounts of WT and mutant cDNAs to mimic the heterozygous
status of the patient, I
Na
density was intermediate (Figure
1D). Furthermore, as illustrated in Figure 2A, C1850S cur-
rents were almost completely inactivated 2 ms after the
voltage step (arrows), while WT currents were still substan-
tial. This indicates that C1850S I
Na
inactivated faster com-
pared with WT I
Na
. Figure 2B shows simulated current
traces generated by the original I
Na
formulation of the LRd
model (WT) and by the formulation in which the rate of
fast inactivation
b
h
was modified. Acceleration of fast inac-
tivation without modifying other gating kinetics and without
modifying I
Na
conductance resulted in a reduction of peak I
Na
by 63% (at 220 mV) and almost complete inactivation after
2 ms, similar to the experimental findings.
In HEK293 cells, the V
1/2
of voltage-dependence of
steady-state inactivation of C1850S was shifted towards
more negative values by 11.6 mV (Table 1 and Figure 3A).
No significant alteration in the activation curve V
1/2
was
observed. When HEK293 cells were transfected with equal
amount of WT and mutant cDNA, the shift of the
steady-state inactivation curve was intermediate (Figure
3A and Table 1). In the I
Na
model, the shift of the
steady-state inactivation curve towards more negative
potentials without alterations in recovery from fast inacti-
vation (discussed later) was simulated by increasing the
fast inactivation rate
b
h
in a voltage-dependent manner
while leaving
a
h
unchanged (see Methods).
No change in the kinetics of recovery from fast inacti-
vation was seen (Figure 4A). In the mathematical model of
I
Na
(Figure 4B), modification of the rate
b
h
did not alter
recovery from inactivation.
As illustrated in Figure 2A, fast inactivation is accelerated
for the currents mediated by mutant channels compared
with WT (arrows in Figure 2A). Individual inactivating
current traces were fitted using a mono-exponential func-
tion and the resulting time constants (
t
) were plotted as a
function of the different voltages (Figure 4C). At voltages
more negative than 5 mV, the mutant
t
values were signifi-
cantly shorter compared with WT. In Figure 4D, the same
analysis is presented for the model WT and mutant I
Na
.
The voltage-dependent behaviour of the
t
values in the
model is comparable with the behaviour observed for the
experiments. In transfected HEK293 cells, we also tested
the entry into the intermediate inactivated state,
13
but no
difference was observed (see Supplementary material
online, Figure S1).
3.3 Action potential simulations
It has been postulated that the clinical manifestations of BrS
are related to the propensity of right epicardial tissue to
premature repolarization, which can explain ST-segment
elevation in the right precordial leads and phase 2
re-excitation by deeper layers of myocardium.
6
This suscep-
tibility of the right epicardial AP to lose its dome is
explained by the predominance of outward currents over
inward currents during phase 1 (notch). During this phase,
the balance of transmembrane currents is principally deter-
mined by the inactivating I
Na
, the activating L-type Ca
2þ
current, and the rapidly activating transient outward K
þ
current, I
to
. Thus, under conditions of high levels of I
to
as
encountered in the right epicardium, the loss of the depolar-
izing contribution of I
Na
caused by its accelerated
Table 1 Biophysical properties of I
Na
recorded in HEK293 cells transfected with wild-type (WT), C1850S, or
1
/
2
WTand
1
/
2
C1850S cDNA. For
the activation and steady-state inactivation curves (Figure 3A), data points from individual cells were fitted with Boltzmann relationship,
and the V
1/2
and K values of the different conditions are presented; *P , 0.05, ***P , 0.001 vs. WT, Bonferroni correction after ANOVA test
Steady state V
1/2
inactivation (mV)
Slope factor: K
inactivation (mV)
V
1/2
activation
(mV)
Slope factor: K
activation (mV)
Time to half recovery from
fast inactivation (ms)
WT 278.6 + 0.8 (n = 9) 6.4 + 0.4 (n =9) 236.9 + 1.4 (n = 7) 7.0 + 0.4 (n = 7) 3.4 + 0.3 (n = 13)
C1850S 290.2 + 0.5*** (n = 7) 5.9 + 0.4 (n =7) 235.5 + 1.3 (n = 5) 7.4 + 0.4 (n = 5) 3.8 + 0.4 (n = 15)
1
/
2
WT þ
1
/
2
C1850S
283.8 + 1.8* (n = 9) 6.9 + 0.3 (n =9) 233.6 + 1.3 (n = 8) 6.4 + 0.6 (n = 8) 3.5 + 0.6 (n =8)
S. Petitprez et al.496
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
inactivation can promote early repolarization.
14
In
addition, it was reported that both hyper- and
hypokalemia could precipitate the clinical manifestations
of BrS.
15,16
This background motivated us to investigate the conse-
quences of the C1850S mutation on the epicardial AP as a
function of the level of I
to
expression and of extracellular
potassium concentration ([K
þ
]
o
). Single cell AP simulations
were conducted using the LRd model. For every [K
þ
]
o
tested (2.5 to 12.0 mmol/L in steps of 0.5 mmol/L), we
determined the critical value of g
to
above which the AP
loses its dome-shaped plateau and repeated this approach
for the WT, HZ, and homozygote C1850S (CS) genotypes.
As shown in Figure 5A, the critical g
to
was, as expected,
the largest for the WT, the smallest for the CS and inter-
mediate for the HZ over the entire range of [K
þ
]
o
tested.
Interestingly, there was a biphasic dependence of critical
g
to
on [K
þ
]
o
for all three genotypes with a maximum near
[K
þ
]
o
= 6 mmol/L. Figure 5B illustrates this loss of the AP
dome with increasing levels of I
to
at a normal [K
þ
]
o
of
4 mmol/L. Increasing g
to
led first to a loss of the dome in
the CS (b), then in the HZ (c) and finally in the WT (d).
Thus, accelerated inactivation of I
Na
may lead to premature
epicardial repolarization under conditions of elevated I
to
expression, which, in vivo, can be encountered in the right
ventricular epicardium.
17
However, the critical g
to
values
Figure 1 Electrocardiogram of the index case, mutations in Na
v
1.5 C-terminus, and expression of wild-type (WT) and C1850S Na
v
1.5 channels. (A) ECG of the
index case at baseline without fever. ( B) Membrane topology of Na
v
1.5. Location of the novel C1850S mutation (black circle) and five BrS mutations located in the
C-terminus (white circles). (C ) Western blot of HEK293 cell lysates expressing WT, C1850S, and 50/50% WT/C1850S channels. Controls are non-transfected cells.
(D) I
Na
density at 378C of HEK293 cells transiently transfected with WT, 50/50% WT/C1850S and C1850S cDNA; n = 14, 10, 14 cells, respectively; ***P , 0.001. n.s.
not significant vs. 50/50% WT/C1850S.
Conduction vs. repolarization disorder in Brugada syndrome 497
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
differed only by a few percent between WT and CS, and the
difference between WT and HZ conditions was ,1% for all
values of [K
þ
]
o
.
It is well known that the clinical manifestations of BrS are
rate-dependent.
18
Moreover, AP characteristics during
pacing at a physiological rate differ from those of an isolated
AP elicited in a resting cell. Therefore, to evaluate the
behaviour of the model cell paced at a physiological rate,
we conducted simulations in which trains of stimuli were
applied during 30 s following the initial 1-min period
without pacing. Figure 5C compares APs in the simulated
WT, HZ and C1850S mutant cells during 30 s of pacing at a
rate of 1 Hz, for [K
þ
]
o
= 4 mmol/L and with g
to
= 1.315 mS/
mF. This value of g
to
was inferior to the thresholds at
which premature repolarization occurred for the isolated
AP (see Figure 5A). While no premature repolarization
occurred during the entire train in the WT cell, premature
repolarization occurred for the second last AP of the train
in the HZ cell (asterisk). In the C1850S cell, premature repo-
larization occurred earlier in the train. When g
to
was sub-
stantially increased (.1.5 mS/mF, not shown), the APs
exhibited an irregular sequence of spike and dome and pre-
mature repolarization morphologies for all three genotypes
and the average rate at which premature repolarization
occurred was not manifestly different.
