Download The cardiac pacemaker current Journal of Molecular and Cellular

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Cyclic nucleotide–gated ion channel wikipedia , lookup

Cell culture wikipedia , lookup

Cell encapsulation wikipedia , lookup

Tissue engineering wikipedia , lookup

List of types of proteins wikipedia , lookup

Cellular differentiation wikipedia , lookup

Mechanosensitive channels wikipedia , lookup

Amitosis wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Transcript
Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
Contents lists available at ScienceDirect
Journal of Molecular and Cellular Cardiology
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / y j m c c
Review article
The cardiac pacemaker current
Mirko Baruscotti ⁎, Andrea Barbuti, Annalisa Bucchi
Department of Biomolecular Sciences and Biotechnology, Laboratory of Molecular Physiology and Neurobiology, Università degli Studi di Milano; Centro Interuniversitario di Medicina
Molecolare e Biofisica Applicata (CIMMBA), via Celoria 26, 20133 Milano, Italy
a r t i c l e
i n f o
Article history:
Received 29 April 2009
Received in revised form 15 June 2009
Accepted 26 June 2009
Available online 8 July 2009
Keywords:
Pacemaker current
SAN
HCN clones
HCN knockout mice
HCN-associated pathologies
a b s t r a c t
In mammals cardiac rate is determined by the duration of the diastolic depolarization of sinoatrial node
(SAN) cells which is mainly determined by the pacemaker If current. f-channels are encoded by four
members of the hyperpolarization-activated cyclic nucleotide-gated gene (HCN1–4) family. HCN4 is the most
abundant isoform in the SAN, and its relevance to pacemaking has been further supported by the discovery of
four loss-of-function mutations in patients with mild or severe forms of cardiac rate disturbances. Due to its
selective contribution to pacemaking, the If current is also the pharmacological target of a selective heart
rate-reducing agent (ivabradine) currently used in the clinical practice. Albeit to a minor extent, the If current
is also present in other spontaneously active myocytes of the cardiac conduction system (atrioventricular
node and Purkinje fibres). In working atrial and ventricular myocytes f-channels are expressed at a very low
level and do not play any physiological role; however in certain pathological conditions over-expression of
HCN proteins may represent an arrhythmogenic mechanism. In this review some of the most recent findings
on f/HCN channels contribution to pacemaking are described.
© 2009 Elsevier Inc. All rights reserved.
Contents
The mechanism of cardiac pacemaking and the If current .
Basic biophysical properties of the pacemaker current . .
Molecular structure of pacemaker channels . . . . . . .
Pacemaker channels during embryonic development . . .
Pacemaker channels in the adult heart . . . . . . . . .
5.1.
SAN . . . . . . . . . . . . . . . . . . . . . .
5.2.
Atrioventricular node (AVN) . . . . . . . . . . .
5.3.
Purkinje fibres (PFs) . . . . . . . . . . . . . . .
5.4.
Working myocardium . . . . . . . . . . . . . .
6.
Basis of functional heterogeneity of pacemaker currents .
6.1.
MiRP1 . . . . . . . . . . . . . . . . . . . . .
6.2.
PI(4,5)P2 . . . . . . . . . . . . . . . . . . . .
6.3.
Caveolin 3 . . . . . . . . . . . . . . . . . . .
7.
Genetics of HCN: HCN knockout mice and HCN-associated
8.
Biological pacemaker . . . . . . . . . . . . . . . . . .
9.
f-channels blockers. . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . .
1.
2.
3.
4.
5.
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
pathologies in
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
1. The mechanism of cardiac pacemaking and the If current
Cardiac pacemaking originates in the sinoatrial node (SAN) as a
consequence of spontaneous firing of rhythmic action potentials
⁎ Corresponding author. Tel.: +39 02 5031 4931; fax: +39 02 5031 4932.
E-mail address: [email protected] (M. Baruscotti).
0022-2828/$ – see front matter © 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.yjmcc.2009.06.019
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
humans
. . . .
. . . .
. . . .
. . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
55
56
56
57
57
58
58
58
59
59
59
59
59
59
61
61
61
61
generated by specialized myocytes. Although the electrical behavior of
a typical SAN cell differs in several aspects from that of a working
myocyte, the functional hallmark can be precisely identified in the
events that take place during the diastolic interval. During this phase
atrial and ventricular myocytes rest in a standby-like condition at a
stable voltage (∼−80 mV); a quite different situation characterizes
SAN cells, where the cell potential slowly creeps up from the
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
56
maximum diastolic potential of about −60 mV to the threshold for
the ignition of a new action potential. Since this time interval sets the
pace of the heart, this phase is named “pacemaker depolarization”.
Given the large spectrum of heart rates observed in mammals the
duration of this phase can vary substantially, however the voltage
range encompassed is extremely constant and roughly extends from
−60 to −40 mV [1–3]. To sustain this phase several ionic currents and
pumps enter in action at variable times and voltages [4–6], and this
complexity allows for a highly flexible system since the chronotropic
fine tuning operated by neuro-hormonal regulators can target
different effectors.
In this review we will focus on the If current which is responsible
for initiating the diastolic depolarization of SAN cells. Due to its
fundamental role and its unusual characteristics of being activated in
hyperpolarization, this current was named “pacemaker current” or
“funny” (If) current [7–9]. The unique property of a reverse voltage
dependence, together with the inward nature of the current at
diastolic potentials, makes this current apt to initiate and support the
diastolic depolarization. In addition, the direct modulation of the
current operated by the second messenger cAMP, represents one of
the main pathways by which the autonomic nervous system controls
cardiac chronotropism [10]. Two recent clinical findings further
confirm the role of f-channels in setting the cardiac rate: one is the
evidence of a causative link between the presence of loss-of-function
mutations found in these channels and the arrhythmic state of
individuals carrying the mutations, and the other is the specific heart
rate reduction observed in patients treated with ivabradine, a drug
that at therapeutic doses selectively reduces the If current (see
specific sections in this review).
Although originally discovered in the heart, the If current is also
abundantly present in a large fraction of neuronal elements, where it
contributes to rhythmic firing, synaptic integration, and dendritic
integration [11].
2. Basic biophysical properties of the pacemaker current
The If current is carried by Na+ and K+ ions and its reversal
potential is between − 10 and −20 mV, and permeation occurs
according to a multiion, single-file mechanism [9,12,13]. Interestingly,
recent single channel experiments carried out in rat and human
working myocytes have also reported a weak permeability to Ca2+
ions [14]. Voltage-clamp studies have shown that activation is a
complex event determined initially by an intrinsic rearrangement of
the closed structure of the channel, which originates a “shoulder-like”
delay, and followed by the proper close-to-open transition which
proceeds according to an exponential relaxation [15,16]. The quantitative aspects of this activation process, i.e. the time-course of the
current activation, and the steady-state current level reached at the
end of each pulse, are finely controlled both by the membrane voltage
and by a direct interaction with the second messenger cAMP [17,18].
Although variable data have been reported, the voltage threshold for
sinoatrial If activation is compatible with a functional presence of the
current at diastolic potentials (Table 1 in [19]); for example, recent
experiments carried out on human SAN cells indicate that threshold
for If activation occurs between −50 and − 60 mV, a voltage range
well comprised in the diastolic range [3].
3. Molecular structure of pacemaker channels
The molecular constituents of pacemaker channels belong to the
hyperpolarization-activated cyclic nucleotide-gated (HCN) channel
family which is part of the superfamily of K+ channels. In mammals
four HCN isoforms are known (HCN1–4), and heterologous expression
and in vivo investigations have shown that they can assemble both as
homotetramers and heterotetramers (with the exceptions of HCN2–
HCN3 heteromers) to yield functional channels. Currents resulting
from homomeric and heteromeric assembly have biophysical and
modulatory properties qualitatively similar to the native If [20]; but
when quantitative aspects are taken into account noticeable differences are evident in terms of kinetics, voltage dependence, and cAMP
sensitivity [21–26]. Each HCN isoform is composed of three large
macrodomains: a cytoplasmic N-terminus, a 6 transmembranespanning “core” region, and a cytoplasmic C-terminus. Each of these
regions mediates important functional aspects. The N-terminus
contains a conserved stretch of residues which appears to be
important in mediating channel trafficking [27]. The core transmembrane region is the more conserved region among isoforms and
contains both the voltage sensor (S4 segment) and residues involved
in pore formation (S5-P-S6 regions), and therefore constitutes the
proper gating module of the channel. All HCN isoforms exhibit within
the permeation pathway the GYG triplet which, as in pure K+
channels, constitutes the selectivity filter. It is not immediately
apparent why HCN channels are also permeable to Na+ ions, but it
is now believed that the inner pore of HCN channels is somewhat less
rigid than that of K + channels, and for this reason it also
accommodates partially hydrated Na+ ions [28]. Further information
Table 1
Effects of “heart rate-lowering” agents on electrical properties of cardiac pacemaker cells.
