ArticlePDF AvailableLiterature Review

Genesis and Regulation of the Heart Automaticity

Authors:

Abstract and Figures

The heart automaticity is a fundamental physiological function in higher organisms. The spontaneous activity is initiated by specialized populations of cardiac cells generating periodical electrical oscillations. The exact cascade of steps initiating the pacemaker cycle in automatic cells has not yet been entirely elucidated. Nevertheless, ion channels and intracellular Ca(2+) signaling are necessary for the proper setting of the pacemaker mechanism. Here, we review the current knowledge on the cellular mechanisms underlying the generation and regulation of cardiac automaticity. We discuss evidence on the functional role of different families of ion channels in cardiac pacemaking and review recent results obtained on genetically engineered mouse strains displaying dysfunction in heart automaticity. Beside ion channels, intracellular Ca(2+) release has been indicated as an important mechanism for promoting automaticity at rest as well as for acceleration of the heart rate under sympathetic nerve input. The potential links between the activity of ion channels and Ca(2+) release will be discussed with the aim to propose an integrated framework of the mechanism of automaticity.
Content may be subject to copyright.
88:919-982, 2008. doi:10.1152/physrev.00018.2007 Physiol Rev
Matteo E. Mangoni and Joël Nargeot
You might find this additional information useful...
541 articles, 337 of which you can access free at: This article cites
http://physrev.physiology.org/cgi/content/full/88/3/919#BIBL
on the following topics:
http://highwire.stanford.edu/lists/artbytopic.dtlcan be found at Medline items on this article's topics
Physiology .. Mice
Medicine .. Pacemaker
Medicine .. Genetics
Medicine .. Cardiovascular Genetics
Physiology .. Nerves
Physiology .. Ion Channels
including high-resolution figures, can be found at: Updated information and services
http://physrev.physiology.org/cgi/content/full/88/3/919
can be found at: Physiological Reviewsabout Additional material and information
http://www.the-aps.org/publications/prv
This information is current as of July 15, 2008 .
http://www.the-aps.org/.website at
MD 20814-3991. Copyright © 2005 by the American Physiological Society. ISSN: 0031-9333, ESSN: 1522-1210. Visit our
published quarterly in January, April, July, and October by the American Physiological Society, 9650 Rockville Pike, Bethesda
provides state of the art coverage of timely issues in the physiological and biomedical sciences. It isPhysiological Reviews
on July 15, 2008 physrev.physiology.orgDownloaded from
Genesis and Regulation of the Heart Automaticity
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Institute of Functional Genomics, Department of Physiology, Centre National de la Recherche Scientifique
UMR5203, Institut National de la Sante´ et de la Recherche Me´dicale U661, University of Montpellier I and II,
Montpellier, France
I. Introduction 920
II. Automaticity in Cardiac Cells and Distribution of Pacemaker Activity in the Heart Tissue 922
A. Molecular determinants of SAN formation 922
B. Automaticity in the sinoatrial node 923
C. Properties of isolated SAN pacemaker cells 925
D. Pacemaker shift and extranodal supraventricular automaticity 926
E. Automaticity in the atrioventricular node 927
F. Automaticity in Purkinje fibers 929
III. Molecular Determinants of Ion Channels in Automatic Heart Cells 930
A. f-Channels 930
B. Voltage-dependent Ca
2
Channels 931
C. St-channels 932
D. Voltage-dependent Na
channels 932
E. Voltage-dependent K
channels 932
F. G protein-activated, ATP-dependent, and inward rectifier K
channels 933
G. Cl
channels, volume-activated channels, and stretch-activated cationic channels 934
IV. Pumps and Exchange Currents 934
A. The Na
-K
pump current I
p
934
B. The Na
-Ca
2
exchanger current I
NCX
and the Na
-H
exchanger 935
V. Patterns of Gene Expression in Adult Pacemaker Tissue 936
VI. Genesis of Cardiac Automaticity: Mechanisms of Pacemaking 937
A. Concepts 937
B. Ion channels and cardiac automaticity: general considerations 937
C. Role of I
Kr
and I
Ks
in automaticity 938
D. Role of I
f
in automaticity 939
E. Role of I
Ca,L
in automaticity 942
F. Role of I
Ca,T
in automaticity 945
G. Role of N- and R-type channels in heartbeat regulation 946
H. Role of I
Na
in automaticity 947
I. Role of I
st
in automaticity 948
L. SR Ca
2
release and automaticity 948
VII. Autonomic Regulation of Pacemaker Activity 951
A. Principles 951
B. Sympathetic regulation of pacemaker activity 952
C. Parasympathetic regulation of pacemaking 955
VIII. Cardiac Automaticity as an Integrated Mechanism: Numerical Modeling of Pacemaker Activity 957
A. General models of automaticity 957
B. Dedicated models of automaticity 958
IX. Additional Regulators of Cardiac Automaticity 960
A. Neuropeptides 960
B. Adenosine 961
C. Hormones 961
D. Mechanical load and atrial stretch 962
E. Electrolytes and temperature 962
X. Genetic and Acquired Diseases of Cardiac Automaticity 963
A. Inherited dysfunction of SAN automaticity 963
B. Automaticity in heart failure and cardiac ischemia 965
XI. Heart Automaticity and Cardioprotection 965
A. Heart rate and cardiac morbidity 966
Physiol Rev 88: 919–982, 2008;
doi:10.1152/physrev.00018.2007.
www.prv.org 9190031-9333/08 $18.00 Copyright © 2008 the American Physiological Society
on July 15, 2008 physrev.physiology.orgDownloaded from
B. Automaticity in engineered “biological pacemakers” 966
XII. Concluding Remarks 968
Mangoni ME, Nargeot J. Genesis and Regulation of the Heart Automaticity. Physiol Rev 88: 919-982, 2008;
doi:10.1152/physrev.00018.2007.—The heart automaticity is a fundamental physiological function in higher organ-
isms. The spontaneous activity is initiated by specialized populations of cardiac cells generating periodical electrical
oscillations. The exact cascade of steps initiating the pacemaker cycle in automatic cells has not yet been entirely
elucidated. Nevertheless, ion channels and intracellular Ca
2
signaling are necessary for the proper setting of the
pacemaker mechanism. Here, we review the current knowledge on the cellular mechanisms underlying the
generation and regulation of cardiac automaticity. We discuss evidence on the functional role of different families
of ion channels in cardiac pacemaking and review recent results obtained on genetically engineered mouse strains
displaying dysfunction in heart automaticity. Beside ion channels, intracellular Ca
2
release has been indicated as
an important mechanism for promoting automaticity at rest as well as for acceleration of the heart rate under
sympathetic nerve input. The potential links between the activity of ion channels and Ca
2
release will be discussed
with the aim to propose an integrated framework of the mechanism of automaticity.
I. INTRODUCTION
The heart pacemaker activity sets the rhythm and
rate of cardiac chamber contraction. Autonomic regula-
tion of heart rate plays a fundamental role in the integra-
tion of vital functions and influences animal behavior and
capability to respond to changing environmental condi-
tions. Heart rate has also been recently linked to the
overall risk of cardiovascular mortality and morbidity
(232).