We then extended the investigation of the critical g
to
level
leading to premature repolarization to the situation of the
paced cell. The critical level was defined as the boundary
between the complete absence of premature repolarization
during the 30-s train and the presence of at least one AP
exhibiting premature repolarization. These simulations
were conducted at pacing rates of 1 and 2 Hz (Figure 5D
and E, respectively). For all three genotypes, the boundaries
exhibited a similar shape as in the single AP simulations but
were shifted towards smaller values of g
to
. This shift was
more prominent at 2 Hz than at 1 Hz. Nevertheless, for
both pacing rates and for all values of [K
þ
]
o
, the critical
g
to
values differed only by a few percent (,1% between
WT and HZ), similar to the single AP simulations.
3.4 The C1850S mutation causes conduction block
in branching tissue
It has recently been proposed that right ventricular fibrosis
may contribute to the genesis of conduction disorders
Figure 2 Wild-type (WT) and C1850S currents recorded in HEK293 cells and
their simulations. (A) Current traces recorded at 378C from cells expressing
WT or C1850S channels in response to a series of 10-ms pulses (inset).
Arrows at 2 ms after onset of the voltage-step indicate the faster inactivation
of C1850S I
Na
compared with WT. The fast transient capacitive current is
blanked for sake of clarity. (B) Reconstruction of WT and C1850S currents
using the original I
Na
of the Luo-Rudy model (WT) and the I
Na
model in
which
b
h
was modified (see ‘Methods’).
Figure 3 Voltage-dependence of activation and inactivation. (A) Activation
(triangles) and inactivation (squares) curves at 378C. Activation properties
were determined from I/V relationships (obtained after applying the step pro-
tocol in inset) by normalizing peak I
Na
to driving force and maximal I
Na
, and
plotting normalized conductance vs. V
m
. Voltage-dependence of steady-state
inactivation was obtained by plotting the normalized peak current (25-ms
test pulse to 220 mV after a 500-ms conditioning pulse, see inserted proto-
col) vs. V
m
. Boltzmann curves were fitted to both activation and steady-state
inactivation data. Averaged values and the number of cells used are pre-
sented in Table 1.(B) Steady-state activation and inactivation curves of the
model I
Na
. These curves were generated using exactly the same voltage-
clamp protocols and analyses as in the experiments.
S. Petitprez et al.498
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
leading to the characteristic BrS ECG.
19,20
Fibrous cardiac
tissue is characterized by sparsely connected strands and
sheets of myocardium intermingled with unexcitable tissue
(Figure 6A), and such structures are known to form the sub-
strate of slow conduction or conduction block.
21
Therefore, we investigated the consequences of the
C1850S mutation on conduction characteristics along a
121-cell strand releasing two branches from its centre
(Figure 6A). These simulations were conducted for a single
AP. In the simulation presented in Figure 6B, the branches
were 20 myocytes long. This branch length was several
times the space constant of the strand, which amounted to
533 mm (approximately five cells) at [K
þ
]
o
= 4 mmol/L (the
space constant was not influenced by g
to
or by the I
Na
geno-
type). This simulation corresponds to scarce lateral coupling
between fibres. The branching strand was characterized by a
mismatch between the current generated by the strand
before the branch point and the increased load represented
by the distal segment of the strand and the two branches.
22
Figure 6B shows AP upstrokes, WT and mutant I
Na
currents
as well as I
to
in the vicinity of the branch point for both the
WT and the HZ (g
to
= 0.5 mS/mF, [ K
þ
]
o
= 4 mmol/L). In the
WT, the current-to-load mismatch resulted in slow upstrokes
and a local conduction delay (0.92 ms). In the HZ, the
reduced total I
Na
shifted the current-to-load balance in
favour of the load, which resulted in conduction block.
In Figure 6C, the dependence of the conduction delay and
the occurrence of block were investigated as a function of
g
to
in strands releasing branches of different lengths (20,
10, 7, 5, and 2 myocytes).
In the absence of I
to
(g
to
= 0), conduction was successful
for both genotypes but, irrespective of branch length, the
conduction delay for the HZ was always longer than for
the WT. For branch lengths of 20, 10, 7 and 5 myocytes,
increasing g
to
led to a progressively increasing conduction
delay, until conduction block occurred. The prolongation
of the conduction delay was much more prominent for the
HZ and the level of g
to
at which block occurred was consider-
ably lower. While, for both genotypes, the g
to
level at which
block occurred was almost the same for branch lengths of 20
and 10 myocytes, this g
to
level was larger for shorter
branches. Only when branches were much shorter (two
Figure 4 Recovery from inactivation and onset of inactivation. (A) Recovery from inactivation recorded at 378C (protocol in inset) was characterized by the time
to half recovery (Table 1). Data were fitted using mono-exponentia l functions (solid lines). Averaged values and the number of cells used are presented in Table 1.
(B) Simulated recovery from inactivation, using the same protocol as in A.(C) Onset of inactivation is accelerated for C1850S at 378C. t values were obtained by
fitting the decaying phase of individual current traces (Figure 2A) with a mono-exponential function and plotted vs. V
m
; n = 7 and 5 cells for WT and C1850S,
respectively; *P , 0.05, **P , 0.01, ***P , 0.001. (D) Corresponding values of t for the simulated I
Na
. These values were obtained using the same protocols
and analyses as in C.
Conduction vs. repolarization disorder in Brugada syndrome 499
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
cells) than the space constant, conduction did not fail and
the conduction delay remained small for both the WT and
the HZ.
Thus, successful conduction in the branched HZ strand is
strongly dependent on I
to
and branch length. This contrasts
with the minimal dependence of the velocity of continuous
conduction on I
to
: as illustrated in Figure 6D, increasing g
to
from 0 to 4 mS/mF slowed conduction by a few percent only.
When the local conduction delay across the branch point
was longer than 1 ms, it provided enough time for I
to
to
activate in the cells immediately before the branch point.
The activated I
to
then acted to diminish the amount of
current available to depolarize the cells beyond the
branch point to threshold. When g
to
was increased, it there-
fore resulted in a prolongation of the conduction delay and
precipitated block. In the presence of mutant I
Na
character-
ized by accelerated inactivation, the marked diminution of
I
Na
after 1
2ms(Figure 2) further exacerbated these altera-
tions of conduction.
In Figure 6E, the (g
to
,[K
þ
]
o
) parameter space was explored
to identify the combinations of g
to
and [K
þ
]
o
leading to con-
duction block in the HZ but not in the WT (light grey area).
In this series of simulations, branches were 20 myocytes
long. [K
þ
]
o
was varied from 2.5 to 12.0 mmol/L in steps of
0.5 mmol/L, and, for each tested [K
þ
]
o
, the critical value
of g
to
at which conduction block occurred was assessed
with a precision of 0.005 mS/mF. Conduction block in the
HZ but not in the WT was obtained in a wide region in the
parameter space, delimited by bell-shaped boundaries.
Furthermore, for all [K
þ
]
o
, there was a several fold differ-
ence between WT and HZ regarding the critical levels of
g
to
at which block occurred. A qualitatively similar differ-
ence was observed for branches 10, 7, and 5 myocytes
long (data not shown). These findings therefore suggest
that in discontinuous cardiac tissue, mutations of I
Na
exhibit-
ing biophysical characteristics similar to those of the C1850S
mutant result in cardiac excitation that is definitely more
prone to slow conduction and conduction block, the key
ingredients of re-entrant arrhythmogenesis.