If reduction
IK reduction
ICa reduction
DDS
APD50
Rate reduction
Refs
Alinidine
Zatebradine
(UL-FS49)
Cilobradine
ZD7288
Ivabradine
∼80%
30 μM
rbSAN cells
30%
30 μM
rbSAN cells
11%
30 μM
rbSAN cells
↓
30 μM
rbSAN cells
+ 23%
30 μM
rbSAN cells
22%
30 μM
rbSAN cells
[146]
∼65%
1 μM
rbSAN cells
∼20%
1 μM
rbSAN cells
No effect
1 μM
rbSAN cells
− 42%
3 μM
rbSAN tissue
+ 29%
3 μM
rbSAN tissue
28%
3 μM
rbSAN tissue
[147–149]
∼60%
1 μM
mSAN cells
∼22%
5 μM
mSAN cells
Not tested
78%
1 μM
gpSAN cells
No effect
1 μM
gpSAN cells
14%
1 μM
gpSAN cells
− 54%
1 μM
gpSAN tissue
+ 8%
1 μM
gpSAN tissue
∼35%
1 μM
gp right atrium
[150,151]
∼60%
3 μM
rbSAN cells
No effect
3 μM
rbSAN cells
No effect
3 μM
rbSAN cells
−67%
3 μM
rbSAN tissue
+ 9%
3 μM
rbSAN tissue
∼24%
3 μM
rbSAN tissue
[10,13,148,149]
− 70%
1 μM
mSAN cells
+ 60%⁎
1 μM
mSAN cells
∼55%
1 μM
mSAN cells
[139]
DDS: diastolic depolarization slope; APD50: action potential duration measured at 50% of repolarization (⁎APD measured from the threshold to the following maximum diastolic
potential); rb, rabbit; m, mouse; gp, guinea-pig.
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
on the structural organization of the pore comes from the observation
that the pore blocker ZD7288 directly interacts with residues of the S6
segment, and that closing of the channel entraps the drug into the
inner vestibule of the HCN1 channel [29]. Taken together, these data
and the homology model based on the KcsA K+ channel crystal [30],
suggest that the four S6 segments line the inner wall of HCN pore, and
that access to the S6 residues involved in ZD7288 binding is guarded
by a gating structure likely to comprise part of the S6 segment. The Cterminus region comprises three separate structural elements: the Clinker, which is organized in 6 α-helices, the cyclic nucleotide binding
domain (CNBD), and the proper C-terminus. Collectively, the CNBD
and the C-linker act as a functional unit that modulates the open
probability of a pacemaker channel. Detailed information on several
aspects of the HCN channel structure–function relation is now
available [21,31,32] and here we will only provide some general
concepts. It is now clear that voltage-dependent ion channels do not
behave like rigid structures and therefore the rotational/translational
movement of the S4 segments imposed by changes in the membrane
electrical field does not induce a general rearrangement of the protein
leading to the opening of the pore [33]. Due to its direct interaction
with the S4 segment a particular focus was placed on the role of the
S4–S5 linker, which indeed was shown to undergo a spatial
reorganization upon S4 movement [33,34]. Single alanine substitution
of aminoacids E324, Y331, and R339 of the HCN2 S4–S5 linker was able
to disrupt channel closing thus indicating that this linker plays a
crucial role in gating [34]. There is now evidence that the S4–S5 linker
functionally interacts with the C-linker region; for example in HCN2
channels the electrostatic interaction between R339 of the S4–S5
linker and D443 of C-linker stabilizes the closed state of the channel
[35]. These data indicate that the interaction between the S4–S5 linker
and the C-linker mediates the coupling of voltage sensing and channel
opening and closing. In addition to the voltage, the HCN channel open
probability is also controlled by the second messenger cAMP through
a direct binding to the channel [17]. Crystallization data indicate that
in the presence of cAMP the A′ and B′ helices of the C-linker of each
subunit interact with the C′ and D′ helices of the neighboring subunit,
and this tetrameric arrangement favors the open state of the channel.
In the absence of cAMP, the tetrameric assembly is lost and this favors
the closed channel conformation [21,36,37]. This mechanism thus
represents the final molecular event in the modulation of the If
current exerted by the autonomic nervous system.
4. Pacemaker channels during embryonic development
Pacemaker activity characterizes the developing heart at a very
early stage; for example, in chick embryos electrical pacemaker
activity is present even before the onset of regular contractions of the
primitive heart tube [38]. In mammals, the linear heart tube starts to
beat around embryonic day 8.5 (ED8.5), a stage at which all
cardiomyocytes are autorhythmic and express the If current [39].
With further development (ED18) only 33% of myocytes present
spontaneous but irregular action potentials and the If current is
decreased by 82% [39]. In agreement with functional data the same
authors also reported a significant down-regulation of HCN1 and
HCN4 mRNAs and a moderate increase of HCN2; HCN3 was not
detected [39]. A detailed analysis of HCN4 localization in the
developing heart, by in situ hybridization, indicates that the HCN4
mRNA is already detectable in the cardiac crescent at ED7.5, and its
expression remains confined to the venous pole, the cardiac region
from which develops the mature SAN [40]. Specific expression of
HCN4 mRNA in the SAN and conduction system has also been
demonstrated in a series of studies addressing the molecular pathway
leading to the formation of these regions. Tbx3 is a transcriptional
repressor whose expression during heart development specifically
delineates the SAN/conduction system and match the HCN4 expression [41,42]. Interestingly, ectopic expression of Tbx3 in the atria
57
generates foci of autorhythmic cells in which HCN4 results upregulated [43]. Also, constitutive deletion of the transcription factors
Shox2 causes embryonic lethality at mid-gestation (ED11.5–ED12.5),
and this genetic alteration prevents the expression of both HCN4 and
Tbx3 and causes formation of an underdeveloped SAN and significant
bradycardia [44]. Taken together these data indicate that HCN4 is a
major component of the f-channel in the developing SAN, but they
leave some uncertainty regarding its presence in the embryonic
working myocytes. The embryological lethality of HCN4 knockout
mice further indicates that the presence of the If current is necessary
for proper cardiac development [45]. Despite the observation that
Tbx3 and Shox2 are important in the modulation of HCN4 expression
during normal heart development, they do not account for basal HCN
gene transcription. Indeed, the lack of Tbx3 does not influence HCN4
expression [43], and the lack of Shox2 does not modify HCN4
expression until ED10 [44]. Two conserved sequences accounting for
HCN4 gene transcription have recently been found in the non-coding
region of HCN4 [46]. The first, an 846 bp long sequence is located at
the 5′UTR, the second sequence is a neuron-restrictive silencer
element (NRSE) located in the intronic region between exons 1 and
2, which specifically represses the promoter activity of the first
sequence [46].
A different approach in studying early events of cardiac development consists of the in vitro differentiation of embryonic stem cells
(ESC) through the formation of embryoid bodies (EBs), which are
three-dimensional cell aggregates able to recapitulate early embryonic developmental events. 7–8 days after cell aggregation mouse EBs
start to display foci of spontaneous contraction generated by early
cardiomyocytes [47–49]. The If current has been found both in human
and in mouse ESC-derived cardiomyocytes and its direct involvement
in the generation and modulation of rate has been assessed by
applying specific If inhibitors [48,50–54]. In mouse ESC-derived
cardiomyocytes, ivabradine (3 μM) slowed beating rate by 25% and
reduced the If current by 50%, while ZD7288 (0.3 μM) slowed beating
rate by approximately 50% and reduced the If current by 15% [52,53].
Similarly, zatebradine, another If blocker, slowed spontaneous rate of
human ESC-derived cardiomyocytes [50]. Patch-clamp investigations
have shown that the If current is already present in cells isolated from
10 day-old mouse EBs, and that its density increases significantly with
the progression of differentiation [52–54].
There is some variability concerning the mRNA and protein
expression of HCN isoforms in mouse ESC-derived myocytes. While
two studies reported the expression of mRNAs of all four isoforms
[52,55], another study showed that mouse ESC-derived myocytes only
express HCN1 and HCN4 [56]. At the protein level, western blot
experiments revealed the expression of HCN2 and HCN3 isoforms
[53], while immunofluorescence experiments reported that HCN1 and
HCN4 are the only isoforms expressed on the membrane of pacemaker
cells [52,57]. Investigations of human ESC-derived cardiomyocytes
revealed a high and stable expression of HCN2 mRNA while HCN1 and
HCN4 appear to be present in undifferentiated human ESC and their
expression decreases with differentiation [50,51].
Despite this variability, likely originating from differences in ES
clones, species and in the techniques employed, it is clear that the If
current and the underlying HCN channels are expressed very early
during cardiac development and significantly contribute to the
pacemaking mechanism.
5. Pacemaker channels in the adult heart
The If current has been described in all tissues of the heart,
however its functional contribution in non pathological conditions is
limited to the cardiac conduction system. Here we provide an
overview of the characteristics of the If current in various heart
regions.
58
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
Fig. 1. Molecular and functional properties of SAN myocytes. (A) Spontaneous action potentials (left) and If current traces (right) recorded from typical rabbit SAN myocytes; currents
were elicited by hyperpolarizing voltage steps in the range − 45 to − 75 mV. (B) Immunofluorescence analysis of rabbit SAN tissue slice labelled with anti-connexin 43 (Cx43, red)
and anti-HCN4 (green) antibodies. HCN4 is strongly expressed in the central region of the SAN, while the opposite staining is observed for Cx43; crista terminalis (CT), interatrial
septum (IS). (C) HCN4 labelling of single myocytes isolated from CT, SAN and IS (top), and representative current traces recorded at − 125 mV from myocytes isolated from the same
regions (bottom). Both If current density and HCN4 labelling are more abundant in the central nodal area. (Panels B and C from [61] with permission).