In the adult heart of higher vertebrates, pacemaking
is generated by specialized “pacemaker” cells having low
contractility and generating a periodical electrical oscil-
lation (Fig. 1). Gene expression in cardiac pacemaker
cells seems to be qualitatively similar to that of working
myocytes, yet a quantitatively different level of expres-
sion of some ion channels, connexins (Cx), and transcrip-
tion factors generates the distinctive phenotype of spon-
taneously active myocytes (179, 317, 477). In spite of the
physiological importance of cardiac automaticity, some
aspects of the pacemaker mechanism have not been elu-
cidated and are still under debate. Indeed, different views
of the ionic and cellular basis of the pacemaker mecha-
nism and its regulation by the autonomic nervous system
have been proposed [see previous articles in this Journal
(63, 215) and Refs. 74, 120, 307].
Unraveling the mechanisms of pacemaking in spon-
taneously active cells is a fascinating and complex task.
Experimental and theoretical approaches, such as in vitro
and in vivo electrophysiology, pharmacology, and genet-
ics, as well as numerical modeling of pacemaking are
necessary. Automatic cells of the adult heart are charac-
terized by the presence of the diastolic depolarization, a
depolarizing phase that drives the membrane voltage at
the end of repolarization to the following action potential
threshold (Fig. 1B).
The pacemaker mechanism has been intensively
studied for more than 40 years. The diastolic depolariza-
tion is an electrical phenomenon, so pacemaking has been
first interpreted in terms of activation of specific ionic
currents (see Refs. 63, 215). Between 1960 and 1980,
pacemaker activity has been investigated by intracellular
recording of automaticity and ionic currents on sponta-
neously active tissue strips coming from the Purkinje fiber
network (118, 192, 485, 495) and the sinoatrial node (SAN)
(62, 364, 533). During these pioneering years, some ionic
currents involved in the generation of automaticity have
been described. Key breakthroughs were the discovery of
the hyperpolarization activated current (I
f
) in SAN (64)
and Purkinje fibers (118, 122) and the description of
the role of dihydropyridine (DHP)-sensitive Ca
2
current
(I
Ca
) in the SAN action potential generation (364). Very
low inward rectifier current (I
K1
) density is found in the
rabbit SAN. In contrast, automatic tissue expresses strong
I
f
(366). The SAN action potential is predominantly con-
trolled by I
Ca
so that the action potential upstroke veloc-
ity is much lower in the SAN than in the ventricle. The
SAN repolarization phase also differs from that of the
working myocardium. Indeed, as the SAN expresses low
levels of the transient outward current (I
to
), the repolar-
ization phase is predominantly controlled by the delayed
rectifier current (I
K
) (62). Finally, the SAN action poten-
tial also lacks the plateau phase. These electrophysiolog-
ical properties confer to the SAN action potential its
typical form (Fig. 1B).
The patch-clamp technique has become popular for
studying the ionic basis of pacemaker activity on individ-
ual pacemaker cells (215). Different families of ion chan-
nels have been described (58). Two distinct I
Ca
compo-
nents have been identified (172): the L-type Ca
2
current
(I
Ca,L
) and the T-type current (I
Ca,T
). A fast (I
Kr
) (451) and
a slow (I
Ks
) (278) delayed rectifier has been reported.
Furthermore, application of the patch-clamp technique
has improved the study of the sensitivity of these chan-
nels to autonomic agonists. Two major insights into the
physiological role of ion channels in the generation and
920
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
regulation of pacemaker activity have been obtained: the
relevance of f-channels in the regulation of heart rate at
low parasympathetic tone (124) and the physiological
significance of the heterogeneity of ion channel expres-
sion in the SAN (58). The importance of the Na
-Ca
2
exchanger (NCX) in pacemaking in the amphibian sinus
venosus (227, 228) and mammalian SAN (44) has been
recently highlighted, and the contribution of spontaneous
intracellular Ca
2
release in the generation of automatic-
ity has been emphasized (510).
New important insights into the pacemaker mecha-
nism are now coming from the study of genetically mod-
ified mouse strains in which specific ion channels have
been inactivated or modified. This genetic approach con-
stitutes a necessary implementation to electrophysiologi-
cal and pharmacological evidence, since selective inhibi-
tors of ion channels are not always available. Gene tar-
geting in the mouse has generated interesting models of
dysfunction of pacemaker activity (393, 467, 522). The
possibility to study pacemaker activity in mouse models is
very recent (312, 313) and has provided exciting insights
into the specific functional role of L-type Ca
v
1.3 (309, 546)
and T-type Ca
v
3.1 (314) channels in the generation of SAN
automaticity. Because of the tiny size of the dominant
pacemaker region in the mouse heart (498), isolation of
mouse SAN cells is technically challenging. Our group has
been the first to obtain recordings of automaticity and ion
channels from isolated mouse SAN cells (312, 313). Other
groups have successfully employed this new preparation
to study pacemaker activity in wild-type (86, 93, 279) and
genetically modified mouse strains (196, 294, 309, 314,
546). Limitations of the use of the mouse for the study of
cardiac automaticity are correlated with the very high
basal heart rate of this species. We can thus expect that
pacemaker mechanisms may assume a different role in
mice than in larger mammals and humans. Nevertheless,
the new possibility of isolating mouse SAN tissue and
primary pacemaker cells has created a new interest into
the study of the ionic basis of pacemaker activity gener-
ation and regulation.
Until the last decade, research on the physiology of
heart automaticity was limited to the domain of basic
science. However, there is currently a renewal of interest
on cardiac pacemaking, and an increasing number of
laboratories are now focusing their efforts on the regula-
tion of heart rate. Indeed, the pharmacological control of
cardiac automaticity is now becoming an important issue
in the management of ischemic heart diseases (123). Iden-
FIG.1.A: the mammalian heart with the cardiac conduction system.
The sinoatrial node (SAN) is located at the entry of the superior vena
cava (SCV) in the right atrium (RA). The atrioventricular node (AVN)
extends in a region delimited by the inferior vena cava (ICV), the central
fibrous body (CFB), and the tricuspid valve (TV). The atrioventricular
bundle (AVB) divides in the bundle branches (BB) and originates the left
and right Purkinje fibers network (PFN). Other abbreviations are: LA,
left atrium; PV, pulmonary veins; MV, mitral valve; RV, right ventricle;
LV, left ventricle. [Adapted from Moorman and Christoffels (341).]
B: recordings of automaticity and action potential waveforms of isolated
mouse SAN, AVN, and PFN cells. Cl, cycle length; APD, action potential
duration; E
th
, action potential threshold; LDD, linear part of the diastolic
depolarization; EDD, exponential part of the diastolic depolarization.
(From Marger Nargeot and Matteo Mangoni, unpublished observations.)
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
921
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
tification of ion channels involved in the generation of
automaticity also constitutes the basis for the develop-
ment of putative genetic and cellular therapies for dys-
functions of pacemaking (94, 423, 424). Finally, knowl-
edge of pacemaker mechanisms is fundamental for under-
standing the basis of inherited diseases of the heart
rhythm.
In this review, we will focus on the cellular mechanisms
underlying heart automaticity and its regulation by the au-
tonomic nervous system in the adult heart. Different views
of automaticity will be compared and commented. To this
aim, we will try to give an overview of how ion channels and
Ca
2
signaling can influence pacemaking under different
physiological conditions. Finally, we will provide the reader
with a short discussion on how ion channels involved in
pacemaker activity are becoming useful targets in the devel-
opment of new therapies for heart diseases.