21
4. Discussion
In this study, we investigated the biophysical characteristics
of a novel SCN5A BrS mutation, C1850S, at physiological
temperature. In addition, we performed AP and conduction
simulations using the LRd model to investigate the
Figure 5 Investigation of the action potential (AP) phenotype in single cell simulations. (A) Boundaries between the spike and dome and premature repolariza-
tion phenotypes in the ([K
þ
]
o
, g
to
) parameter space for the WT, HZ and C1850S genotypes. Combinations of [K
þ
]
o
and g
to
below the corresponding curves produce
a spike and dome phenotype and combinations above the curves lead to loss of the dome and premature repolarization. The points labelled a
d correspond to the
APs shown in B.(B) APs for the three genotypes at a constant [K
þ
]
o
(4 mmol/L) but at increasing levels of g
to
.a:g
to
= 1.450 mS/mF; b: g
to
= 1.515 mS/mF; c: g
to
=
1.535 mS/mF; d: g
to
= 1.600 mS/mF. ( C ) Trains of 30 APs elicited at 1 Hz following a resting period of 1 min, for the three genotypes. [K
þ
]
o
= 4 mmol/L and g
to
=
1.315 mS/mF. Asterisks indic ate premature repolarization. (D) Boundaries between the spike and dome and premature repolarization phenotypes in the ([K
þ
]
o
,
g
to
) parameter space for the three genotypes, for trains of 30 action potentials elicited at 1 Hz. Combinations of [K
þ
]
o
and g
to
above the corresponding curves
lead to loss of the dome and premature repolarization in at least one AP. (E) Same as D, but for trains of 60 APs elicited at 2 Hz.
S. Petitprez et al.500
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
mechanisms underlying the arrhythmogenic alterations of
repolarization and conduction. C1850S channels generated
smaller peak I
Na
, and many of the Na
v
1.5 steady-state and
time-dependent inactivation properties were altered.
These alterations significantly reduce I
Na
during the early
phases of the AP. The simulations suggest that at the hetero-
zygous state, the mutation can depress conduction in myo-
cardium exhibiting sites of current-to-load mismatch, as
exemplified in our study by a myocyte strand releasing
branches.
Figure 6 Conduction in a discontinuous structure is more prone to block in the presence of the C1850S mutation. (A) Schematic of discontinuous cardiac tissue
(top) and its representa tion using the Luo-Rudy model as a strand releasing two branches (bottom). Gray areas represent inexcitable clefts (e.g. connective tissue
in fibrosis) between scarcely connected parallel fibres of cardiac myocytes. (B) Simulated action potentials (V
m
), WT and mutant I
Na
as well as I
to
for the wild-type
(WT) and heterozygote (HZ) genotypes. Branches were 20 myocytes long in this simulation. Data are shown for the 10 cells proximal and distal to the branch
point. The bold traces correspond to the cell at the branch point. g
to
= 0.5 mS/mF and [K
þ
]
o
= 4 mmol/L. ( C ) Conduction delay across the branch point for
[K
þ
]
o
= 4 mmol/L, as a function of g
to
for the WT and HZ genotypes. These simulations were conducted with branches 20, 10, 7, 5, and 2 myocytes long
(labels). Round symbols denote the occurrence of conduction block. (D) Conduction velocity in unbranched WT and HZ strands as a function of g
to
([K
þ
]
o
=
4 mmol/L). (E) Boundaries between successful conduction and conduction block in the ([K
þ
]
o
, g
to
) parameter space for the WT and HZ genotypes (branch
length: 20 cells). Combinations of [K
þ
]
o
and g
to
above the corresponding curves lead to conduction block (white area: successful conduction for both genotypes;
light grey area: successful conduction for the WT genotype but block for the HZ genotype; dark grey area: conduction block for both genotypes). The dot corre-
sponds to the simulation represented in B.
Conduction vs. repolarization disorder in Brugada syndrome 501
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
4.1 Heterogeneity of the biophysical properties of
Brugada syndrome Na
v
1.5 channels
Thus far, the cardiac arrhythmias database
2
of the European
Society of Cardiology lists more than 90 SCN5A BrS
mutations. Among them, five are located in the intracellular
C-terminus of Na
v
1.5 (Figure 1B). Functional characteriz-
ation has been performed for only 23 mutations (three of
them are in the C-terminus).
23
The vast majority of these
mutations reduce the I
Na
mediated by Na
v
1.5. However,
the molecular mechanisms underlying this reduction are
very heterogeneous,
23
and most of the mutant channels
have a unique pattern of alterations. One can divide these
mechanisms into (1) alterations leading to a reduced I
Na
membrane density, and (2) changes in biophysical proper-
ties.
23
The mutation found in this BrS patient replaces
Cys-1850 by a Ser located in a highly conserved part of the
C-terminus of Na
v
1.5. It appears that the first 150 residues
of the C-terminus form a highly structured domain compris-
ing six a-helices.
24
Without more knowledge about the
detailed role of this domain, one cannot but speculate
about the consequences of this mutation. The biophysical
alterations induced by C1850S are mainly related to the I
Na
fast inactivation process. The negative shift of the
steady-state availability curve and faster onset of fast
inactivation are both consistent with a ‘stabilization’ of
inactivated states (fast and slow/intermediate) of mutant
channels. Altogether, these findings suggest that Ser-1850,
by enhancing the inactivation process, contributes to a
‘loss-of-function’ of mutant channels. Our simulations
suggest that acceleration of inactivation alone (determined
by the rate
b
h
) can underlie both the acceleration of inacti-
vation and the shift of steady-state inactivation (by render-
ing inactivated states more stable) without influencing the
activation process.
4.2 Action potential and conduction alterations in
the Brugada syndrome
The molecular, cellular, and arrhythmogenic mechanisms
underlying the ECG alterations seen in BrS are still a
matter of debate.
1
According to the ‘repolarization disorder
hypothesis’, re-entrant arrhythmias are caused by a hetero-
geneous loss of the AP epicardial dome leading to epicardial
dispersion of refractoriness, which forms the substrate for
reentry. According to the ‘conduction disorder hypothesis’,
the typical ECG signs can be explained by slow conduction
and activation delays in the right ventricle (in particular in
the right ventricular outflow tract).
Our computational results support the conduction dis-
order hypothesis. On the one hand, loss of the dome was
more prone to occur in the presence of mutant I
Na
in
single cells. However, the level of I
to
at which abrupt repo-
larization occurred was only minimally smaller (1%) for the
HZ compared with the WT. On the other hand, in discontinu-
ous tissue, conduction delays and block were markedly
potentiated by the presence of mutant I
Na
. Interestingly,
the levels of I
to
conductance at which these manifestations
occurred were strikingly different from those necessary to
result in premature repolarization. Increasing g
to
to
1.3 mS/mF or more was necessary to result in a loss of the
dome in the HZ cell. However, in the branched HZ tissue,
a prominent prolongation of the branching-induced local
conduction delay and conduction block were already
observed at g
to
levels ,0.6 mS/mF for branches consisting
of five myocytes or more. In the original formulation of
right ventricular I
to
by Dumaine et al.,
12
a value of
1.1 mS/mF was used. In human right ventricular tissue,
17
maximal peak I
to
was measured to be 9.8 pA/pF (at
þ60 mV at 368C), which translates, assuming a reversal
potential of 290 mV and a maximal channel open prob-
ability of 0.25, to 0.25 mS/mF. Our study thus suggests
that I
to
-mediated premature repolarization would require
a considerable increase in I
to
density, while conduction
alterations in discontinuous tissue could occur at I
to
levels
comparable with those measured previously. In view of the
recent findings indicating that structural abnormalities
may be involved in the pathogenesis of BrS,
19,20
we
therefore propose the ‘conduction disorder in discontinuous
tissue hypothesis’ as a further mechanism that should
deserve a particular attention.