5.1. SAN
In agreement with the leading role of the SAN in pacing, this tissue
presents the highest level of If current density and HCN expression
[58–62]. As previously mentioned, the presence of the If current has
been documented in adult SAN myocytes of many species including
humans [3,19], and its activation threshold (∼− 50.8 mV, n = 9, from
Table 1 in [19]) falls within the voltage excursion encompassed by SAN
cells during the diastolic depolarization. In SAN tissues of lower
mammals and humans, the predominant molecular constituent of fchannel is the HCN4 isoform (Fig. 1); HCN1 and HCN2 have also been
detected, but at low to moderate levels depending on the species [58–
66]. HCN3 is absent from the SAN [58,67].
5.2. Atrioventricular node (AVN)
The AVN is able to pace the heart in the absence of proper sinus
rhythm, and the If current has been recorded in the majority of cells
isolated from the AVN of rabbit, mouse and guinea-pig [68–70]. A
more detailed investigation on ovoid- and rod-shaped cells isolated
from rabbit AVN indicates that in ovoid cells the pacemaker current is
abundantly present (− 5.18 pA/pF at − 100 mV) and its activation
threshold is about −60 mV, while in rod-shaped cells the current is
nearly absent [68]. Studies on mRNA distribution indicate that in the
AVN, as well as in the SAN, the HCN4 isoform is the most abundant
isoform [71], and, at least in mice, the overall HCN4 protein expression
is about one third of that found in the SA node [72].
5.3. Purkinje fibres (PFs)
Isolated PFs are able to beat spontaneously, and for this reason they
were largely used in early electrophysiological experiments investigating the nature of the pacemaker currents responsible for the
diastolic depolarization [8,9,73,74]. The diastolic depolarization of
these fibres develops between −90 mV and − 70 mV, thus in a range
of potential more negative than that of a typical SAN cell. Interestingly,
the threshold for If current activation in these cells is also negatively
shifted (∼ −80 mV), and this suggests an active role of the current in
the generation of the spontaneous activity [8,9]. The molecular
composition of f-channels in PFs is extremely dependent on the
species investigated. For example, canine PFs, which exhibit a
significant automaticity and a large If current, have high levels of
HCN mRNA (∼ 35% of the HCN signal recorded in the SAN), with ∼90%
of the transcripts constituted by HCN4 and the remaining by HCN2
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
[75,76]; interestingly, at the protein level HCN2 is the major isoform
expressed [76]. In contrast, rabbit PFs, which tend not to be automatic
and exhibit little If current, show minimal levels of HCN messenger
(∼4% of SAN) with an equivalent presence of HCN4 and HCN1, and a
minor contribution of HCN2 (∼10% [65]). In human PFs HCN4 is the
predominant isoform expressed [77].
5.4. Working myocardium
The presence of the If current in the working myocardium is well
documented but its range of activation is much more negative than in
the SAN. Reported values of thresholds are extremely variable and
range from −60 to − 120 mV [78–83]. Furthermore, even when
assessed at −130 mV the current densities are extremely low; for
example in human atrial and ventricular myocytes the reported values
are −0.8 pA/pF and −0.47 pA/pF, respectively [79,81]. For comparison at − 130 mV the If density measured in human SAN cells is
− 8 pA/pF [3]. Taken together these data indicate that under
physiological conditions the If current is not expected to play a
functional role in atria and ventricles. The molecular composition of fchannels in working myocytes is extremely complex due to interspecies variability and sometimes due to contrasting results between
transcriptional and protein data. In general, it can be stated that the
major protein isoform both in atria and ventricles is HCN2. HCN4 is
only occasionally reported, while HCN1 does not appear to be present
[62,71,76,84,85]. Despite the fact that pacemaker currents do not play
a role in the healthy atria and ventricles, growing evidence in humans
and animals demonstrates that the over-expression of the If current in
these tissues is often associated to some cardiac diseases, and this
over-expression may represent an important arrhythmogenic source
[82,83,85–90]. For example, in human ventricular myocytes isolated
from failing hearts of patients with ischemic cardiomyopathy, the If
current is over-expressed by approximately two-fold and its threshold
of activation is shifted positively by 9.6 mV [82]. In agreement with
these observations up-regulation of HCN2 and HCN4 mRNAs/proteins
was found in both atria and ventricle myocytes obtained from failing
hearts explanted from patients with end-stage ischemic cardiomyopathy [85]. A regulatory mechanism controlling the expression of HCN
isoforms may consist of post-translational events. One such mechanism, known to control the functional expression of both HCN2 and
HCN4 is operated by two muscle-specific micro RNAs, miR-133 and
miR-1 [91,92]. For example, during cardiac development, up-regulation of miR-133 and miR-1 decreases the ventricular If current by
reducing HCN2 and HCN4 proteins without significantly affecting
mRNA levels. The opposite event occurs in hypertrophied hearts
where miR-133 and miR-1 are decreased and this down-regulation
causes the re-expression of HCN2 and HCN4 channels [91,92]. This
same mechanism could also explain the variability between mRNA
and protein expression data found in different systems.
6. Basis of functional heterogeneity of pacemaker currents
The identification of the exact molecular composition of native
pacemaker channels was initially challenged by the attempt to
reproduce the native kinetic and modulatory properties of the native
f through heterologous expressions of HCN isoforms. Unfortunately,
this approach proved unsuccessful. Although HCN4 and HCN1 are the
main isoforms of the SAN, their expression either alone or in
combination failed to reproduce the sinoatrial If current. For example,
even though homomeric HCN4 channels retain a cAMP modulation
similar to that observed in native channels, both the activation and
deactivation kinetics and the voltage dependence are much different.
On the contrary HCN4–HCN1 heteromers generate currents with
kinetics approaching those of native SAN f-channels, but they do not
reproduce the same voltage dependence of activation [25]. Furthermore, it has been shown that even when the same isoform is
59
expressed in different cells, the resulting currents do not display
identical properties. To this regard Qu et al. [93] have shown that
currents resulting from HCN2 expression in neonatal and in adult
ventricle myocytes have different biophysical properties such as a
different position of the voltage-dependent curve (V1/2, − 76 mV and
−96 mV, respectively). These data clearly suggest that, in addition to
the heterogeneity provided by different HCN isoform and cAMP
modulation, the large phenotypic variation of If properties observed
both in the healthy heart and in pathological conditions likely reflects
the presence of additional modulatory factors. Recent evidence
indicates that protein–protein and protein–phospholipid interactions,
and modulatory cytoplasmic factors play a relevant role in modulating
the pacemaker current. Taken together all these regulatory elements
are defined as “context dependence”. Some of the most relevant
modulatory factors are MiRP1, phosphatidylinositol 4,5-bisphosphate
(PI(4,5)P2), and caveolin 3.
6.1. MiRP1
Mink-related protein 1 (MiRP1 or KCNE2) is a single transmembrane-spanning protein that acts as a β-subunit of HCN channels.
Although its modulatory actions depend on the HCN isoform and on
the expression system used, the most relevant effect is to increase
current density [94–97].
6.2. PI(4,5)P2
Recent data indicate that the voltage dependence of native
sinoatrial f- and HCN1/2 channels is also regulated by local pools of
PI(4,5)P2 [98,99]. Further experiments suggested that stimulation of
receptors coupled to phospholipase C (PLC), such as bradykinin BK2
receptor and the muscarinic M1 receptor, can modulate the gating of
both recombinant and native HCN channels via an increase of PI(4,5)
P2 [100]. Specifically, activation of the BK2 receptor induces a positive
shift of the voltage dependence of HCN2 (∼ 20 mV), HCN1 (∼ 6 mV),
and sinoatrial f-channels (∼8 mV), and also affects time constants of
activation and deactivation. Noteworthy is the fact that basal levels of
PI(4,5)P2 may change with development, stress, and pathological
conditions [101–103].
6.3. Caveolin 3
It has also been shown that in rabbit SAN cells HCN4 localizes in
caveolin-rich membrane microdomains (caveolae) and interacts with
caveolin 3 [104,105]. Disorganization of caveolae strongly affects fchannel kinetics, shifting the activation curve toward more positive
potentials and slowing the deactivation kinetics. Similar effects were
observed for HCN4 current after lipid raft disruption in HEK cells
[106].
Post translational modification events may also contribute to
modulate HCN properties: both native and HCN channels can indeed
be modulated by phosphorylation processes and particularly by
tyrosine-kinase phosphorylation, but a clear understanding of the
functional role of this regulatory mechanism remains in part elusive
[107–115].