II. AUTOMATICITY IN CARDIAC CELLS AND
DISTRIBUTION OF PACEMAKER ACTIVITY
IN THE HEART TISSUE
In the mammalian heart, three major structures are
endowed with automaticity and are able to drive the
heartbeat: the SAN, which constitutes the physiological
pacemaker, the atrioventricular node (AVN), and the Pur-
kinje fibers network (Fig. 1A). Functional evidence in
favor of a supraventricular dominant pacemaker was ob-
tained by Gaskell in the slowly beating tortoise heart (154,
155). Anatomically, the SAN was identified a century ago
by Keith and Flack (237), shortly after Tawara’s descrip-
tion of the atrioventricular junction (476; see Ref. 460 for
review). The intrinsic SAN beating rate is normally faster
than that of the cardiac conduction system and sup-
presses pacemaking in the AVN and Purkinje network.
However, automaticity in AVN can become dominant in
case of SAN block or failure (220). Purkinje fibers can
also generate a viable rhythm in conditions of atrioven-
tricular block. For these reasons, the SAN region is indi-
cated as the primary pacemaker, while the AVN and Pur-
kinje fibers are indicated as secondary (or accessory)
pacemakers. In this section, we review basic principles of
automaticity in primary and secondary pacemaking re-
gions. Some aspects of the structure and function of SAN,
AVN, and the His-Purkinje system are commented on.
A. Molecular Determinants of SAN Formation
In the early embryonic heart, all cardiomyocytes dis-
play automaticity. The embryonic myocardium of the pri-
mary heart tube is characterized by slow conduction ve-
locity, and even if dominant pacemaker activity is present
at the venous pole, automaticity can be originated at any
point of the heart tube (341). The differentiation of the
cardiac chambers (heart ballooning) and the maturation
of myocytes forming the working myocardium of the late
embryo are associated with the disappearance of this kind
of “diffuse” automaticity. Working myocytes of the ma-
ture cardiac chambers have completely lost automaticity
and display fast conduction velocity (see Refs. 341, 386
for review). In the adult heart, pacemaker activity is con-
fined to the SAN and AVN, two regions in which the
intercellular conduction velocity is relatively slow. The
mechanisms underlying the confinement of pacemaker
activity and the patterning of the SAN and AVN during
development are mostly unresolved. However, progress
has been recently obtaining in the identification of tran-
scription factors involved in the differentiation of cell
lineages contributing to the SAN formation (207, 208, 338,
341).
Mommersteeg et al. (338) reported that the SAN de-
velops in the inflow tract region of the embryonic heart
(Fig. 2). This region is characterized by expression of the
T-box transcription factor Tbx3 and the Hcn4 ion channel
gene (coding for f-channels). This SAN primordium is also
devoid of Cx40, which is typical of fast-conducting work-
ing myocardium (Fig. 2). Mommersteeg et al. (338) have
also provided evidence that the Nkx2–5 transcription fac-
tor (NK2 transcription factor related to locus 5 of Dro-
sophila) may act as a repressor of the SAN lineage gene
program, since mice lacking Nkx2–5 show ectopic expres-
sion of Tbx3 and Hcn4 throughout the heart tube. Intact
activity of Nkx2–5 is also required to suppress Tbx3 ex-
pression outside the developing SAN and to trigger the
atrial gene expression program (338). If Nkx2–5 appears
to suppress SAN formation through inhibition of HCN4
expression and other markers, Tbx3 constitutes an acti-
vator of the SAN gene expression program (208) and a
repressor of heart chamber formation. Expression of Tbx3
in the whole heart tube suppresses chamber formation
(207). Mouse embryos in which Tbx3 has been inactivated
show that this gene is essential to activate the SAN related
gene expression program, and to prevent expression of myo-
cardial markers in the SAN region. Ectopic expression of
Tbx3 in the adult atria of transgenic mice is sufficient to
reactivate focal automaticity and some SAN markers (208).
Hoogars et al. (208) have proposed that SAN forma-
tion (at least starting from E10) is due to proliferation of
Tbx3-expressing cells rather than to recruitment of myo-
cytes having initiated the atrial gene program. Also, label-
ing of cell lineages suggests that the SAN is to be consid-
ered as a remnant of cell populations that did not enter
the myocardial expression program (338). Blaschke et al.
(40) have reported that the homeodomain transcription
factor Shox2 is involved in the formation of the SAN
region. Particularly, Shox2 knockout mice die around E12
and show severe SAN hypoplasia. Furthermore, zebrafish
embryos lacking Shox2 are bradycardic, an observation
922
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
which is consistent with the involvement of Shox2 in the
determination of the pacemaker tissue (40).
B. Automaticity in the Sinoatrial Node
In this section, we aim to provide a concise overview
of automaticity in the SAN. Some fundamental structural
aspects of the SAN cellular and electrophysiological or-
ganization will be discussed. Because the rabbit has been
extensively used as an animal model to study the mech-
anisms of pacemaking, we will focus on results obtained
from the rabbit atrial-SAN preparation. Further discus-
sion of the SAN structure and variations between species
can be found in some excellent reviews by Opthof (374)
and Boyett and co-workers (57, 58).
SAN pacemaker tissue is located in the intercaval
region (Fig. 3A) and extends towards the endocardial side
of the crista terminalis (Fig. 3, B–D). Spontaneously ac-
tive cells are found in the area delimited by the crista
terminalis, the left and right branch of the sinoatrial ring
bundle and the interatrial septum (Fig. 3, B–D). Beside
pacemaker cells, the SAN also contains atrial cells, fibro-
blasts, and adipocytes (130). The collagen content of the
SAN region is relatively high (377). However, collagen
does not seem to have an adverse effect on SAN conduc-
tion, because species having different SAN collagen con-
tent display similar conduction times (377). During aging,
the size and position of SAN do not change, but its struc-
ture undergoes remodeling associated with an augmenta-
tion of collagen content (6).
The overall electrical coupling within the SAN is
weak (see Ref. 58 for review). Two lines of evidence
indicate that SAN automaticity is in fact partially sup-
pressed by the electrotonic load imposed by the atrium.
First, removal of the right atrium results in an accelera-
tion of SAN pacing rate (241, 248). Second, direct cou-
pling of a SAN cell to an atrial cell suppresses pacemaker
activity (224, 482). However, a relatively low electrical
coupling between pacemaker cells seems to be necessary
for the SAN to be able to drive the atrium. Indeed, by
employing numerical modeling, Joyner and van Cappele
(225) have indicated that the SAN should have a minimal
size to drive the atrium, but also that intranodal cellular
coupling should be relatively weak to protect pacemaking
from the hyperpolarizing electrical load imposed by the
atrium. Watanabe et al. (514) have also used numerical
modeling to show that stronger electrical coupling be-
tween SAN and atrium would stop pacemaking.
The SAN is structurally heterogeneous. Two views of
the cellular organization of the SAN have been proposed.
In the “gradient” model of SAN (58), there is a progressive
transition in the size and electrical properties of pace-
maker cells between the SAN center and the periphery.