4.3 Role of I
to
in discontinuous conduction
It has been shown
21
that the L-type calcium current (I
Ca,L
)
greatly contributes to the success of conduction when con-
duction is characterized by local conduction delays that
extend beyond I
Na
inactivation. In such situations, suppres-
sion of I
Ca,L
can precipitate conduction block. Because the
time to peak of I
Ca,L
and I
to
is very similar, our study indi-
cates that I
to
modulates discontinuous conduction via an
analogous mechanism with an opposed polarity. Thus, an
increased I
to
density may precipitate block across a discon-
tinuous structure, while blocking I
to
may result in a recovery
from conduction block. Moreover, our ‘conduction disorder
in discontinuous tissue’ concept is in line with the recent
finding that a BrS ECG pattern can also be caused by
mutations causing loss of function of I
Ca,L
.
25
4.4 Effects of extracellular [K
1
]
Previous work has shown that both hypo- and hyperkalemia
can exacerbate the ECG phenotype of BrS.
15,16
In our simu-
lations, we observed a biphasic dependence on [K
þ
]
o
of the
level of g
to
leading to conduction block in the branching
strand (Figure 6E). This biphasic dependence is reminiscent
of the well-known biphasic dependence of conduction vel-
ocity (CV) on [K
þ
]
o
, which was investigated computationally
in detail by Shaw and Rudy.
26
As shown in their work, when
the resting membrane is hyperpolarized by a decrease of
[K
þ
]
o
, CV decreases because the charge necessary to bring
the membrane to threshold increases. However, during
elevation of [K
þ
]
o
, CV initially increases (supernormal con-
duction) until K
þ
-induced membrane depolarization results
in an increasing fraction of inactivated I
Na
, leading then to
conduction slowing. In the setting of a current-to-load mis-
match, as exemplified by tissue branching, this consider-
ation regarding the charge necessary to reach threshold
(load) vs. the available I
Na
also explains the biphasic beha-
viour of conduction block in the branching strand. Thus,
our results are in agreement with the notion that changes
in [K
þ
]
o
in both directions may exacerbate the clinical mani-
festations of carriers of mutations leading to a loss of I
Na
function. Moreover, it could be hypothesized that, among
other factors, variations in plasma [K
þ
]
o
may play a role in
the transient nature of the BrS ECGs.
S. Petitprez et al.502
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
4.5 Study limitations
The clinical manifestations of BrS are rate-dependent.
18
Because of the computational cost of the LRd model, we
did not investigate the rate-dependence of conduction
characteristics in the branching strands and left the rate-
dependent aspects of BrS during long term pacing of single
cells out of our study. The simulations of conduction there-
fore suffer from the limitation that a single stimulus was
applied instead of continuous pacing until intracellular ion
concentrations have reached their true steady state. As
shown in this study, the AP phenotype and conduction
block in discontinuous tissue critically depend on the
balance between inward and outward currents during early
repolarization. Furthermore, the rate-dependence of the
AP and of conduction is governed by very intricate
dynamic interactions between repolarization, recovery of
I
Na
from inactivation and intracellular Na
þ
and Ca
2þ
over-
load at rapid pacing rates.
10,27
Pacing at different rates or
during different periods of time will therefore affect the
availability of I
Na
, I
Ca,L
, and I
to
during the early phases of
the AP and thus the quantitative outcome of the simulations
in comparison to an isolated stimulus.
BecausenumerousmembranetransportersrelevanttoNa
þ
and K
þ
homeostasis are not incorporated into the LRd model
(e.g. the Na
þ
/glucose symport and the Na
þ
/H
þ
exchanger), it
cannot be absolutely guaranteed that the intracellular ion
concentrations in the model exactly match those in the
human heart in vivo. One can nevertheless speculate that
at a given heart rate, [Na
þ
]
i
might be lower in the presence
of the mutation because less Na
þ
enters the cell during the
AP upstroke. This would increase the driving force for I
Na
.
While this increase might partially compensate for the I
Na
loss due to the mutation, it appears however unlikely that
the extent of this increase would suffice to fully offset the
loss of I
Na
that we observed experimentally.
In addition, since the patient also showed a BrS ECG at
normal temperature, we did not test for effects of higher
temperature on the biophysical characteristics of mutant
cells.
4.6 Conclusions
In conclusion, these results confirm that mutations of the C-
terminal domain of Na
v
1.5 significantly alter the inactivation
properties of the channel, and support the notion that con-
duction alterations in discontinuous tissue may be involved
in the pathogenesis of BrS. Finally, the computational ana-
lyses allowed us to formulate clinically-relevant hypotheses
regarding the cellular basis of arrhythmogenesis in the
context of BrS, which deserve further investigations.
Supplementary material
Supplementary material is available at Cardiovascular
Research online.
Conflict of interest: none declared.
Funding
Supported by grants of the Swiss National Science Foundation
(PP00-110638/1 to HA and 3100A0-100285 to J.P.K.), CardioMet
Center, Fondations Vaudoise de Cardiologie, Rita et Richard
Barme
´,
Carlsberg Foundation to T.J., and Swiss Heart Foundation
and L.&Th. La Roche Foundation (Basel) to D.K.
References
1. Meregalli PG, Wilde AA, Tan HL. Pathophysiological mechanisms of
Brugada syndrome: depolarization disorder, repolarization disorder, or
more? Cardiovasc Res 2005;67:367
378.
2. Study group on molecular basis of arrhythmias. Inherited Arrhythmias
Database http://www.fsm.it/cardmoc/
3. Wilde AA, Antzelevitch C, Borggrefe M, Brugada J, Brugada R, Brugada P
et al. Proposed diagnostic criteria for the Brugada syndrome: consensus
report. Circulation 2002;106:2514
2519.
4. Priori SG, Napolitano C, Schwartz PJ, Bloise R, Crotti L, Ronchetti E. The
elusive link between LQT3 and Brugada syndrome: the role of flecainide
challenge. Circulation 2000;102:945
947.
5. Keller DI, Rougier JS, Kucera JP, Benammar N, Fressart V, Guicheney P
et al. Brugada syndrome and fever: genetic and molecular characterization
of patients carrying SCN5A mutations. Cardiovasc Res 2005;67:510
519.
6. Antzelevitch C, Brugada P, Brugada J, Brugada R, Shimizu W, Gussak I
et al. Brugada syndrome: a decade of progress. Circ Res 2002;91:
1114
1118.
7. Wang Q, Li Z, Shen J, Keating MT. Genomic organization of the human
SCN5A gene enco ding the cardiac sodium channel. Genomics 1996;34:
9
16.
8. van Bemmelen MX, Rougier J-S, Gavillet B, Apotheloz F, Daidie D,
Tateyama M et al. Cardiac voltage-gated sodium channel Na
v
1.5 is
regulated by Nedd4-2 mediated ubiquitination. Circ Res 2004;95:
284
291.
9. Luo CH, Rudy Y. A dynamic model of the cardiac ventricular action
potential. I. Simulations of ionic currents and concentration changes.
Circ Res 1994;74:1071
1096.
10. Faber GM, Rudy Y. Action potential and contractility changes in [Na
þ
]
i
overloaded cardiac myocytes: a simulatio n study. Biophys J 2000;78:
2392
2404.
11. Shaw RM, Rudy Y. Ionic mechanisms of propagation in cardiac tissue.
Roles of the sodium and L-type calcium currents during reduced
excitability and decreased gap junction coupling. Circ Res 1997;81:
727
741.
12. Dumaine R, Towbin JA, Brugada P, Vatta M, Nesterenko DV, Nesterenko VV
et al. Ionic mechanisms responsible for the electrocardiographic pheno-
type of the Brugada syndrome are temperature dependent. Circ Res
1999;85:803
809.