7. Genetics of HCN: HCN knockout mice and HCN-associated
pathologies in humans
Genetically manipulated knockout mice and genetic analysis in
human patients affected by severe or mild cardiac rhythm alterations
have brought new information on the role and relevance of HCN4
channels to cardiac pacemaking. In recent years four HCN4 knockout
mouse models have been developed to evaluate the functional
contribution of this isoform to pacemaking in vivo [45,116,117]. Both
global and cardiac specific constitutive knockouts determine the
60
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
premature death of mice at mid-gestation (ED9.5–ED11.5), period in
which a regular contractile activity of the developing heart normally
appears [45]. Hearts isolated from knockout embryos at day ED9.5
exhibited a decreased spontaneous rate (−36.7%) and If current
(−75/−90%). Interestingly, the knockout process also led to the
complete loss of cAMP-mediated β-adrenergic chronotropic modulation in both intact hearts and single cells; this finding is consistent
with the observation that the residual If current could be carried by
the HCN1 and HCN3 isoforms which are known to be mildly or not
affected by cAMP modulation [67,118,119]. Taken together these data
indicate that expression of HCN4 is a necessary element for proper
embryonic heart development, but unfortunately they do not provide
any hint on the role of the HCN4 current in the adult animal. In order
to overcome this limitation, temporally-controlled global [116], and
cardiac specific [117] HCN4 knockout models were developed. ECG
recordings in freely moving knockout animals showed the presence of
sinus pauses (mean duration = 321 ms, mean frequency = 8.1/min) in
otherwise normal rhythmic activity. Interestingly, β-modulation of
rate was not lost in the adult knockout animals, and any acceleration
in heart rate led to a reduction of sinus pauses. When the pacemaking
activity was investigated at the single cell level, it was found that 90%
of sinoatrial node myocytes of the global knockout and 50% of the
cardiac specific knockout were quiescent [116,117], and that the If
current was reduced by about 75%–80%. Interestingly, β-adrenergic
stimulation was able to rescue spontaneous activity in knockout cells
[116]. As pointed out by the same authors, the presence in vivo of a
sympathetic tone and of additional network properties may explain
the persistence of spontaneous activity in vivo, however it does not
fully explain the presence of sinus pauses [116].
The members of the HCN channel family were among the last ion
channel genes to be cloned; for this reason research aiming at
identifying a link between alterations in the HCN genes and cardiac
pathologies has only recently begun. During the course of independent screenings of patients affected by various forms of rhythm
disturbances, four mutations/deletions of the HCN4 open reading
frame have been identified and directly correlated to the phenotypes
[120–123]. Although all the identified mutations determine a reduced
contribution of the pacemaker If current to the diastolic depolarization
(loss-of-function), the functional mechanisms by which this reduction is achieved are quite different (Fig. 2).
The first evidence of a mutation of the HCN4 in humans was found
by Schulze-Bahr et al. [120] who reported the presence of the
heterozygous 1-bp deletion (1631delC) in a single patient affected by
idiopathic sinus node dysfunction (SND) with severe bradycardia
(41 beats per minute, bpm), intermittent atrial fibrillation, and
chronotropic incompetence. The resulting HCN4 protein (573X) had
a shorter C-terminus, lacking the CNBD, caused by the presence of an
earlier stop codon. In vitro heterologous expression of the 573X
channels revealed that the major alteration in the current was a
dominant negative loss of cAMP modulation, which could explain the
chronotropic incompetence of the patient both at rest and during
maximal workload [120]. Although these data are extremely intriguing, the causative relation between the altered function of the
channel and the pathological phenotype of the patient is not
conclusive since the study is based on a single-case report. In another
study Ueda et al. [121] reported the case of arrhythmic patients
affected by SND with recurrent syncope, severe bradycardia, QT
prolongation, polymorphic ventricular tachycardia. The study, based
on a two generation family, suggested a possible linkage between the
presence of the D553N mutation in the C-linker region and the disease
since members carrying the mutation (n = 3) also presented the
clinical phenotype. Patch-clamp and immunofluorescence analysis
showed that this mutation exerts a dominant negative effect and
severely impairs channel trafficking to the membrane leading to an
almost complete loss of the pacemaker current. Despite the strong
suggestion of a link between the mutation and the disease, the
Fig. 2. hHCN4 mutations associate to cardiac rhythm alterations. Schematic topology of
a single HCN4 subunit with the 6 transmembrane segments, and the N- and C-termini.
The C-terminal region is composed by the C-linker (A′ to F′ helices), the CNBD (A helix,
β-roll, B and C helices) and the proper C-terminus. Orange dots indicate the position of
the point mutations (G480R, D553N, and S672R), while the orange X indicates the site
where the mutant protein (573X), originated by the 1631delC, is truncated. Insets show
the main effects on If properties induced by the mutations. Top insets show bar-graphs
indicating the reduction in current densities observed for mutant G480R (at −100 mV)
and D553N (at − 120 mV) proteins (measured by eye from [121,123]). Bottom left inset
shows that cAMP does not modulate the position of the voltage-dependent curve of the
mutant 573X (reproduced from [120]), while bottom right inset shows that S672R
mutant channels display a negative shift of the voltage-dependent curve (reproduced
from [122]). For clarity data relative to homomeric channel expressions have been used.
F.A., fractional activation.
complexity of the clinical manifestations cannot be easily explained by
in vitro results.
A direct correlation between a heterozygous mutation and cardiac
rhythm alteration has been demonstrated for the S672R mutation
which induces a form of familial asymptomatic bradycardia [122]. An
extensive familial analysis carried out on 27 members allowed the
conclusion that the mutation co-segregated with the bradycardic
phenotype according to an autosomal dominant pattern (LOD score
N5). The mean heart rate of affected individuals (52.2 bpm) was
reduced by 29% when compared to unaffected members (73.2 bpm) of
the same family. In vitro heterologous expression of mutant and wild
type channels demonstrated that the half-activation voltage (V1/2) of
homomeric mutant channels was 8.4 mV (heteromeric channels:
4.9 mV) more negative than that of wild type channels; a slower
kinetics of deactivation was also observed. Despite the fact that this
mutation is localized in the CNBD region, cAMP retained its normal
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
modulatory action. Interestingly, the quantitative and qualitative
effects of the mutation (4.9 mV shift and 29% heart rate reduction)
completely resemble those of a low dose (30 nM) of acetylcholine.
This study not only identifies a tight linkage between the mutation
and bradycardia but it is also the first one that fully describes the
underlying molecular mechanism.
A form of familial sinus bradycardia was also found to co-segregate
with the mutation G480R which is located in the pore region of the
HCN4 channel [123]. In vitro studies seem to indicate both an extreme
negative shift (∼ 40 mV) of the voltage dependence and trafficking
defects.
Taken together studies on HCN4 channelopathies seem to suggest a
tight relation between the amount of current suppressed by the
presence of the mutation and the severity of the associated disease.
Indeed, while the limited reduction of the If current caused by the
S672R mutation only determines asymptomatic bradycardia, progressively more severe clinical manifestations appear with the deletion
1631delC which eliminates the cAMP modulation, and with the D553N
mutation which suppresses the presence of functional channels.
8. Biological pacemaker
The search of new therapeutic tools consisting of gene- and/or
cell-based intervention aimed to restore compromised cardiac
functions has prompted researchers to exploit the use of HCN
channels to alter cellular electrical activity in order to generate, in
normally quiescent substrates, stable rhythmic activity similar to that
of native pacemaker myocytes. The specific features of pacemaker
channels and in particular the fact that they are activated only at
diastolic potentials and do not contribute to other phases of the action
potentials, make them particularly suitable for such purpose. Early in
vitro studies demonstrated that virus-mediated over-expression of
HCN2 channels induced a significant increase in the rate of
spontaneously beating neonatal ventricular myocytes by causing an
If-mediated increase of the diastolic depolarization slope [93]. This
approach was later confirmed in vivo by showing that direct injection
of the HCN2-adenovirus in the left atrium or into the ventricular
conduction system of dogs, was able to induce ectopic regular
spontaneous activity after AV block [93,124–126]. Similarly, adenovirus-mediated over-expression of HCN1 or HCN4 was sufficient to
induce a regular rhythm in quiescent cardiomyocytes [127,128].
Alternative cell-based strategies, aimed to avoid the use of viruses,
have been developed by engineering cells in order to express high
levels of HCN channels. Engineered human mesenchymal stem cells
(hMSCs) expressing either HCN2 or HCN4 have been shown in vitro to
properly connect to neonatal cardiomyocytes and to increase their
intrinsic spontaneous rhythm [129,130]. HCN2-expressing hMSCs
have also been successfully transplanted in canine left ventricular
wall where they were able to induce stable ectopic beats [129].
Furthermore, spontaneously beating heterokarion cells generated by
the fusion of HCN1 expressing lung fibroblast and ventricular
myocytes were shown to induce in vivo ectopic beats at the site of
injection [131].
A different approach to generate a biological pacemaker would be
to use a cellular substrate as close as possible to native pacemaker
myocytes. Both murine and human embryonic stem cells differentiate
into spontaneously beating cell aggregates (EBs) which contain
myocytes with functional and molecular properties typical of pacemaker cells [50,52–54,132–135]. Two separate studies have demonstrated that spontaneously beating portions of human EBs are able to
pace either cultures of neonatal rat cardiac myocytes in vitro or the
whole heart in vivo [136,137]. Although neither the If current nor HCN
expression was directly addressed in these studies, the increase in rate
by β-adrenergic stimulation and, more importantly, the decrease by
the f-channel blocker ZD7288 [136,137] strongly suggest a role of the
pacemaker current in the generation of the rhythmic activity.