Cells in the center (labeled in red in Fig. 3) are small and
have intrinsically slower pacing rate and upstroke veloc-
ity compared with that at the periphery (labeled in blue in
Fig. 3), which have faster pacing rate, upstroke velocity,
and intermediate properties between “pure” pacemaker
and atrial cells. Verheijck et al. (499) have proposed that
the SAN region is constituted by automatic cells having
FIG.2.A: computer-based 3-dimensional reconstructions (465) of the mouse embryonic heart at 14.5 post coitum (E14.5). In this reconstruction,
the left ventricular myocardium has been removed to show the blood-filled lumen (orange), from the dorsal (top panels) and the right side (bottom
panels). The Tbx3-positive myocardium is in red, and the Cx40-negative myocardium is in grey. Note that expression of Tbx3 is restricted to areas
in which Cx40 expression is absent. B: schematic representation which shows the patterns of expression of critical markers in the developing fast
conducting myocardium or SAN primordia (see sect. IIA). Left panel delineates the expression pattern until E9.5, and middle panel shows the pattern
between E9.5 and E14.5. Dotted lines identify borders of Nkx2–5 expression domains. Except for the yellow-colored mesenchyme, only myocardium
is depicted (see legend in right panel). Abbreviations are as follows: as, atrial septum; avc, atrioventricular canal; ift, inflow tract; la/ra, left and right
atrium; laa/raa, left and right atrial appendage; lsh/rsh, left and right sinus horn; lv/rv, left and right ventricle; pv, pulmonary vein; san, sinoatrial node;
vv, venous valves. [From Mommersteeg et al. (338).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 923
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
variable intrinsic pacing rate and action potential config-
uration. Contrary to the “gradient” model, there would be
no preferential distribution of small pacemaker cells in
the SAN center. Indeed, pacemaker cells having different
degrees of automaticity and action potential configuration
are supposed to be uniformly distributed in the SAN
region. Verheijck et al. (499) have indicated that atrial
cells are present in the SAN, their density being maximal
in the SAN periphery and minimal in the center. This
latter view of the SAN structure has been called the
“mosaic” model.
Consistent with the gradient model of SAN is the
observation that tissue balls isolated from the rabbit SAN
periphery display a distinct electrophysiological profile
compared with that from the center (57, 60, 248, 250, 355,
370) (Fig. 4A). More importantly, a “central” and “periph-
eral” SAN can be defined by using electrophysiological
and histochemical criteria (57, 130). Consistently with the
mosaic model, atrial cells have been unambiguously iden-
tified in the peripheral SAN by using a Cx40
EGFP/
knock-in mouse line (331) and by direct staining with
Cx40 antibodies (130). Verhejick et al. (498) have re-
corded atrial action potential activity in the intact mouse
SAN (Fig. 4B). However, atrial cells seem to be absent in
the rabbit central SAN. In conclusion, even if the gradient
model seems to explain better SAN electrophysiological
behavior, certain elements of the mosaic model are to be
included in future structural SAN models.
Dobrzinsky et al. (130) have recently proposed the
first comprehensive structural computer model of the
rabbit SAN (see Fig. 3). In this model, peripheral SAN
tissue constitutes the SAN impulse exit pathway. Periph-
eral cells can be defined by membrane expression of Cx43
and middle neurofilament (NF-M) protein. SAN peripheral
cells are large and predominantly arranged in parallel.
Cells in the SAN center are smaller and express the
low-conductance Cx45 and NF-M. Central pacemakers
are arranged in a mesh and wrapped around bundles of
connective tissue. The mouse SAN seems to share some
common features with the rabbit SAN (289). In the central
mouse SAN, cells are packed in a compact structure
(compact node) and are oriented perpendicularly to the
crista terminalis (289). Cells at the periphery are loosely
packed and are oriented parallel to the crista. Numerical
simulations of impulse propagation from the leading
pacemaker site to the atrium have indicated that the SAN
could not sustain atrial beating if cellular coupling and
conduction were uniformly low throughout the SAN
(224). Expression of different connexins in the SAN cen-
ter and periphery contributes to create a gradient in im-
pulse conduction velocity from the leading pacemaker
site to SAN periphery and exit zone, so that conduction
velocity in central SAN is lower than that measured in the
periphery (see Ref. 61 for review). Beside Cx45, expres-
sion of Cx30.2 has been also reported in the mouse SAN
FIG. 3. A 3-dimensional computer model of the rabbit SAN as
proposed by Dobrzynski et al. (130). In all panels, central SAN tissue is
labeled in red, while peripheral SAN tissue is depicted in green. The
yellow dot indicates the leading SAN pacemaking site, and the blue line
indicates the SAN ring. A: topographical organization of the rabbit SAN,
as viewed in an ideal intact heart diagram. Note the extension of
peripheral SAN tissue around the superior vena cava (see also Fig. 6C).
Ao, aorta; PA, pulmonary artery; CS, coronary sinus. Other abbreviations
are as in Fig. 1. B and C: endocardial (left) and epicardial (right) view of
SAN central and peripheral tissue. In B, SAN myocytes are shown in a
simple framework model of the right atrium. In C, myocyte orientation
is shown. Note the mesh organization of central SAN tissue. Peripheral
SAN myocytes are predominantly oriented perpendicularly to the crista
terminalis and extend toward the interatrial septum (SEP). D: model
section of the SAN cut at the level of the white line in C. Note that
peripheral SAN tissue extends onto the endocardial side of the crista
terminalis. Central SAN myocytes are embedded in connective tissue.
[From Dobrzynski et al. (130).]
924 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
and AVN (257). Cx30.2 has smaller conductance than
Cx45 (70). It has been proposed that Cx30.2 contributes to
the intrinsically slow conduction velocity in the SAN and
in the AVN (258). This view is supported by the observa-
tion that mice in which Cx30.2 has been inactivated have
shortened P-Q intervals (256).
Expression of Cx43 in the SAN periphery enhances
cell-to-cell electrical coupling thereby forming a “transi-
tional” zone between the slow-conducting leading pace-
maker site and the fast-conducting atrial tissue (130).
Kodama et al. (248) have shown that small tissue
balls from the SAN periphery have intrinsically faster
pacing rate than balls from the central SAN (Fig. 3A).
However, as reported by Bleeker (41), the central SAN
can lead pacemaker activity in spite of its lower beating
rate, because cells at the periphery are electrotonically
inhibited by the right atrium. Differences in intrinsic firing
rates of central and peripheral SAN are due to heteroge-
neous expression of ion channels (60, 250, 355) and pro-
teins involved in Ca
2
homeostasis (267, 349). Particu-
larly, larger cells from SAN periphery express higher I
f
and I
Na
densities than small cells in the SAN center.
Enhanced expression of I
f
and I
Na
can explain the faster
pacing rate and upstroke velocity recorded in tissue from
the SAN periphery. The stronger pacemaker activity in
cells of the SAN periphery can help overcome the sup-
pressive effect of the right atrium on the overall SAN
automaticity (58). It would be important to develop a
comprehensive model of SAN activity including both
structural and electrophysiological data. Theoretically,
such a model could be integrated in an anatomical and
biophysically detailed three-dimensional model of the
heart (102) and would be an important step for under-
standing some integrative properties of heart automatic-
ity at the organ level.
The physiological role of fibroblasts in the SAN has
not been completely elucidated. Fibroblasts form an ex-
tensive network in the SAN (73, 105). Camelliti et al. (73)
have shown that fibroblasts can functionally connect with
SAN myocytes, possibly by Cx45-mediated gap junctions.