13. Wang DW, Makita N, Kitabatake A, Balser JR, George AL Jr. Enhanced Na
þ
channel intermediate inactivation in Brugada syndrome. Circ Res 2000;
87:E37
E43.
14. Yan GX, Antzelevitch C. Cellular basis for the Brugada syndrome and
other mechanisms of arrhythmogenesis associated with ST-segment
elevation. Circulation 1999;100:1660
1666.
15. Littmann L, Monroe MH, Taylor L III, Brearley J. The hyperkalemic
Brugada sign. J Electrocardiol 2007;40:53
59.
16. Araki T, Konno T, Itoh H, Ino H, Shimizu M. Brugada syndrome with ventri-
cular tachycardia and fibrillatio n related to hypokalemia. Circ J 2003;67:
93
95.
17. Li GR, Feng J, Yue L, Carrier M. Transmural heterogeneity of action
potentials and Ito1 in myocytes isolated from the human right ventricle.
Am J Physiol Heart Circ Physiol 1998;275:H369
H377.
18. Extramiana F, Seitz J, Maison-Blanche P, Badilini F, Haggui A, Takatsuki S
et al. Quantitative assessment of ST segment elevation in Brugada
patients. Heart Rhythm 2006;3:1175
1181.
19. Coronel R, Casini S, Koopmann TT, Wilms-Schopman FJG, Verkerk AO, de
Groot JR et al. Right ventricular fibrosis and conduction delay in a patient
with clinical signs of Brugada syndrome: a combined electrophysiological,
genetic, histopathologic, and computational study. Circulation 2005;
112:2769
2777.
20. Frustaci A, Priori SG, Pieroni M, Chimenti C, Napolitano C, Rivolta I et al.
Cardiac histological substrate in patients with clinical phenotype of
Brugada syndrome. Circulation 2005;112:3680
3687.
21. Kle
´b
er AG, Rudy Y. Basic mechanisms of cardiac impulse propagation and
associated arrhythmias. Physiol Rev 2004;84:431
488.
22. Kucera JP, Rudy Y. Mechanistic insights into very slow conduction in
branching cardiac tissue: a model study. Circ Res 2001;89:799
806.
Conduction vs. repolarization disorder in Brugada syndrome 503
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
23. Tan H. Biophysical analysis of mutant sodium channels in Brugada syndrome.
In: Antzelevitch C, Brugada P, Brugada J, Brugada R ed.The Brugada syn-
drome: From Bench to Bedside. 1st ed. Blackwell Publishing; 2005. p26
41.
24. Cormier JW, Rivolta I, Tateyama M, Yang AS, Kass RS. Secondary structure
of the human cardiac Na
þ
channel C terminus: evidence for a role of
helical structures in modulation of channel inactivation. J Biol Chem
2002;277:9233
9241.
25. Antzelevitch C, Pollevick GD, Cordeiro JM, Casis O, Sanguinetti MC,
Aizawa Y et al. Loss-of-function mutations in the cardiac calcium
channel underlie a new clinical entity characterized by ST-segment
elevation, short QT intervals, and sudden cardiac death. Circulation
2007;115:442
449.
26. Shaw RM, Rudy Y. Electrophysiologic effects of acute myocardial
ischemia: a mechanistic investigation of action potential conduction
and conduction failure. Circ Res 1997;80:124
138.
27. Kondratyev AA, Ponard JG, Munteanu A, Rohr S, Kucera JP. Dynam ic
changes of cardiac conduction during rapid pacing. Am J Physiol Heart
Circ Physiol 2006;292:H1796
H1811.
S. Petitprez et al.504
by guest on June 8, 2013http://cardiovascres.oxfordjournals.org/Downloaded from
... By contrast, it is well known that both loss-offunction I Na mutations in BrS and Na channel blocking drugs do more than just reduce peak I Na amplitude. They also alter Na channel kinetics by shifting steady-state activation and inactivation curves and altering inactivation and recovery time constants 26,[36][37][38][39][40][41][42][43][44][45][46][47][48][49][50] (Table 1). The goal of this computational study was to explore whether such kinetic alterations, in combination with reduced I Na peak amplitude, can account for the clinically observed robust effects of Na channel mutations at causing BrS as well as Na channel blockers unmasking its features, including P2R-mediated life-threatening arrhythmias. ...
... The AP did not exhibit a spike-and-dome pattern, and conduction was normal. However, when I Na features were altered to mimic quantitatively the reported effects of BrS Na channel mutations and Na channel blocking drugs on I Na kinetics as well as amplitude, 26,[36][37][38][39][40][41][42]48 certain combinations induced both spike-and-dome AP morphology and P2R, as illustrated in the example in Figure 2B. Figure 2C shows another example of a P2R-induced PVC that propagated throughout the cable even as the paced AP upstroke developed conduction block at the cable's lower end. Figure 2D shows an example in which multiple spontaneous P2R-induced PVCs occurred at intervals of <100 ms. ...
... Our findings are consistent with experimental evidence that many of the Na channel mutations causing BrS exhibit a negative voltage shift in steady-state inactivation or a reduction in inactivation time constant(Table 1).26,[36][37][38][39][40][41][42][43][44][45]48 Moreover, a reported Na channel mutation that increases the inactivation time constant and has no effect on steady-state inactivation showed no features of BrS despite significantly reducing I Na amplitude, although the resulting conduction abnormalities caused a high incidence of sudden cardiac death.74 ...
... Consistent with this, the EF-hand domain and III-IV linker peptide show qualitatively weaker binding with a K d of about 900 μM. Several residues with perturbed chemical shifts, including L1786, S1787, D1790, Y1795, and C1850, have been previously associated with disease-causing mutations 24,25,30,31,[34][35][36][37] . Altered interactions between the EF-hand domain and III-IV linker may thus be responsible for part or all of the disease phenotype. ...
... As it is clear from both the steady-state inactivation curve and on individual traces, the currents remain at a level of ~9% Fig. S3). This so-called late sodium current has been implicated in cardiac arrhythmia, underlying Long-QT syndrome type 3 37,42 . It thus appears that the interaction between the III-IV linker and EF-hand domain has a dual role in preventing inactivation at hyperpolarized potentials, and in producing a stable inactivated state 19 . ...
... Using NMR spectroscopy, we mapped the interface of the III-IV linker on the Na V 1.5 EF-hand domain. Residues showing chemical shift perturbations upon linker binding include a number of sites for mutations associated with Brugada and Long-QT syndromes, namely L1786Q, S1787N, D1790G, Y1795, C1850 24,25,30,31,[34][35][36][37] . Table S3 outlines the disease phenotypes and reported functional impacts. ...
Article
Full-text available
Voltage-gated sodium channels (NaV) are responsible for the rapid depolarization of many excitable cells. They readily inactivate, a process where currents diminish after milliseconds of channel opening. They are also targets for a multitude of disease-causing mutations, many of which have been shown to affect inactivation. A cluster of disease mutations, linked to Long-QT and Brugada syndromes, is located in a C-terminal EF-hand like domain of NaV1.5, the predominant cardiac sodium channel isoform. Previous studies have suggested interactions with the III-IV linker, a cytosolic element directly involved in inactivation. Here we validate and map the interaction interface using isothermal titration calorimetry (ITC) and NMR spectroscopy. We investigated the impact of various disease mutations on the stability of the domain, and found that mutations that cause misfolding of the EF-hand domain result in hyperpolarizing shifts in the steady-state inactivation curve. Conversely, mutations in the III-IV linker that disrupt the interaction with the EF-hand domain also result in large hyperpolarization shifts, supporting the interaction between both elements in intact channels. Disrupting the interaction also causes large late currents, pointing to a dual role of the interaction in reducing the population of channels entering inactivation and in stabilizing the inactivated state.