61
9. f-channels blockers
The specific and restricted contribution of If to the generation and
modulation of the sinoatrial diastolic depolarization phase has long
made this current a crucial target for pharmacological applications. In
principle, a selective reduction of the If current should cause a slowing of
heart rate devoid of undesired side effects. For this reason, drugs able to
block f-channels are expected to have the potential for treatment of
heart diseases characterized by a deficiency of oxygen supply to the
working myocardium such as angina pectoris, heart failure and ischemic
heart disease. Several f-channel blockers, called specific “heart ratelowering” agents, have been developed and extensively characterized in
in vitro and in vivo studies. This family includes alinidine (ST567),
zatebradine (UL-FS49), cilobradine (DK-AH26), ZD-7288, and ivabradine (S16257) [19]. The effects of these agents on the action potential
parameters and membrane currents of pacemaker cells/tissue are
shown in Table 1. The only If blocker extremely selective at therapeutic
doses and with mild side effects is ivabradine [19]. The action of
ivabradine on native f-channels displays both a marked use-dependence, since the drug can only access its binding site when channels are
in the open state, and current-dependence since the direction of current
through the channel affects the stability of the binding [13]. The “usedependence” is the mechanism responsible for block accumulation
during repetitive channel opening–closing cycles with the possible
consequence that the higher the initial cardiac rate the more effective is
the ivabradine block [66]. Experiments investigating the affinity of
ivabradine for homomeric HCN channels do not reveal any substantial
isoform-specificity although the drug was shown to act with a higher
degree of co-operativity in binding to HCN1 than to HCN4 and native fchannels [138,139]. A striking difference however exists between HCN4
and HCN1 in the state-dependence of channel block: while ivabradine
behaves as an open-channel blocker of HCN4 and native channels, it can
only block HCN1 channels when they are in the closed state [138].
Currently, ivabradine is marketed for treatment of chronic stable angina
in patients with normal sinus rhythm who have a contraindication or
intolerance to β-blockers; clinical studies of patients with chronic stable
angina have shown that ivabradine acts as a pure heart rate-reducing
agent and has anti-ischemic and anti-anginal properties equivalent to βblockers and Ca2+ channel blockers and presents a good safety and
tolerability profile even during long-term treatment [140–143]. Mild
visual symptoms (phosphenes) were occasionally reported, but were
generally well tolerated [140,143]. Additional information comes from
results from a recent large clinical trial (BEAUTIFUL) which indicate that
ivabradine treatment of patients with stable coronary artery disease
(CAD) and heart rate ≥70 bpm can reduce the incidence of some CAD
outcomes such as hospitalization for myocardial infarction and coronary
revascularization [144,145].
Acknowledgments
This work was supported by grants from the Ministero dell'
Istruzione dell'Università e della Ricerca (Cofin 2007WB35CW) to MB
and by the European Union (Normacor) grant. We would like to thank
Prof. DiFrancesco for his helpful suggestions and discussions.
References
[1] Mangoni ME, Nargeot J. Properties of the hyperpolarization-activated current (I
(f)) in isolated mouse sino-atrial cells. Cardiovasc Res 2001;52:51–64.
[2] DiFrancesco D, Ferroni A, Mazzanti M, Tromba C. Properties of the hyperpolarizing-activated current (if) in cells isolated from the rabbit sino-atrial node. J
Physiol 1986;377:61–88.
[3] Verkerk AO, Wilders R, van Borren MM, Peters RJ, Broekhuis E, Lam K, et al.
Pacemaker current (I(f)) in the human sinoatrial node. Eur Heart J 2007;28:
2472–8.
[4] Mangoni ME, Nargeot J. Genesis and regulation of the heart automaticity. Physiol
Rev 2008;88:919–82.
[5] Lakatta EG, DiFrancesco D. What keeps us thicking, a funny current, a calcium
clock, or both? J Mol Cell Cardiol 2009;47:157–70.
62
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
[6] DiFrancesco D. Pacemaker mechanisms in cardiac tissue. Annu Rev Physiol
1993;55:455–72.
[7] Brown HF, DiFrancesco D, Noble SJ. How does adrenaline accelerate the heart?
Nature 1979;280:235–6.
[8] DiFrancesco D. A new interpretation of the pace-maker current in calf Purkinje
fibres. J Physiol 1981;314:359–76.
[9] DiFrancesco D. A study of the ionic nature of the pace-maker current in calf
Purkinje fibres. J Physiol 1981;314:377–93.
[10] Bucchi A, Baruscotti M, Robinson RB, DiFrancesco D. Modulation of rate by
autonomic agonists in SAN cells involves changes in diastolic depolarization and
the pacemaker current. J Mol Cell Cardiol 2007;43:39–48.
[11] Wahl-Schott C, Biel M. HCN channels: structure, cellular regulation and
physiological function. Cell Mol Life Sci 2009;66:470–94.
[12] Frace AM, Maruoka F, Noma A. External K+ increases Na+ conductance of the
hyperpolarization-activated current in rabbit cardiac pacemaker cells. Pflugers
Arch 1992;421:97–9.
[13] Bucchi A, Baruscotti M, DiFrancesco D. Current-dependent block of rabbit sinoatrial node I(f) channels by ivabradine. J Gen Physiol 2002;120:1–13.
[14] Michels G, Brandt MC, Zagidullin N, Khan IF, Larbig R, van Aaken S, et al. Direct
evidence for calcium conductance of hyperpolarization-activated cyclic nucleotide-gated channels and human native If at physiological calcium concentrations.
Cardiovasc Res 2008;78:466–75.
[15] DiFrancesco D, Ferroni A. Delayed activation of the cardiac pacemaker current
and its dependence on conditioning pre-hyperpolarizations. Pflugers Arch
1983;396:265–7.
[16] DiFrancesco D. Characterization of the pace-maker current kinetics in calf
Purkinje fibres. J Physiol 1984;348:341–67.
[17] DiFrancesco D, Tortora P. Direct activation of cardiac pacemaker channels by
intracellular cyclic AMP. Nature 1991;351:145–7.
[18] DiFrancesco D, Mangoni M. Modulation of single hyperpolarization-activated
channels (i(f)) by cAMP in the rabbit sino-atrial node. J Physiol 1994;474:473–82.
[19] Baruscotti M, Bucchi A, DiFrancesco D. Physiology and pharmacology of the
cardiac pacemaker (“funny”) current. Pharmacol Ther 2005;107:59–79.
[20] Accili EA, Proenza C, Baruscotti M, DiFrancesco D. From funny current to HCN
channels: 20 years of excitation. News Physiol Sci 2002;17:32–7.
[21] Zagotta WN, Olivier NB, Black KD, Young EC, Olson R, Gouaux E. Structural basis
for modulation and agonist specificity of HCN pacemaker channels. Nature
2003;425:200–5.
[22] Chen S, Wang J, Siegelbaum SA. Properties of hyperpolarization-activated
pacemaker current defined by coassembly of HCN1 and HCN2 subunits and
basal modulation by cyclic nucleotide. J Gen Physiol 2001;117:491–504.
[23] Ulens C, Tytgat J. Functional heteromerization of HCN1 and HCN2 pacemaker
channels. J Biol Chem 2001;276:6069–72.
[24] Much B, Wahl-Schott C, Zong X, Schneider A, Baumann L, Moosmang S, et al. Role
of subunit heteromerization and N-linked glycosylation in the formation of
functional hyperpolarization-activated cyclic nucleotide-gated channels. J Biol
Chem 2003;278:43781–6.
[25] Altomare C, Terragni B, Brioschi C, Milanesi R, Pagliuca C, Viscomi C, et al.
Heteromeric HCN1–HCN4 channels: a comparison with native pacemaker
channels from the rabbit sinoatrial node. J Physiol 2003;549:347–59.
[26] Moroni A, Barbuti A, Altomare C, Viscomi C, Morgan J, Baruscotti M, et al. Kinetic
and ionic properties of the human HCN2 pacemaker channel. Pflugers Arch
2000;439:618–26.
[27] Tran N, Proenza C, Macri V, Petigara F, Sloan E, Samler S, et al. A conserved domain
in the NH2 terminus important for assembly and functional expression of
pacemaker channels. J Biol Chem 2002;277:43588–92.
[28] Giorgetti A, Carloni P, Mistrik P, Torre V. A homology model of the pore region of
HCN channels. Biophys J 2005;89:932–44.
[29] Shin KS, Rothberg BS, Yellen G. Blocker state dependence and trapping in
hyperpolarization-activated cation channels: evidence for an intracellular
activation gate. J Gen Physiol 2001;117:91–101.
[30] Doyle DA, Morais CJ, Pfuetzner RA, Kuo A, Gulbis JM, Cohen SL, et al. The structure
of the potassium channel: molecular basis of K+ conduction and selectivity.
Science 1998;280:69–77.
[31] Craven KB, Zagotta WN. CNG and HCN channels: two peas, one pod. Annu Rev
Physiol 2006;68:375–401.
[32] Zhou L, Siegelbaum SA. Gating of HCN channels by cyclic nucleotides: residue
contacts that underlie ligand binding, selectivity, and efficacy. Structure 2007;15:
655–70.
[33] Long SB, Campbell EB, MacKinnon R. Voltage sensor of Kv1.2: structural basis of
electromechanical coupling. Science 2005;309:903–8.
[34] Chen J, Mitcheson JS, Tristani-Firouzi M, Lin M, Sanguinetti MC. The S4–S5 linker
couples voltage sensing and activation of pacemaker channels. Proc Natl Acad Sci
U S A 2001;98:11277–82.
[35] Decher N, Chen J, Sanguinetti MC. Voltage-dependent gating of hyperpolarization-activated, cyclic nucleotide-gated pacemaker channels: molecular coupling
between the S4–S5 and C-linkers. J Biol Chem 2004;279:13859–65.
[36] Wainger BJ, DeGennaro M, Santoro B, Siegelbaum SA, Tibbs GR. Molecular mechanism
of cAMP modulation of HCN pacemaker channels. Nature 2001;411:805–10.