Apparently, SAN fibroblasts do not participate to intran-
odal conduction (105). However, they may act in the
transduction of the atrial wall stress to SAN myocytes. In
this respect, Kohl et al. (252) have proposed that fibro-
blasts participate in the mechanoelectrical regulation of
SAN rate.
C. Properties of Isolated SAN Pacemaker Cells
Isolated SAN pacemaker cells are a widespread ex-
perimental model for studying the pacemaker mecha-
nism. Several research groups have described the electro-
physiological properties of enzymatically isolated calci-
um-tolerant (216) pacemaker cells from a variety of
mammals including the rabbit (114, 125, 492), guinea pig
(12, 333), pig (372), rat (457), and mouse (86, 313). The
gross morphology of spontaneously active cells is con-
served between species. Cells from the rabbit and mouse
SAN have been empirically classified in three distinct
morphologies, namely, “spindle,” “elongated,” and “spi-
der” (125, 313, 499) (Fig. 5, A and B). In the rabbit and
mouse, spindle-shaped cells are generally smaller than
elongated and spider cells, their capacitance varying be-
tween 15 and 30 pF. Under Nomarski optics, the cyto-
plasm of spindle cells appears transparent and poor in
myofilaments. Elongated cells have a longer longitudinal
axis and larger capacitance (between 35 and 50 pF) than
spindle-shaped cells. In some elongated cells, the cyto-
plasm is darker than in spindle cells. This is possibly due
FIG.4.A: pacemaker activity of isolated tissue balls from the SAN center, periphery and transitional zone. Tissue from the SAN center has
intrinsically slower automaticity and more positive maximum diastolic potential, while peripheral and transitional tissues are intrinsically faster and
have a more negative maximum diastolic potential. [Original recordings from Boyett et al. (59).] B: intracellular microelectrode recordings of
automaticity in an isolated mouse SAN. The leading pacemaker site is indicated by asterisk. Note the presence of the diastolic depolarization phase
in this recording site (right panel) and suppression of automaticity in a location in the vicinity of the leading site (left panel), and in the SAN
periphery close to the interatrial septum (IAS). Atrial-type action potentials are recorded in the mouse SAN (right panel). Automaticity is not
recorded in the crista terminalis (CT) and in the right branch of the sinoatrial ring (SARB). [Original recordings from Verheijck et al. (498), with
permission from Elsevier.]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
925
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
to a higher myofilament density than in spindle cells (Fig.
5A). Spider cells have branched cytoplasm (see Fig. 5C).
The functional significance of such a peculiar morphology
is not known. In the rabbit, small spindle-shaped cells are
considered to be leading pacemaker cells (presumably
from the SAN center), while larger elongated cells come
from the SAN periphery (204). However, this may not be
an absolute rule, since small clusters of mouse SAN cells
containing both spindle and elongated cells have been
observed in the mouse SAN (499) (Mangoni and Nargeot,
unpublished observations).
D. Pacemaker Shift and Extranodal
Supraventricular Automaticity
The position of the leading pacemaker site is not
fixed. The pacemaker initiation site can shift in different
zones of the SAN. Pacemaker shifts can be induced by
autonomic regulatory inputs, temperature, atrial stretch,
as well as by pharmacological interventions (Fig. 6A) (58,
442). Experimentally, pacemaker shift has been studied in
the rabbit SAN (374), but naturally occurs also in dogs
(47) and humans. Variability in the position of the leading
pacemaker site in human subjects has been shown by
Boineau et al. (46). It is possible that pacemaker shift
underlies dynamic changes in the morphology of the P-
wave observed in ECG recordings (442). Schuessler et al.
(442) have reported that pacemaking can be located at
any point of the intercaval region as well as, in some
cases, in the left atrium. Interestingly, cells from the
sleeves of pulmonary veins have recently been shown to
display spontaneous activity (83, 205).
The leading pacemaker site is situated where local
automaticity is faster. We can thus expect that sympa-
thetic input will shift the leading site where automaticity
is most sensitive to adrenergic stimulation, while para-
sympathetic activity shifts pacemaking where automatic-
ity is less inhibited by cholinergic transmitters (300).
Boyett et al. (58) have attributed pacemaker shift to the
heterogeneous expression of ion channels in the SAN
center and periphery. Accordingly, a given channel blocker
will shift pacemaking to a site where automaticity is less
dependent from the targeted channel. This hypothesis has
been experimentally verified by using Ca
2
(250), K
(249),
and f-channel inhibitors (355). In the rabbit, shifts induced
by epinephrine and acetylcholine roughly correspond to
SAN zones in which I
f
block has very moderate effect on
pacing rate, an observation that matches the framework
proposed for channel blockers (58). Yamamoto et al. (531)
have shown that application of isoproterenol to intact atria
can shift the leading pacemaker site from the central SAN, to
a location between the superior vena cava and the interatrial
groove. Under basal conditions, automaticity in this site is
less robust than in the SAN, but under isoproterenol, it can
be as fast as in the SAN. These observations are indicative of
the existence of multiple functional pacemaker sites outside
the “classical” SAN region (Fig. 6B).
There is now substantial evidence indicating that
myocytes capable to drive pacemaking are present out-
side the SAN. Indeed, spontaneously beating cells can be
isolated from the right atrium (408, 427) and from the
bundle branches of the mouse heart (Mangoni, unpub-
lished observations). Automatic cells are also present
around the tricuspid valve (11). These cells are very likely
to be part of an extranodal extension of the atrial con-
duction system. Importantly, the existence of such an
extension has been recently demonstrated by staining of
FIG. 5. Morphology of isolated mouse (A) and rabbit (B) SAN cells.
Atrial cells can be isolated from the SAN in both species. C: a simple
structural model of a rabbit SAN spider cell. These cells have branched
cytoplasm (arrows) and are mononucleated. [Photographs in A from
Mangoni and Nargeot (312), with permission from Elsevier; pictures in B
and C from Verheijck et al. (499).]
926 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
rabbit and rat atria with antibodies directed against Cx45
and HCN4 (Fig. 6C) (531). Particularly, it has been shown
that cells having “nodal” phenotype extend from the rear
of the superior vena cava to the interatrial groove. Nodal-
like cells with high expression of both Cx45 and HCN4 are
present around the atrioventricular valves (the atrioven-
tricular ring bundles) and reach the AVN (531). The pres-
ence of an atrial conduction system extension may be due
to common embryonic origin and/or development of
these cells with that of SAN and AVN (411).
The physiological significance of automaticity in la-
tent pacemaker cells of the right atrium is unclear. These
cells have been shown to be capable of assuming the
control of the heartbeat under conditions of SAN failure
and to play a role in the generation of atrial arrhythmias
(426). In the left atrium, cells generating focal discharge
have been found at the boundary between the left atrium
and the pulmonary veins in human subjects affected of
paroxystic atrial fibrillation (532). The key role of these
cells in triggering and sustaining atrial fibrillations has
been clearly recognized (175, 176). Tissue sleeves from
rabbit pulmonary veins display transient automaticity
triggered by rapid pacing, application of ryanodine (205,
531), and stretch (83). A diastolic depolarization phase
can be recorded in these cells, but automaticity is insen-
sitive to I
f
block by Cs
(531). The expression of ionic
currents in these cells resembles much more that of non-
automatic atrial than pacemaker cells (141). The ionic
mechanism responsible for pacemaking will need inves-
tigation in the near future.