... Ajmaline is usually described as a sodium-channel blocker (3), and most research into the mechanism of BrS has centered around this idea that the sodium channel is somehow impaired in BrS (21,22), and thus the genetics research has placed much emphasis on sodium channel gene mutations, especially the gene SCN5A, whereas systematic studies on other genes are Abbreviations: AP, action potential; BrS, Brugada syndrome; Ca 2+ , calcium; ECG, electrocardiogram; HEK, human embryonic kidney; HERG, human ether a-go-go related gene; I Ca−L , L-type calcium current; I K , delayed rectifier potassium current; I K1 , inwardly rectifying potassium current; I K(ATP) , ATP-sensitive potassium current; I K,end , the current measured at the end of 300 ms depolarizing pulse; I Kur , ultrarapid outward potassium current; I Na , sodium current; I to , transient outward potassium current; K + , potassium; K ATP , ATP-sensitive potassium channel; Na + , sodium; NCX, sodium-calcium exchanger; NPA, N-propyl ajmaline, a quaternary derivative of ajmaline; PVCs, pre-mature ventricular complexes; RV, right ventricle; Vas, ventricular arrhythmias; VF, ventricular fibrillation; VT, ventricular tachycardia. ...
Article
Full-text available
Ajmaline is an anti-arrhythmic drug that is used to unmask the type-1 Brugada syndrome (BrS) electrocardiogram pattern to diagnose the syndrome. Thus, the disease is defined at its core as a particular response to this or other drugs. Ajmaline is usually described as a sodium-channel blocker, and most research into the mechanism of BrS has centered around this idea that the sodium channel is somehow impaired in BrS, and thus the genetics research has placed much emphasis on sodium channel gene mutations, especially the gene SCN5A , to the point that it has even been suggested that only the SCN5A gene should be screened in BrS patients. However, pathogenic rare variants in SCN5A are identified in only 20–30% of cases, and recent data indicates that SCN5A variants are actually, in many cases, prognostic rather than diagnostic, resulting in a more severe phenotype. Furthermore, the misconception by some that ajmaline only influences the sodium current is flawed, in that ajmaline actually acts additionally on potassium and calcium currents, as well as mitochondria and metabolic pathways. Clinical studies have implicated several candidate genes in BrS, encoding not only for sodium, potassium, and calcium channel proteins, but also for signaling-related, scaffolding-related, sarcomeric, and mitochondrial proteins. Thus, these proteins, as well as any proteins that act upon them, could prove absolutely relevant in the mechanism of BrS.
... Instead, they take information about variants, for example the site of the substitution, the prevalence of this variant in the population, or the physico-chemical properties of the affected amino acids, and then try to establish a direct, 'black-box' link to pathogenicity. In several cases, however, the proarrhythmic mechanisms associated with a specific SCN5A mutation have been established using a systems approach that follows genetic screening by current measurements in expression systems, and then uses mathematical modelling to show how they can lead to the observed cell, tissue, and ECG effects [10][11][12] . While this approach has traditionally been applied to investigate single mutations with known clinical phenotypes, a study on long-QT syndrome type 1 by Hoefen et al. 13 clearly demonstrated that (1) this approach can be extended to cover multiple mutations (provided their effects on the ion current are known), and (2) that the resulting models can predict clinical outcomes and improve risk stratification. ...
Article
Full-text available
Mutations in SCN5A can alter the cardiac sodium current INa and increase the risk of potentially lethal conditions such as Brugada and long-QT syndromes. The relation between mutations and their clinical phenotypes is complex, and systems to predict clinical severity of unclassified SCN5A variants perform poorly. We investigated if instead we could predict changes to INa, leaving the link from INa to clinical phenotype for mechanistic simulation studies. An exhaustive list of nonsynonymous missense mutations and resulting changes to INa was compiled. We then applied machine-learning methods to this dataset, and found that changes to INa could be predicted with higher sensitivity and specificity than most existing predictors of clinical significance. The substituted residues’ location on the protein correlated with channel function and strongly contributed to predictions, while conservedness and physico-chemical properties did not. However, predictions were not sufficiently accurate to form a basis for mechanistic studies. These results show that changes to INa, the mechanism through which SCN5A mutations create cardiac risk, are already difficult to predict using purely in-silico methods. This partly explains the limited success of systems to predict clinical significance of SCN5A variants, and underscores the need for functional studies of INa in risk assessment.
... During arrhythmias, takeoff potential (minimal diastolic potential) prior to the next upstroke was depolarized to around ≈−65 mV in contrast to ≈−75 mV during SR. This 10 mV shift has profound effect on Na + channel availability, due to increased voltage-dependent inactivation (Petitprez et al., 2008). This effect is observed by Resting membrane potential (RMP), (C) atrial effective refractory period (aERP), (D) action potential duration at 90% repolarization (APD90), (E) action potential duration at 50% repolarization (APD50), (F) upstroke velocity (dV/dt max ), with the time-matched control group being unchanged (from 136 ± 12 to 137 ± 22 mV/s) and the flecainide group being reduced (from 152 ± 9 to 103 ± 6 mV/s), and (G) data on AF duration normalized to baseline (%), at pacing rates of 5 Hz. ...
Article
Full-text available
Sympathetic and vagal activation is linked to atrial arrhythmogenesis. Here we investigated the small conductance Ca²⁺-activated K⁺ (SK)-channel pore-blocker N-(pyridin-2-yl)-4-(pyridine-2-yl)thiazol-2-amine (ICA) on action potential (AP) and atrial fibrillation (AF) parameters in isolated rat atria during β-adrenergic [isoprenaline (ISO)] and muscarinic M2 [carbachol (CCh)] activation. Furthermore, antiarrhythmic efficacy of ICA was benchmarked toward the class-IC antiarrhythmic drug flecainide (Fleca). ISO increased the spontaneous beating frequency but did not affect other AP parameters. As expected, CCh hyperpolarized resting membrane potential (-6.2 ± 0.9 mV), shortened APD90 (24.2 ± 1.6 vs. 17.7 ± 1.1 ms), and effective refractory period (ERP; 20.0 ± 1.3 vs. 15.8 ± 1.3 ms). The duration of burst pacing triggered AF was unchanged in the presence of CCh compared to control atria (12.8 ± 5.3 vs. 11.2 ± 3.6 s), while β-adrenergic activation resulted in shorter AF durations (3.3 ± 1.7 s) and lower AF-frequency compared to CCh. Treatment with ICA (10 μM) in ISO -stimulated atria prolonged APD90 and ERP, while the AF burden was reduced (7.1 ± 5.5 vs. 0.1 ± 0.1 s). In CCh-stimulated atria, ICA treatment also resulted in APD90 and ERP prolongation and shorter AF durations. Fleca treatment in CCh-stimulated atria prolonged APD90 and ERP and abbreviated the AF duration to a similar extent as with ICA. Muscarinic activated atria constitutes a more arrhythmogenic substrate than β-adrenoceptor activated atria. Pharmacological inhibition of SK channels by ICA is effective under both conditions and equally efficacious to Fleca.
... Our simulations are certainly not the first to assess the effects of a specific mutation in SCN5A on the electrical activity of cardiac myocytes. However, in most studies conducted thus far, e.g., [54][55][56][57][58][59][60][61][62][63][64][65][66][67][68][69][70][71], the emphasis was on ventricular cells, some also considering atrial or Purkinje cells. Fewer studies, e.g., by Butters et al. [72], Wu et al. [73], and Zhang et al. [74], focused on the effects on SA nodal cells, using the rabbit SA nodal cell models by Zhang et al. [75,76]. ...