[37] Craven KB, Zagotta WN. Salt bridges and gating in the COOH-terminal region of
HCN2 and CNGA1 channels. J Gen Physiol 2004;124:663–77.
[38] Van Mierop LH. Location of pacemaker in chick embryo heart at the time of
initiation of heartbeat. Am J Physiol 1967;212:407–15.
[39] Yasui K, Liu W, Opthof T, Kada K, Lee JK, Kamiya K, et al. I(f) current and
spontaneous activity in mouse embryonic ventricular myocytes. Circ Res
2001;88:536–42.
[40] Garcia-Frigola C, Shi Y, Evans SM. Expression of the hyperpolarization-activated
cyclic nucleotide-gated cation channel HCN4 during mouse heart development.
Gene Expr Patterns 2003;3:777–83.
[41] Hoogaars WM, Tessari A, Moorman AF, de Boer PA, Hagoort J, Soufan AT, et al. The
transcriptional repressor Tbx3 delineates the developing central conduction
system of the heart. Cardiovasc Res 2004;62:489–99.
[42] Remme CA, Verkerk AO, Hoogaars WM, Aanhaanen WT, Scicluna BP, Annink C, et al.
The cardiac sodium channel displays differential distribution in the conduction
system and transmural heterogeneity in the murine ventricular myocardium. Basic
Res Cardiol 2009 PMID: 19255801 [Electronic publication ahead of print].
[43] Hoogaars WM, Engel A, Brons JF, Verkerk AO, de Lange FJ, Wong LY, et al. Tbx3
controls the sinoatrial node gene program and imposes pacemaker function on
the atria. Genes Dev 2007;21:1098–112.
[44] Espinoza-Lewis RA, Yu L, He F, Liu H, Tang R, Shi J, et al. Shox2 is essential for the
differentiation of cardiac pacemaker cells by repressing Nkx2-5. Dev Biol
2009;327:376–85.
[45] Stieber J, Herrmann S, Feil S, Loster J, Feil R, Biel M, et al. The hyperpolarizationactivated channel HCN4 is required for the generation of pacemaker action
potentials in the embryonic heart. Proc Natl Acad Sci U S A 2003;100:15235–40.
[46] Kuratomi S, Kuratomi A, Kuwahara K, Ishii TM, Nakao K, Saito Y, et al. NRSF
regulates the developmental and hypertrophic changes of HCN4 transcription in
rat cardiac myocytes. Biochem Biophys Res Commun 2007;353:67–73.
[47] Maltsev VA, Rohwedel J, Hescheler J, Wobus AM. Embryonic stem cells
differentiate in vitro into cardiomyocytes representing sinusnodal, atrial and
ventricular cell types. Mech Dev 1993;44:41–50.
[48] Maltsev VA, Wobus AM, Rohwedel J, Bader M, Hescheler J. Cardiomyocytes
differentiated in vitro from embryonic stem cells developmentally express
cardiac-specific genes and ionic currents. Circ Res 1994;75:233–44.
[49] Hescheler J, Fleischmann BK, Lentini S, Maltsev VA, Rohwedel J, Wobus AM, et al.
Embryonic stem cells: a model to study structural and functional properties in
cardiomyogenesis. Cardiovasc Res 1997;36:149–62.
[50] Sartiani L, Bettiol E, Stillitano F, Mugelli A, Cerbai E, Jaconi ME. Developmental
changes in cardiomyocytes differentiated from human embryonic stem cells: a
molecular and electrophysiological approach. Stem Cells 2007;25:1136–44.
[51] Satin J, Kehat I, Caspi O, Huber I, Arbel G, Itzhaki I, et al. Mechanism of
spontaneous excitability in human embryonic stem cell derived cardiomyocytes. J
Physiol 2004;559:479–96.
[52] Barbuti A, Crespi A, Capilupo D, Mazzocchi N, Baruscotti M, DiFrancesco D.
Molecular composition and functional properties of f-channels in murine
embryonic stem cell-derived pacemaker cells. J Mol Cell Cardiol 2009;46:343–51.
[53] Qu Y, Whitaker GM, Hove-Madsen L, Tibbits GF, Accili EA. Hyperpolarizationactivated cyclic nucleotide-modulated ‘HCN’ channels confer regular and faster
rhythmicity to beating mouse embryonic stem cells. J Physiol 2008;586:701–16.
[54] Abi-Gerges N, Ji GJ, Lu ZJ, Fischmeister R, Hescheler J, Fleischmann BK. Functional
expression and regulation of the hyperpolarization activated non-selective cation
current in embryonic stem cell-derived cardiomyocytes. J Physiol 2000;523:
377–89.
[55] White SM, Claycomb WC. Embryonic stem cells form an organized, functional
cardiac conduction system in vitro. Am J Physiol, Heart Circ Physiol 2005;288:
H670–9.
[56] van Kempen M, van Ginneken A, de Grijs I, Mutsaers N, Opthof T, Jongsma H, et al.
Expression of the electrophysiological system during murine embryonic stem cell
cardiac differentiation. Cell Physiol Biochem 2003;13:263–70.
[57] Yanagi K, Takano M, Narazaki G, Uosaki H, Hoshino T, Ishii T, et al. Hyperpolarization-activated cyclic nucleotide-gated channels and T-type calcium channels
confer automaticity of embryonic stem cell-derived cardiomyocytes. Stem Cells
2007;25:2712–9.
[58] Liu J, Dobrzynski H, Yanni J, Boyett MR, Lei M. Organisation of the mouse
sinoatrial node: structure and expression of HCN channels. Cardiovasc Res
2007;73:729–38.
[59] Marionneau C, Couette B, Liu J, Li H, Mangoni ME, Nargeot J, et al. Specific pattern
of ionic channel gene expression associated with pacemaker activity in the
mouse heart. J Physiol 2005;562:223–34.
[60] Tellez JO, Dobrzynski H, Greener ID, Graham GM, Laing E, Honjo H, et al.
Differential expression of ion channel transcripts in atrial muscle and sinoatrial
node in rabbit. Circ Res 2006;99:1384–93.
[61] Brioschi C, Micheloni S, Tellez JO, Pisoni G, Longhi R, Moroni P, et al. Distribution
of the pacemaker HCN4 channel mRNA and protein in the rabbit sinoatrial node. J
Mol Cell Cardiol 2009;47:221–7.
[62] Chandler NJ, Greener ID, Tellez JO, Inada S, Musa H, Molenaar P, et al. Molecular
architecture of the human sinus node: insights into the function of the cardiac
pacemaker. Circulation 2009;119:1562–75.
[63] Moosmang S, Stieber J, Zong X, Biel M, Hofmann F, Ludwig A. Cellular expression
and functional characterization of four hyperpolarization-activated pacemaker
channels in cardiac and neuronal tissues. Eur J Biochem 2001;268:1646–52.
[64] Ishii TM, Takano M, Xie LH, Noma A, Ohmori H. Molecular characterization of the
hyperpolarization-activated cation channel in rabbit heart sinoatrial node. J Biol
Chem 1999;274:12835–9.
[65] Shi W, Wymore R, Yu H, Wu J, Wymore RT, Pan Z, et al. Distribution and
prevalence of hyperpolarization-activated cation channel (HCN) mRNA expression in cardiac tissues. Circ Res 1999;85:e1–6.
[66] Thollon C, Bedut S, Villeneuve N, Coge F, Piffard L, Guillaumin JP, et al. Usedependent inhibition of hHCN4 by ivabradine and relationship with reduction in
pacemaker activity. Br J Pharmacol 2007;150:37–46.
[67] Stieber J, Stockl G, Herrmann S, Hassfurth B, Hofmann F. Functional expression of
the human HCN3 channel. J Biol Chem 2005;280:34635–43.
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
[68] Munk AA, Adjemian RA, Zhao J, Ogbaghebriel A, Shrier A. Electrophysiological
properties of morphologically distinct cells isolated from the rabbit atrioventricular node. J Physiol 1996;493:801–18.
[69] Hancox JC, Levi AJ, Lee CO, Heap P. A method for isolating rabbit atrioventricular
node myocytes which retain normal morphology and function. Am J Physiol
1993;265:H755–66.
[70] Yuill KH, Hancox JC. Characteristics of single cells isolated from the atrioventricular node of the adult guinea-pig heart. Pflugers Arch 2002;445:311–20.
[71] Dobrzynski H, Nikolski VP, Sambelashvili AT, Greener ID, Yamamoto M, Boyett
MR, et al. Site of origin and molecular substrate of atrioventricular junctional
rhythm in the rabbit heart. Circ Res 2003;93:1102–10.
[72] Liu J, Noble PJ, Xiao G, Abdelrahman M, Dobrzynski H, Boyett MR, Lei M, Noble D.
Role of pacemaking current in cardiac nodes: insights from a comparative study
of sinoatrial node and atrioventricular node. Prog Biophys Mol Biol 2008;96:
294–304.
[73] Weidmann S. Effect of current flow on the membrane potential of cardiac muscle.
J Physiol 1951;115:227–36.
[74] Vassalle M. Analysis of cardiac pacemaker potential using a “voltage clamp”
technique. Am J Physiol 1966;210:1335–41.
[75] Shi W, Yu H, Wu J, Zuckerman J, Wymore R. The distribution and prevalence of
HCN isoforms in the canine heart and their relation to the voltage dependence of
If. Biophys J 2000;78:353A (Abstract).