In conclusion, one can wonder why some automatic
sites outside the SAN (e.g., the interatrial groove exten-
sion) can effectively sustain normal pacemaking while
others can trigger atrial arrhythmias. In this respect, it is
worth noting that pacemaking in these regions is intrinsi-
cally less robust than that of the SAN (or less sensitive to
autonomic regulation). Consequently, in case of activa-
tion of these sites, atrial activation may not have the
correct space- and frequency-dependent properties.
E. Automaticity in the Atrioventricular Node
The AVN sets the appropriate frequency-dependent
conduction delay between the atria and ventricles. It also
limits ventricular activation during atrial tachyarrhyth-
mias, thereby protecting ventricular rhythm. The AVN has
a dual electrical input from the atrium, namely, the “fast”
FIG. 6. Pacemaker shift and extranodal supraventricular automa-
ticity in the rabbit heart. A: shift of the leading pacemaker site inside the
rabbit SAN region. In the left panel, SAN activation after impulse gen-
eration in a central leading pacemaker site (marked as the point acti-
vating at zero time) is depicted. Isochronal curves indicate the spread of
impulse conduction toward the crista terminalis. After activation of the
crista terminalis, the impulse propagates in the right atrium (30 ms
isochronal) and to the right and left branches of the SAN ring bundles
(RBSARB and LBARB). Note the block of impulse conduction in the
direction of the interatrial septum. [From Boyett et al. (59).] The right
panel indicates the shift of the leading pacemaker site under the influ-
ence of different ionic and pharmacological agents. [From Boyett et al.
(58).] B: variability of the position of the leading pacemaking site in
isolated rabbit hearts. The leading site can be found inside and some-
times outside the “classical” SAN region. Asterisks indicate the position
of the impulse origin in each of the 18 preparations tested. C: dorsal view
of the left and right atria showing the extension of the atrial conduction
system in the rabbit heart. Extension is defined according to extranodal
distribution of HCN4 (red) and Cx45 (yellow) immunoreactivity (see
text of sect. IIC). Positively stained cells are found around the mitral and
tricuspid valves, as well in the interatrial groove. [Modified from
Yamamoto et al. (531).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 927
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
and “slow” pathways of conduction (336, 356). However,
the AVN is also endowed of automaticity. AVN automa-
ticity can effectively drive the heart after SAN failure or
block.
The intact AVN is a complex and highly heteroge-
neous structure. For a recent review on the AVN structure
and conduction, the reader is referred to Efimov et al.
(140). Here, we will discuss the principal aspects of the
atrioventricular junction and automaticity. In spite of a
growing knowledge about the histology and electrophys-
iology of AVN, a detailed correlation between the differ-
ent anatomical components and action potential configu-
rations recorded from the AVN is not yet available. De-
bate still exists between morphologists and physiologists
about which structures of the atrioventricular junction
should be included in an ideal “true” AVN (140). In this
review, we will endorse the physiological definition of the
AVN, as indicated by Billette (39). This includes all struc-
tures contributing to the atrial-His conduction interval.
We thus define the AVN as the region comprised in the
Koch’s triangle, with the heterogeneous structure in-
cluded in the central fibrous body, as well as the posterior
nodal extension (PNE) projecting in the isthmus below
the coronary sinus (Fig. 7A). The central fibrous body
contains the enclosed part of the AVN with the compact
node (CN) and lies at the apex of the triangle. The action
potential configuration and expression of ion channels in
the AVN are heterogeneous. Three types of action poten-
tial waveforms have been identified in early studies (see
Ref. 140 for review): atrionodal (AN), nodal (N), and
nodo-His (NH). The N-type action potential is character-
ized by slow upstroke velocity and action potential am-
plitude, while AN and NH action potentials have interme-
diate properties between nodal and atrial or His action
potentials, respectively. Billette (38) has presented a mi-
croelectrode mapping study of the AVN and defined six
cell types based on action potential morphology and re-
fractoriness. It would be very important to build a com-
prehensive model of the functioning of the AVN. How-
ever, it is still difficult to correlate action potential types
with the structural organization of the atrioventricular
tissue. Histologically, the AVN is composed by a “super-
ficial” (subendocardial) layer and “middle” and “deep”
innermost layers (37a). Based on simultaneous microelec-
trode recording and optical mapping of rabbit AVN acti-
vation under SAN rhythm, Efimov and Mazgalev (139)
have supported the hypothesis that the superficial layer is
predominantly composed by AN-type cells, while the mid-
dle layer is composed of cells having N-type properties.
The AVN superficial layer is rich in Na
channels (391)
and Cx43, while the intermediate layer forming the CN is
essentially poor in Na
channels (391, 441). Furthermore,
the intermediate layer expresses the low-conductance
Cx45 rather than Cx43 or Cx40 (61, 140). Consistently,
lower conduction velocities are recorded in the middle
N-type layer than in the superficial AN-type layer (139).
FIG.7.A: a computer model of the rabbit atrioventricular junction
shows the basic structural organization of the AVN. The AVN is a hetero-
geneous structure delimited by the tricuspid valve, the tendon of Todaro,
and the coronary sinus. The posterior nodal extension (PNE) is depicted in
red and constitutes a major site of origin of junctional AVN automaticity. It
can also be part of the slow pathway of atrioventricular conduction. The
enclosed node is shown in purple and is continuous to the PNE. The green
zone corresponds to AVN tissue composed by loosely packed atrial cells. It
is possible that these atrial cells are part of the fast AVN conduction
pathway. [From Boyett et al. J Electrocardiol 38 Suppl: 113-120, 2005, with
permission from Elsevier.] B: mapping of automaticity in the rabbit AVN.
Pacemaking originates in the PNE and spreads to the atrial muscle and to
the enclosed node (top panel). Optical recording of automaticity in the PNE
is shown in the bottom panel.
928 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
The location of the leading AVN pacemaker site has
been a matter of debate. Initiation of automaticity has
been identified in the N-NH part of the node (140, 515).
Recently, Dobrzynski et al. (131) attempted to locate the
site of origin of pacemaker activity in isolated rabbit AVN
preparations. These authors have employed optical map-
ping with voltage-sensitive dyes to study the spread of
excitation from the automatic site to the fast and slow
AVN conduction pathways. In 14 preparations studied,
pacemaker activity originated in the PNE in 10 prepara-
tions (Fig. 7B). Pacemaking originated in the N/NH region
of AVN in four other preparations. Dobrzynski et al. (131)
also reported that the PNE pacemaker site expresses high
density of neurofilament 160 (NF160), Cx45, and the
HCN4 protein which codes for “pacemaker” f-channels.
Finally, the PNE can constitute a region of slow conduc-
tion during AVN reentry and premature beats (131).
In conclusion, the PNE extension of AVN is able to
generate pacemaking and can effectively pace the atrio-
ventricular junction. It is thus possible that the leading
pacemaker site in AVN more frequently originates in the
PNE, which is also part of the slow AVN conduction
pathway. It is possible that in vivo both the PNE and
NH-CS region can generate junctional automaticity. It
would be interesting to test if pacemaker shift exists in
the AVN. Indeed, it cannot be excluded that the dominant
site can shift between the PNE and the NH-CS region
depending on the physiological conditions.