Article
Full-text available
The SCN5A gene encodes the pore-forming α-subunit of the ion channel that carries the cardiac fast sodium current (INa). The 1795insD mutation in SCN5A causes sinus bradycardia, with a mean heart rate of 70 beats/min in mutation carriers vs. 77 beats/min in non-carriers from the same family (lowest heart rate 41 vs. 47 beats/min). To unravel the underlying mechanism, we incorporated the mutation-induced changes in INa into a recently developed comprehensive computational model of a single human sinoatrial node cell (Fabbri–Severi model). The 1795insD mutation reduced the beating rate of the model cell from 74 to 69 beats/min (from 49 to 43 beats/min in the simulated presence of 20 nmol/L acetylcholine). The mutation-induced persistent INa per se resulted in a substantial increase in beating rate. This gain-of-function effect was almost completely counteracted by the loss-of-function effect of the reduction in INa conductance. The further loss-of-function effect of the shifts in steady-state activation and inactivation resulted in an overall loss-of-function effect of the 1795insD mutation. We conclude that the experimentally identified mutation-induced changes in INa can explain the clinically observed sinus bradycardia. Furthermore, we conclude that the Fabbri–Severi model may prove a useful tool in understanding cardiac pacemaker activity in humans.
Article
Full-text available
Voltage-dependent Na+ channel activation underlies action potential generation fundamental to cellular excitability. In skeletal and cardiac muscle this triggers contraction via ryanodine-receptor (RyR)-mediated sarcoplasmic reticular (SR) Ca2+ release. We here review potential feedback actions of intracellular [Ca2+] ([Ca2+]i) on Na+ channel activity, surveying their structural, genetic and cellular and functional implications, translating these to their possible clinical importance. In addition to phosphorylation sites, both Nav1.4 and Nav1.5 possess potentially regulatory binding sites for Ca2+ and/or the Ca2+-sensor calmodulin in their inactivating III-IV linker and C-terminal domains (CTD), where mutations are associated with a range of skeletal and cardiac muscle diseases. We summarize in vitro cell-attached patch clamp studies reporting correspondingly diverse, direct and indirect, Ca2+ effects upon maximal Nav1.4 and Nav1.5 currents (Imax) and their half-maximal voltages (V1/2) characterizing channel gating, in cellular expression systems and isolated myocytes. Interventions increasing cytoplasmic [Ca2+]i down-regulated Imax leaving V1/2 constant in native loose patch clamped, wild-type murine skeletal and cardiac myocytes. They correspondingly reduced action potential upstroke rates and conduction velocities, causing pro-arrhythmic effects in intact perfused hearts. Genetically modified murine RyR2-P2328S hearts modelling catecholaminergic polymorphic ventricular tachycardia (CPVT), recapitulated clinical ventricular and atrial pro-arrhythmic phenotypes following catecholaminergic challenge. These accompanied reductions in action potential conduction velocities. The latter were reversed by flecainide at RyR-blocking concentrations specifically in RyR2-P2328S as opposed to wild-type hearts, suggesting a basis for its recent therapeutic application in CPVT. We finally explore the relevance of these mechanisms in further genetic paradigms for commoner metabolic and structural cardiac disease.
Thesis
The rhythmic beating of the heart results from the interactions between billions of heart muscle cells. This complex process involves factors at the genetic, molecular, cell, tissue and whole-organ scales. To study them, multi-scale (computer) models are employed. In this thesis, a software application is introduced that makes multi-scale models easier to work with, and thereby more broadly applicable. In addition, novel experimental data is presented on the effects of genetic mutations and natural variability on the function of ion channels in the heart. These results add to our understanding of the heart's rhythm, and can accelerate the pace of research into cardiac rhythm disorders.
Article
Full-text available
Limited information is available about transmural heterogeneity in cardiac electrophysiology in man. The present study was designed to evaluate heterogeneity of cardiac action potential (AP), transient outward K+ current (Ito1) and inwardly rectifying K+ current (IK1) in human right ventricle. AP and membrane currents were recorded using whole cell current- and voltage-clamp techniques in myocytes isolated from subepicardial, midmyocardial, and subendocardial layers of the right ventricle of explanted failing human hearts. AP morphology differed among the regional cell types. AP duration (APD) at 0.5-2 Hz was longer in midmyocardial cells (M cells) than in subepicardial and subendocardial cells. At room temperature, observed Ito1, on step to +60 mV, was significantly greater in subepicardial (6.9 +/- 0.8 pA/pF) and M cells (6.0 +/- 1.1 pA/pF) than in subendocardial cells (2.2 +/- 0.7 pA/pF, P < 0.01). Slower recovery of Ito1 was observed in subendocardial cells. The half-inactivation voltage of Ito1 was more negative in subendocardial cells than in M and subepicardial cells. At 36 degrees C, the density of Ito1 increased, the time-dependent inactivation and reactivation accelerated, and the frequency-dependent reduction attenuated in all regional cell types. No significant difference was observed in IK1 density among the regional cell types. The results indicate that M cells in humans, as in canines, show the greatest APD and that a gradient of Ito1 density is present in the transmural ventricular wall. Therefore, the human right ventricle shows significant transmural heterogeneity in AP morphology and Ito1 properties.
Article
Asyndrome characterized by ST-segment elevation in right precordial leads (V1 to V3) that is unrelated to ischemia, electrolyte disturbances, or obvious structural heart disease was reported as early as 1953,1 but was first described as a distinct clinical entity associated with a high risk of sudden cardiac death in 1992.2–4⇓⇓ The Brugada syndrome is a familial disease that displays an autosomal dominant mode of transmission, with incomplete penetrance and an incidence ranging between 5 and 66 per 10 000. In regions of Southeast Asia where it is endemic, the clinical presentation of Brugada syndrome is distinguished by a male predominance (8:1 ratio of male:female) and the appearance of arrhythmic events at an average age of 40 years (range: 1 to 77 years).2,5⇓ Although a number of candidate genes are considered plausible, thus far the syndrome has been linked only to mutations in SCN5A , the gene encoding for the α subunit of the sodium channel.6 A number of ambiguities exist concerning the diagnosis of Brugada syndrome. The electrocardiographic signature of the syndrome is dynamic and often concealed, but can be unmasked by potent sodium channel blockers such as flecainide, ajmaline, and procainamide,7 although the specificity of this effect for uncovering patients at risk for sudden death has been an issue of concern. A recent report by Remme et al8 has shown that the number of idiopathic ventricular fibrillation patients diagnosed as having Brugada syndrome is a sensitive function of the diagnostic criteria applied. What are the proper diagnostic criteria for identifying Brugada syndrome? A definitive answer to this question has been out of reach and is the reason for the establishment of a special Arrhythmia Working Group of the European Society of Cardiology that met from August 31 to …
Chapter
Brugada Syndrome and Sodium Channel MutationsStructure and Function of the Cardiac Na ChannelBiophysical Analysis of Mutant Na Channels: The Patch-Clamp TechniqueReduced Na CurrentBiophysical Mechanisms of Na Current ReductionOverlap Syndromes and Modulating FactorsConclusions References
Article
Typescript. Department of Biomedical Engineering. Thesis (M.S.)--Case Western Reserve University, 2000. Includes bibliographical references (leaves 43-48).