[76] Han W, Bao W, Wang Z, Nattel S. Comparison of ion-channel subunit expression
in canine cardiac Purkinje fibers and ventricular muscle. Circ Res 2002;91:790–7.
[77] Gaborit N, Le Bouter S, Szuts V, Varro A, Escande D, Nattel S, et al. Regional and
tissue specific transcript signatures of ion channel genes in the non-diseased
human heart. J Physiol 2007;582:675–93.
[78] Porciatti F, Pelzmann B, Cerbai E, Schaffer P, Pino R, Bernhart E, et al. The pacemaker
current I(f) in single human atrial myocytes and the effect of β-adrenoceptor and
A1-adenosine receptor stimulation. Br J Pharmacol 1997;122:963–9.
[79] Hoppe UC, Beuckelmann DJ. Characterization of the hyperpolarization-activated
inward current in isolated human atrial myocytes. Cardiovasc Res 1998;38:
788–801.
[80] Yu H, Chang F, Cohen IS. Pacemaker current i(f) in adult canine cardiac ventricular
myocytes. J Physiol 1995;485:469–83.
[81] Hoppe UC, Jansen E, Sudkamp M, Beuckelmann DJ. Hyperpolarization-activated
inward current in ventricular myocytes from normal and failing human hearts.
Circulation 1998;97:55–65.
[82] Cerbai E, Sartiani L, DePaoli P, Pino R, Maccherini M, Bizzarri F, et al. The
properties of the pacemaker current I(F)in human ventricular myocytes are
modulated by cardiac disease. J Mol Cell Cardiol 2001;33:441–8.
[83] Fernandez-Velasco M, Goren N, Benito G, Blanco-Rivero J, Bosca L, Delgado C.
Regional distribution of hyperpolarization-activated current (If) and hyperpolarization-activated cyclic nucleotide-gated channel mRNA expression in ventricular cells from control and hypertrophied rat hearts. J Physiol 2003;553:
395–405.
[84] Zicha S, Fernandez-Velasco M, Lonardo G, L'Heureux N, Nattel S. Sinus node
dysfunction and hyperpolarization-activated (HCN) channel subunit remodeling
in a canine heart failure model. Cardiovasc Res 2005;66:472–81.
[85] Stillitano F, Lonardo G, Zicha S, Varro A, Cerbai E, Mugelli A, et al. Molecular basis
of funny current (If) in normal and failing human heart. J Mol Cell Cardiol
2008;45:289–99.
[86] Opthof T. The membrane current (I(f)) in human atrial cells: implications for
atrial arrhythmias. Cardiovasc Res 1998;38:537–40.
[87] Lai LP, Su MJ, Lin JL, Tsai CH, Lin FY, Chen YS, et al. Measurement of funny
current (I(f)) channel mRNA in human atrial tissue: correlation with left atrial
filling pressure and atrial fibrillation. J Cardiovasc Electrophysiol 1999;10:
947–53.
[88] Workman AJ, Rankin AC. Serotonin, I(f) and human atrial arrhythmia. Cardiovasc
Res 1998;40:436–7.
[89] Cerbai E, Barbieri M, Mugelli A. Occurrence and properties of the hyperpolarization-activated current If in ventricular myocytes from normotensive and
hypertensive rats during aging. Circulation 1996;94:1674–81.
[90] Cerbai E, Pino R, Porciatti F, Sani G, Toscano M, Maccherini M, et al.
Characterization of the hyperpolarization-activated current, I(f), in ventricular
myocytes from human failing heart. Circulation 1997;95:568–71.
[91] Xiao J, Yang B, Lin H, Lu Y, Luo X, Wang Z. Novel approaches for gene-specific
interference via manipulating actions of microRNAs: examination on the
pacemaker channel genes HCN2 and HCN4. J Cell Physiol 2007;212:285–92.
[92] Luo X, Lin H, Pan Z, Xiao J, Zhang Y, Lu Y, et al. Down-regulation of miR-1/miR-133
contributes to re-expression of pacemaker channel genes HCN2 and HCN4 in
hypertrophic heart. J Biol Chem 2008;283:20045–52.
[93] Qu J, Barbuti A, Protas L, Santoro B, Cohen IS, Robinson RB. HCN2 overexpression
in newborn and adult ventricular myocytes: distinct effects on gating and
excitability. Circ Res 2001;89:E8–E14.
[94] Yu H, Wu J, Potapova I, Wymore RT, Holmes B, Zuckerman J, et al. MinK-related
peptide 1: a β subunit for the HCN ion channel subunit family enhances
expression and speeds activation. Circ Res 2001;88:E84–7.
[95] Decher N, Bundis F, Vajna R, Steinmeyer K. KCNE2 modulates current amplitudes
and activation kinetics of HCN4: influence of KCNE family members on HCN4
currents. Pflugers Arch 2003;446:633–40.
[96] Qu J, Kryukova Y, Potapova IA, Doronin SV, Larsen M, Krishnamurthy G, et al.
MiRP1 modulates HCN2 channel expression and gating in cardiac myocytes. J Biol
Chem 2004;279:43497–502.
[97] Proenza C, Angoli D, Agranovich E, Macri V, Accili EA. Pacemaker channels
produce an instantaneous current. J Biol Chem 2002;277:5101–9.
63
[98] Pian P, Bucchi A, Robinson RB, Siegelbaum SA. Regulation of gating and rundown
of HCN hyperpolarization-activated channels by exogenous and endogenous
PIP2. J Gen Physiol 2006;128:593–604.
[99] Zolles G, Klocker N, Wenzel D, Weisser-Thomas J, Fleischmann BK, Roeper J, et al.
Pacemaking by HCN channels requires interaction with phosphoinositides.
Neuron 2006;52:1027–36.
[100] Pian P, Bucchi A, Decostanzo A, Robinson RB, Siegelbaum SA. Modulation of cyclic
nucleotide-regulated HCN channels by PIP(2) and receptors coupled to
phospholipase C. Pflugers Arch 2007;455:125–45.
[101] Suchy SF, Olivos-Glander IM, Nussabaum RL. Lowe syndrome, a deficiency of
phosphatidylinositol 4,5-bisphosphate 5-phosphatase in the Golgi apparatus.
Hum Mol Genet 1995;4:2245–50.
[102] Pasquare SJ, Salvador GA, Giusto NM. Phospholipase D and phosphatidate
phosphohydrolase activities in rat cerebellum during aging. Lipids 2004;39:553–60.
[103] Suh BC, Hille B. Regulation of ion channels by phosphatidylinositol 4,5bisphosphate. Curr Opin Neurobiol 2005;15:370–8.
[104] Barbuti A, Terragni B, Brioschi C, DiFrancesco D. Localization of f-channels to
caveolae mediates specific β2-adrenergic receptor modulation of rate in
sinoatrial myocytes. J Mol Cell Cardiol 2007;42:71–8.
[105] Ye B, Balijepalli RC, Foell JD, Kroboth S, Ye Q, Luo YH, et al. Caveolin-3 associates
with and affects the function of hyperpolarization-activated cyclic nucleotidegated channel 4. Biochemistry 2008;47:12312–8.
[106] Barbuti A, Gravante B, Riolfo M, Milanesi R, Terragni B, DiFrancesco D.
Localization of pacemaker channels in lipid rafts regulates channel kinetics.
Circ Res 2004;94:1325–31.
[107] Accili EA, Redaelli G, DiFrancesco D. Differential control of the hyperpolarizationactivated current (i(f)) by cAMP gating and phosphatase inhibition in rabbit sinoatrial node myocytes. J Physiol 1997;500:643–51.
[108] Wu JY, Cohen IS. Tyrosine kinase inhibition reduces i(f) in rabbit sinoatrial node
myocytes. Pflugers Arch 1997;434:509–14.
[109] Altomare C, Tognati A, Bescond J, Ferroni A, Baruscotti M. Direct inhibition of the
pacemaker (If) current in rabbit sinoatrial node cells by genistein. Br J Pharmacol
2006;147:36–44.
[110] Yu H, Chang F, Cohen IS. Phosphatase inhibition by calyculin A increases i(f) in
canine Purkinje fibers and myocytes. Pflugers Arch 1993;422:614–6.
[111] Zong X, Eckert C, Yuan H, Wahl-Schott C, Abicht H, Fang L, et al. A novel
mechanism of modulation of hyperpolarization-activated cyclic nucleotide-gated
channels by Src kinase. J Biol Chem 2005;280:34224–32.
[112] Arinsburg SS, Cohen IS, Yu HG. Constitutively active Src tyrosine kinase changes
gating of HCN4 channels through direct binding to the channel proteins. J
Cardiovasc Pharmacol 2006;47:578–86.
[113] Li CH, Zhang Q, Teng B, Mustafa SJ, Huang JY, Yu HG. Src tyrosine kinase alters
gating of hyperpolarization-activated HCN4 pacemaker channel through Tyr531.
Am J Physiol, Cell Physiol 2008;294:C355–62.
[114] Huang J, Huang A, Zhang Q, Lin YC, Yu HG. Novel mechanism for suppression of
hyperpolarization-activated cyclic nucleotide-gated pacemaker channels by
receptor-like tyrosine phosphatase-alpha. J Biol Chem 2008;283:29912–9.