AVN cells have been successfully isolated from rab-
bit (187, 347), guinea pig (539), and mouse (314). Individ-
ual rabbit AVN cells display two different phenotypes:
“ovoid” cells resembling spindle SAN cells and “rod-
shaped” cells. Munk et al. (347) have reported action
potential waveforms having AN, N, and NH properties in
isolated rabbit AVN cells. These authors reported that N
and NH action potential waveforms were typical for ovoid
cells, while AN waveforms were associated with rod-
shaped cellular morphology. It seems a consistent finding
that the majority of ovoid cells are spontaneously active
(347, 539). Some rod-shaped cells are also automatic
(187), but pacemaker activity is present in a more nega-
tive diastolic range (347). As to ionic currents, ovoid cells
express relatively high densities of I
f
and I
Ca,L
(347, 539).
However, rod-shaped cells seem to lack I
f
(187, 347).
F. Automaticity in Purkinje Fibers
The intrinsic conduction velocity of ventricular myo-
cardium is not sufficient to achieve synchronized contrac-
tion of the cardiac chambers. The Purkinje fiber network
ensures a proper propagation of the cardiac impulse along
the ventricular myocardium. Purkinje fibers are very fast
conducting. The upstroke velocity in Purkinje fibers (429)
and cells (72) can approach 1,000 V/s, a value at least
threefold that of ventricular muscle (203). Quick conduc-
tion in Purkinje fibers is due to high expression of both
Na
channels (101, 191) and Cxs (245). Furthermore, the
intercellular resistance to impulse conduction in the Pur-
kinje network of different mammals can be as low as 100
/cm, approaching that of the intracellular milieu (72,
335).
The Purkinje fibers network can pace the heart, in
case of complete atrioventricular block. Automaticity in
Purkinje fibers can become important, because conduc-
tion in the AVN and bundle branches is prone to block
under different genetic and physiopathological condi-
tions. Automaticity in Purkinje fibers is slower than in the
SAN and AVN. Depending on the species considered,
pacemaking rate in Pukinje fibers varies between 25
and 40 beats/min, which is sufficient to set a viable
cardiac output. Action potentials in Purkinje fibers are
similar but not identical to those of ventricular myo-
cardium and display longer action potential duration
and a more negative diastolic potential than SAN and
AVN cells (Fig. 8B) (13, 72, 331, 501). Verkerk et al.
(501) have shown heterogeneity in the duration of re-
polarization in sheep Purkinje cells (PCs). According to
these authors, sheep PCs can display either prominent
phase I repolarization and short-lasting plateau phase
or absence of phase I associated with longer plateau
phase. Early afterdepolarizations (EADs) are often ob-
served in intact Purkinje fibers (13) and isolated PCs
(Fig. 8D) (331). Spontaneous EADs in the Purkinje
network can constitute an important trigger of ventric-
ular arrhythmias (13).
Compared with pacemaker cells from the SAN and
AVN, spontaneously beating PCs have distinct morphol-
ogy and electrophysiological properties (Fig. 8, C and D).
After enzymatic isolation, PCs can be distinguished in
vitro from ventricular and SAN cells by their shape and
size (72, 331). PCs are mainly rod-shaped and larger than
SAN and AVN cells, but generally smaller than ventricular
myocytes (72, 331) (Fig. 8C). PCs have been successfully
isolated from the mouse heart by using a transgenic
mouse line in which the enhanced green fluorescent pro-
tein (EGFP) has been knocked in the gene coding for
Cx40 (Cx40
EGFP/
mouse, see Fig. 8, A and C) (331).
Similar to what is observed in intact fibers, spontaneously
beating PFCs have a large action potential overshoot and
a long-lasting plateau phase (Figs. 1B and 8D). EADs can
also be recorded during repolarization (Fig. 8D), a phe-
nomenon that has been observed also in paced intact
fibers (13). The maximum diastolic potential can be close
to 90 mV in calf Purkinje fibers (127). In contrast, mouse
Purkinje fibers seem to have more positive diastolic po-
tentials (75 mV) (13). The negative diastolic potential in
Purkinje fibers is attributed to I
K1
. Indeed, in calf Purkinje
fibers, I
K1
density is high enough to drive the maximum
diastolic potential close to the equilibrium of K
(72, 118).
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 929
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
Consequently, automaticity needs to be generated in a
more negative voltage interval than in the SAN or AVN.
However, automaticity of Purkinje fibers can be increased
by stretch induced by ventricular filling (233). The pres-
ence of a large I
K1
can partly explain the relatively slow
intrinsic beating rate of Purkinje fibers. Ionic currents
underlying SAN pacemaking have been found also in Pur-
kinje fibers (72, 501, 502). The presence of I
f
has been
extensively described in isolated Purkinje fibers by Di-
Francesco (118, 122) and others (72) as that of I
Ca,T
. Other
ionic mechanisms as the NCX can participate to automa-
ticity in Purkinje fibers (307).
III. MOLECULAR DETERMINANTS OF ION
CHANNELS IN AUTOMATIC HEART CELLS
A. f-Channels
Almost all spontaneously active cells coming from
heart rhythmogenic centers express f-channels (21). Di-
Francesco and co-workers (20, 120, 121) have proposed
that I
f
is the predominant ionic mechanism underlying
cardiac automaticity. SAN pacemaker cells robustly ex-
press I
f
(125, 313). In the rabbit, I
f
density is higher in the
SAN periphery than in the center (355). I
f
can be recorded
also in the myocardium, where it can be activated below
the physiological resting potential (535). It has been
shown that disease states such as atrial fibrillation and
heart failure enhance I
f
expression and shift positively the
current voltage dependence (see Ref. 79 for review). I
f
is
a mixed cationic current carried by Na
and K
(2).
Permeability to Na
is predominant, but K
activates I
f
conductance (115, 122, 125). Ca
2
permeability of f-chan-
nels has been reported in HEK cells expressing recombi-
nant HCN2 channels (538) and in rat ventricular cells
(537). In the SAN and AVN, I
f
activates upon membrane
hyperpolarization with variable threshold between 50
and 65 mV (21). The threshold of I
f
is substantially more
negative in Purkinje fibers (118).
Several factors can influence the activation of f-chan-
nels.
-Adrenergic receptor activation stimulates (21) while
muscarinic agonists inhibit I
f
(129). This regulation is medi-
ated by direct activation of f-channels by cAMP (50, 128),
which facilitates channel opening (126). DiFrancesco and
Tortora (128) failed to observe a direct regulation of SAN
FIG.8.A: structural organization of mouse Purkinje fibers (PF). In this mouse line, the EGFP has been knocked in the gene encoding for Cx40.
Asymmetry of the PF network is then visualized by epifluorescence. In the right ventricle, PF originate from a common trunk (RBB) that ramificates
in secondary branches, which laterally terminate in contact with papillary muscle (APM). PF in the left ventricle arise from the left bundle branches
(LBB) in continuity with the His bundle (HB) and form a dense network on the left ventricular free wall (LVW). Dotted lines indicate the axis of
cutting along the right (RS) and left (LS) ventricular septum to expose the ventricular free walls. B: evoked action potentials recorded using the same
preparation as in A. Records are from the ventricular working myocardium (VWM), the right (RBB) and left (LBB) bundle branches. Note the longer
action potential plateau phase in the bundle branches composed by Purkinje fibers. C: line shows the morphology of a spontaneously active mouse
Purkinje cell under visible light (right panel) and EGFP epifluorescence (left panel). D: automaticity and action potential configuration in an isolated
Purkinje cell as shown in C. Spontaneous early afterdepolarizations (EADs) are often observed during the action potential plateau phase. [From
Miquerol et al. (331), with permission from Elsevier.]