Article
A mathematical model of the cardiac ventricular action potential is presented. In our previous work, the membrane Na+ current and K+ currents were formulated. The present article focuses on processes that regulate intracellular Ca2+ and depend on its concentration. The model presented here for the mammalian ventricular action potential is based mostly on the guinea pig ventricular cell. However, it provides the framework for modeling other types of ventricular cells with appropriate modifications made to account for species differences. The following processes are formulated: Ca2+ current through the L-type channel (ICa), the Na(+)-Ca2+ exchanger, Ca2+ release and uptake by the sarcoplasmic reticulum (SR), buffering of Ca2+ in the SR and in the myoplasm, a Ca2+ pump in the sarcolemma, the Na(+)-K+ pump, and a nonspecific Ca(2+)-activated membrane current. Activation of ICa is an order of magnitude faster than in previous models. Inactivation of ICa depends on both the membrane voltage and [Ca2+]i. SR is divided into two subcompartments, a network SR (NSR) and a junctional SR (JSR). Functionally, Ca2+ enters the NSR and translocates to the JSR following a monoexponential function. Release of Ca2+ occurs at JSR and can be triggered by two different mechanisms, Ca(2+)-induced Ca2+ release and spontaneous release. The model provides the basis for the study of arrhythmogenic activity of the single myocyte including afterdepolarizations and triggered activity. It can simulate cellular responses under different degrees of Ca2+ overload. Such simulations are presented in our accompanying article in this issue of Circulation Research.
Article
The voltage-gated cardiac sodium channel, SCN5A, is responsible for the initial upstroke of the action potential. Mutations in the human SCN5A gene cause susceptibility to cardiac arrhythmias and sudden death in the long QT syndrome (LQT). In this report we characterize the genomic structure of SCN5A. SCN5A consists of 28 exons spanning approximately 80 kb on chromosome 3p21. We describe the sequences of all intron/exon boundaries and a dinucleotide repeat polymorphism in intron 16. Oligonucleotide primers based on exon-flanking sequences amplify all SCN5A exons by PCR. This work establishes the complete genomic organization of SCN5A and will enable high-resolution analyses of this locus for mutations associated with LQT and other phenotypes for which SCN5A may be a candidate gene.
Article
A multicellular ventricular fiber model was used to determine mechanisms of slowed conduction and conduction failure during acute ischemia. We simulated the three major pathophysiological component conditions of acute ischemia: elevated [K+]o, acidosis, and anoxia. Elevated [K+]o was the major determinant of conduction, causing supernormal conduction, depressed conduction, and conduction block as [K+]o was gradually increased from 4.5 to 14.4 mmol/L. Only elevated [K+]o caused conduction failure when varied within the range reported for acute ischemia. Before block, depressed upstrokes consisted of two distinct components: the first to the fast Na+ current (INa) and the second to the L-type Ca2+ current (ICa(L)). Even in highly depressed conduction, excitability was maintained by INa, with conduction block occurring at 95% INa inactivation. However, because ICa(L) supported the later phase of the depressed upstroke, ICa(L) enhanced conduction and delayed block by increasing the electrotonic source current. At [K+]o = 18 mmol/L, slow action potentials generated by ICa(L) were obtained with 10% ICa(L) augmentation. However, in the presence of acidosis and anoxia, significantly larger (120%) ICa(L) augmentation was required. The depressant effect was due mostly to anoxic activation of outward ATP-sensitive K+ current, which counteracts inward ICa(L) and, by lowering the action potential amplitude, decreases the electrotonic current available to depolarize downstream cells. The simulations highlight the interactive nature of electrophysiological ischemic changes during propagation and demonstrate that both membrane changes and load factors (by downstream fiber) must be considered.
Article
In cardiac tissue, reduced membrane excitability and reduced gap junction coupling both slow conduction velocity of the action potential. However, the ionic mechanisms of slow conduction for the two conditions are very different. We explored, using a multicellular theoretical fiber, the ionic mechanisms and functional role of the fast sodium current, INa, and the L-type calcium current, ICa(L), during conduction slowing for the two fiber conditions. A safety factor for conduction (SF) was formulated and computed for each condition. Reduced excitability caused a lower SF as conduction velocity decreased. In contrast, reduced gap junction coupling caused a paradoxical increase in SF as conduction velocity decreased. The opposite effect of the two conditions on SF was reflected in the minimum attainable conduction velocity before failure: decreased excitability could reduce velocity to only one third of control (from 54 to 17 cm/s) before failure occurred, whereas decreased coupling could reduce velocity to as low as 0.26 cm/s before block. Under normal conditions and conditions of reduced excitability, ICa(L) had a minimal effect on SF and on conduction. However, ICa(L) played a major role in sustaining conduction when intercellular coupling was reduced. This phenomenon demonstrates that structural, nonmembrane factors can cause a switch of intrinsic membrane processes that support conduction. High intracellular calcium concentration, [Ca]i, lowered propagation safety and caused earlier block when intercellular coupling was reduced. [Ca]i affected conduction via calcium-dependent inactivation of ICa(L). The increase of safety factor during reduced coupling suggests a major involvement of uncoupling in stable slow conduction in infarcted myocardium, making microreentry possible. Reliance on ICa(L) for this type of conduction suggests ICa(L) as a possible target for antiarrhythmic drug therapy.
Article
Background—The Brugada syndrome is characterized by marked ST-segment elevation in the right precordial ECG leads and is associated with a high incidence of sudden and unexpected arrhythmic death. Our study examines the cellular basis for this syndrome. Methods and Results—Using arterially perfused wedges of canine right ventricle (RV), we simultaneously recorded transmembrane action potentials from 2 epicardial and 1 endocardial sites, together with unipolar electrograms and a transmural ECG. Loss of the action potential dome in epicardium but not endocardium after exposure to pinacidil (2 to 5 μmol/L), a K+ channel opener, or the combination of a Na+ channel blocker (flecainide, 7 μmol/L) and acetylcholine (ACh, 2 to 3 μmol/L) resulted in an abbreviation of epicardial response and a transmural dispersion of repolarization, which caused an ST-segment elevation in the ECG. ACh facilitated loss of the action potential dome, whereas isoproterenol (0.1 to 1 μmol/L) restored the epicardial dome, thus reducing or eliminating the ST-segment elevation. Heterogeneous loss of the dome caused a marked dispersion of repolarization within the epicardium and transmurally, thus giving rise to phase 2 reentrant extrasystole, which precipitated ventricular tachycardia (VT) and ventricular fibrillation (VF). Transient outward current (Ito) block with 4-aminopyridine (1 to 2 mmol/L) or quinidine (5 μmol/L) restored the dome, normalized the ST segment, and prevented VT/VF. Conclusions—Depression or loss of the action potential dome in RV epicardium creates a transmural voltage gradient that may be responsible for the ST-segment elevation observed in the Brugada syndrome and other syndromes exhibiting similar ECG manifestations. Our results also demonstrate that extrasystolic activity due to phase 2 reentry can arise in the intact wall of the canine RV and serve as the trigger for VT/VF. Our data point to Ito block (4-aminopyridine, quinidine) as an effective pharmacological treatment.
Article
The Brugada syndrome is a major cause of sudden death, particularly among young men of Southeast Asian and Japanese origin. The syndrome is characterized electrocardiographically by an ST-segment elevation in V1 through V3 and a rapid polymorphic ventricular tachycardia that can degenerate into ventricular fibrillation. Our group recently linked the disease to mutations in SCN5A, the gene encoding for the alpha subunit of the cardiac sodium channel. When heterologously expressed in frog oocytes, electrophysiological data recorded from the Thr1620Met missense mutant failed to adequately explain the electrocardiographic phenotype. Therefore, we sought to further characterize the electrophysiology of this mutant. We hypothesized that at more physiological temperatures, the missense mutation may change the gating of the sodium channel such that the net outward current is dramatically augmented during the early phases of the right ventricular action potential. In the present study, we test this hypothesis by expressing Thr1620Met in a mammalian cell line, using the patch-clamp technique to study the currents at 32 degrees C. Our results indicate that Thr1620Met current decay kinetics are faster when compared with the wild type at 32 degrees C. Recovery from inactivation was slower for Thr1620Met at 32 degrees C, and steady-state activation was significantly shifted. Our findings explain the features of the ECG of Brugada patients, illustrate for the first time a cardiac sodium channel mutation of which the arrhythmogenicity is revealed only at temperatures approaching the physiological range, and suggest that some patients may be more at risk during febrile states.