[115] Kryukova Y, Rybin VO, Qu J, Steinberg SF, Robinson RB. Age-dependent differences
in the inhibition of HCN2 current in rat ventricular myocytes by the tyrosine
kinase inhibitor erbstatin. Pflugers Arch 2009;457:821–30.
[116] Herrmann S, Stieber J, Stockl G, Hofmann F, Ludwig A. HCN4 provides a
‘depolarization reserve’ and is not required for heart rate acceleration in mice.
EMBO J 2007;26:4423–32.
[117] Hoesl E, Stieber J, Herrmann S, Feil S, Tybl E, Hofmann F, et al. Tamoxifeninducible gene deletion in the cardiac conduction system. J Mol Cell Cardiol
2008;45:62–9.
[118] Mistrik P, Mader R, Michalakis S, Weidinger M, Pfeifer A, Biel M. The murine HCN3
gene encodes a hyperpolarization-activated cation channel with slow kinetics
and unique response to cyclic nucleotides. J Biol Chem 2005;280:27056–61.
[119] Moroni A, Gorza L, Beltrame M, Gravante B, Vaccari T, Bianchi ME, et al.
Hyperpolarization-activated cyclic nucleotide-gated channel 1 is a molecular
determinant of the cardiac pacemaker current I(f). J Biol Chem 2001;276:
29233–41.
[120] Schulze-Bahr E, Neu A, Friederich P, Kaupp UB, Breithardt G, Pongs O, et al.
Pacemaker channel dysfunction in a patient with sinus node disease. J Clin Invest
2003;111:1537–45.
[121] Ueda K, Nakamura K, Hayashi T, Inagaki N, Takahashi M, Arimura T, et al.
Functional characterization of a trafficking-defective HCN4 mutation, D553N,
associated with cardiac arrhythmia. J Biol Chem 2004;279:27194–8.
[122] Milanesi R, Baruscotti M, Gnecchi-Ruscone T, DiFrancesco D. Familial sinus
bradycardia associated with a mutation in the cardiac pacemaker channel. N Engl
J Med 2006;354:151–7.
[123] Nof E, Luria D, Brass D, Marek D, Lahat H, Reznik-Wolf H, et al. Point mutation in
the HCN4 cardiac ion channel pore affecting synthesis, trafficking, and functional
expression is associated with familial asymptomatic sinus bradycardia. Circulation 2007;116:463–70.
[124] Qu J, Plotnikov AN, Danilo Jr P, Shlapakova I, Cohen IS, Robinson RB, et al.
Expression and function of a biological pacemaker in canine heart. Circulation
2003;107:1106–9.
[125] Plotnikov AN, Sosunov EA, Qu J, Shlapakova IN, Anyukhovsky EP, Liu L, et al.
Biological pacemaker implanted in canine left bundle branch provides ventricular escape rhythms that have physiologically acceptable rates. Circulation
2004;109:506–12.
[126] Bucchi A, Plotnikov AN, Shlapakova I, Danilo Jr P, Kryukova Y, Qu J, et al. Wild-type
and mutant HCN channels in a tandem biological-electronic cardiac pacemaker.
Circulation 2006;114:992–9.
64
M. Baruscotti et al. / Journal of Molecular and Cellular Cardiology 48 (2010) 55–64
[127] Tse HF, Xue T, Lau CP, Siu CW, Wang K, Zhang QY, et al. Bioartificial sinus node
constructed via in vivo gene transfer of an engineered pacemaker HCN Channel
reduces the dependence on electronic pacemaker in a sick-sinus syndrome
model. Circulation 2006;114:1000–1.
[128] Cai J, Yi FF, Li YH, Yang XC, Song J, Jiang XJ, et al. Adenoviral gene transfer of HCN4
creates a genetic pacemaker in pigs with complete atrioventricular block. Life Sci
2007;80:1746–53.
[129] Potapova I, Plotnikov A, Lu Z, Danilo Jr P, Valiunas V, Qu J, et al. Human
mesenchymal stem cells as a gene delivery system to create cardiac pacemakers.
Circ Res 2004;94:952–9.
[130] Yang XJ, Zhou YF, Li HX, Han LH, Jiang WP. Mesenchymal stem cells as a gene
delivery system to create biological pacemaker cells in vitro. J Int Med Res
2008;36:1049–55.
[131] Cho HC, Kashiwakura Y, Marban E. Creation of a biological pacemaker by cell
fusion. Circ Res 2007;100:1112–5.
[132] Wobus AM, Kleppisch T, Maltsev V, Hescheler J. Cardiomyocyte-like cells
differentiated in vitro from embryonic carcinoma cells P19 are characterized by
functional expression of adrenoceptors and Ca2+ channels. In Vitro Cell Dev Biol,
Anim 1994;30A:425–34.
[133] Kehat I, Kenyagin-Karsenti D, Snir M, Segev H, Amit M, Gepstein A, et al. Human
embryonic stem cells can differentiate into myocytes with structural and
functional properties of cardiomyocytes. J Clin Invest 2001;108:407–14.
[134] He JQ, Ma Y, Lee Y, Thomson JA, Kamp TJ. Human embryonic stem cells develop
into multiple types of cardiac myocytes: action potential characterization. Circ
Res 2003;93:32–9.
[135] Mummery C, Ward-van Oostwaard D, Doevendans P, Spijker R, van den Brink S,
Hassink R, et al. Differentiation of human embryonic stem cells to cardiomyocytes: role of coculture with visceral endoderm-like cells. Circulation 2003;107:
2733–40.
[136] Kehat I, Khimovich L, Caspi O, Gepstein A, Shofti R, Arbel G, et al. Electromechanical integration of cardiomyocytes derived from human embryonic stem
cells. Nat Biotechnol 2004;22:1282–9.
[137] Xue T, Cho HC, Akar FG, Tsang SY, Jones SP, Marban E, et al. Functional integration of
electrically active cardiac derivatives from genetically engineered human embryonic
stem cells with quiescent recipient ventricular cardiomyocytes: insights into the
development of cell-based pacemakers. Circulation 2005;111:11–20.
[138] Bucchi A, Tognati A, Milanesi R, Baruscotti M, DiFrancesco D. Properties of
ivabradine-induced block of HCN1 and HCN4 pacemaker channels. J Physiol
2006;572:335–46.
[139] Stieber J, Wieland K, Stockl G, Ludwig A, Hofmann F. Bradycardic and proarrhythmic
properties of sinus node inhibitors. Mol Pharmacol 2006;69:1328–37.
[140] Borer JS, Fox K, Jaillon P, Lerebours G. Antianginal and antiischemic effects of
ivabradine, an I(f) inhibitor, in stable angina: a randomized, double-blind,
multicentered, placebo-controlled trial. Circulation 2003;107:817–23.
[141] Tardif JC, Ford I, Tendera M, Bourassa MG, Fox K. Efficacy of ivabradine, a new
selective I(f) inhibitor, compared with atenolol in patients with chronic stable
angina. Eur Heart J 2005;26:2529–36.
[142] Ruzyllo W, Tendera M, Ford I, Fox KM. Antianginal efficacy and safety of
ivabradine compared with amlodipine in patients with stable effort angina
pectoris: a 3-month randomised, double-blind, multicentre, noninferiority trial.
Drugs 2007;67:05–393.
[143] Lopez-Bescos L, Filipova S, Martos R. Long-term safety and efficacy of ivabradine
in patients with chronic stable angina. Cardiology 2007;108:387–96.
[144] Fox K, Ford I, Steg PG, Tendera M, Ferrari R. Ivabradine for patients with
stable coronary artery disease and left-ventricular systolic dysfunction
(BEAUTIFUL): a randomised, double-blind, placebo-controlled trial. Lancet
2008;372:807–16.
[145] Fox K, Ford I, Steg PG, Tendera M, Robertson M, Ferrari R. Heart rate as a
prognostic risk factor in patients with coronary artery disease and leftventricular systolic dysfunction (BEAUTIFUL): a subgroup analysis of a
randomised controlled trial. Lancet 2008;372:817–21.
[146] Satoh H, Hashimoto K. Electrophysiological study of alinidine in voltage clamped
rabbit sino-atrial node cells. Eur J Pharmacol 1986;121:211–9.
[147] Goethals M, Raes A, Van Bogaert PP. Use-dependent block of the pacemaker
current I(f) in rabbit sinoatrial node cells by zatebradine (UL-FS 49). On the mode
of action of sinus node inhibitors. Circulation 1993;88:2389–401.
[148] Thollon C, Cambarrat C, Vian J, Prost JF, Peglion JL, Vilaine JP. Electrophysiological effects of S 16257, a novel sino-atrial node modulator, on rabbit and
guinea-pig cardiac preparations: comparison with UL-FS 49. Br J Pharmacol
1994;112:37–42.
[149] Bois P, Bescond J, Renaudon B, Lenfant J. Mode of action of bradycardic agent, S
16257, on ionic currents of rabbit sinoatrial node cells. Br J Pharmacol 1996;118:
1051–7.
[150] BoSmith RE, Briggs I, Sturgess NC. Inhibitory actions of ZENECA ZD7288 on
whole-cell hyperpolarization activated inward current (If) in guinea-pig
dissociated sinoatrial node cells. Br J Pharmacol 1993;110:343–9.
[151] Marshall PW, Rouse W, Briggs I, Hargreaves RB, Mills SD, McLoughlin BJ. ICI D7288, a
novel sinoatrial node modulator. J Cardiovasc Pharmacol 1993;21:902–6.