930 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
f-channels by purified protein kinase A (PKA) (128). In SAN
cells, direct regulation of voltage dependence by cAMP can
quantitatively account for I
f
regulation by autonomic ago-
nists (126). The existence of a PKA-mediated regulation of I
f
has been suggested by Chang et al. (80) in canine Purkinje
cells and in murine stem cells derived from cardiomyocytes
(1). Intracellular Ca
2
have been reported to positively reg-
ulate I
f
by either increasing current conductance (171) or by
positively shifting the current voltage dependence (416). The
mechanisms by which Ca
2
regulate I
f
has not been eluci-
dated. Inside-out patch-clamp experiments seem to exclude
a direct Ca
2
effect on f-channels (540).
Four gene isoforms, named HCN1–4, code for f-chan-
nels and have been cloned from the mouse (295, 435),
rabbit (217), and human (296, 444). These isoforms dis-
play different activation threshold, kinetics, and sensitiv-
ity to cAMP (234). Particularly, HCN1 channels show the
more positive threshold for activation, the fastest activa-
tion kinetics, and the lowest sensitivity to cAMP (295, 435),
while HCN4 channels are slowly gating and strongly sensi-
tive to cAMP (217, 296, 444). HCN2-mediated channels have
intermediate properties between HCN1- and HCN4-medi-
ated channels (234). It has been proposed that the trans-
membrane protein KCNE2 stimulates the surface expres-
sion and accelerates the kinetics of HCN channels in heter-
ologous expression system (536) and in cardiomyocytes
(404). It has also been proposed that KCNE2 can switch
HCN2-mediated currents from time dependent to time
independent (302).
HCN4 is the predominant f-channel isoform ex-
pressed in the SAN (317, 342, 450) and AVN (317). The
HCN1 and HCN2 isoforms are also expressed in these
rhythmogenic centers (317, 342, 344, 450).
However, the molecular composition of native car-
diac f-channels has not been elucidated. The strong ex-
pression of HCN4 mRNA in the SAN and AVN and the high
sensitivity of native f-channels to cAMP (126) suggest that
HCN4 is a major determinant of native I
f
. Remarkably,
native I
f
has faster activation kinetics than I
f
mediated by
HCN1-HCN4 heterotetramers. Such a difference cannot
be accounted for the presence of basal cAMP or KCNE2 in
intact SAN cells (7). Recent studies have described a
“context-dependent” regulation of native f-channels. Qu
et al. (403) overexpressed HCN2 channels in neonatal and
adult ventricular myocytes and showed that I
f
voltage
dependence was dependent on the developmental state of
the cells. Furthermore, transfection of HCN2 and HCN4
channels in HEK cells and neonatal ventricular myocytes
yields I
f
currents that activated more positively in cardi-
omyocytes than in HEK cells, irrespectively of the trans-
fected isoform (402). These observations indicate that
cellular factors other than cAMP and KCNE2 contribute
to the native SAN I
f
kinetics and voltage dependence.
Barbuti et al. (19) have reported that in rabbit SAN cells,
HCN4 channels interact with caveolin-3 and are concen-
trated in caveolar membrane lipid rafts. Chemical disor-
ganization of caveolar structures positively shifted the I
f
activation curve and slowed the current deactivation ki-
netics (19). In a follow-up study, the same group showed
that
2
-receptors colocalize with f-channels in cavelolar
structures and that
2
-dependent regulation of I
f
also
requires caveolar membrane compartimentalization (20).
It thus seems very likely that cAMP-dependent regulation
of f-channels takes place locally, in a subcellular space
delimited by lipid rafts. Pian et al. (392) have reported that
phosphatidylinositol 4,5-bisphosphate (PIP
2
) prevents
rundown of both recombinant HCN2-mediated and native
rabbit SAN I
f
. It is thus possible that a multiplicity of
factors other than cAMP can participate in the regulation
of I
f
in the SAN. Regulation of HCN2 channels by stretch
has been reported by Li et al. (287). These authors have
studied HCN2-mediated I
f
expressed in Xenopus oocytes.
In this preparation, stretch accelerated both activation
and deactivation kinetics of I
f
(287).
B. Voltage-Dependent Ca
2
Channels
Voltage-dependent Ca
2
channels (VDCCs) are an
important pathway of Ca
2
entry in pacemaker cells. L-
and T-type VDCCs have been consistently recorded in
spontaneously active SAN and AVN cells (145, 172, 314,
401, 539). L-type VDCCs are expressed throughout the
myocardium, are sensitive to dihydropyridines (DHPs)
such as nifedipine and BAY K 8644, and are stimulated by
PKA-dependent phosphorylation (491). For further dis-
cussion on the physiology and pharmacology of cardiac
L-type VDCCs, the reader is referred to some recent re-
views (470, 491). In the heart, L-type VDCCs initiate con-
traction (475) and contribute to pacemaker activity (309,
497, 546).
I
Ca,L
in the SAN and AVN activates upon depolariza-
tion at a variable threshold between 50 and 30 mV
(309, 314, 497, 539, 546) and displays Ca
2
- and voltage-
dependent inactivation (172, 311). Expression of I
Ca,L
in
the rabbit SAN is heterogeneous; larger cells (presumably
from the SAN periphery) express less current density than
smaller cells (349). In the SAN, I
Ca,L
is regulated by PKA
(390) and by activated Ca
2
/calmodulin-dependent pro-
tein kinase II (CaMKII) (509), which regulates the current
activation and reactivation kinetics (509). Similarly, in
ventricular cells, CaMKII facilitates opening of L-type
channels in response to Ca
2
permeation, with conse-
quent stimulation of I
Ca,L
amplitude and slowing of cur-
rent inactivation (135, 528).
L-type VDCCs are multisubunit complexes, consti-
tuted by a pore-forming
1 subunit in association with
different accessory subunits (
2-
,
, and
) (470). Four
L-type
1-subunits have been cloned and classified in the
Ca
v
1 gene family (78). The Ca
v
1.1
1-subunit is responsi-
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 931
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
ble for excitation-contraction coupling in skeletal muscle
(474). Ca
v
1.4 is expressed in the retina, spinal cord, and
immune cells (324). Ca
v
1.2 and Ca
v
1.3 are expressed in
neurons, as well as in cardiovascular and neuroendocrine
cells (77). In the ventricle, the Ca
v
1.2
-subunit is the
predominant molecular determinant of I
Ca,L
. This
-sub-
unit is expressed in the SAN, AVN, and the atria (45, 309,
317, 472). Recombinant and native Ca
v
1.3-mediated I
Ca,L
displays a more negative activation threshold and slower
inactivation kinetics than Ca
v
1.2-mediatd I
Ca,L
.Ca
v
1.3
channels have also reduced sensitivity to DHPs (254). In
contrast, SAN Ca
v
1.2-mediated and Ca
v
1.3-mediated I
Ca,L
seem to be similar as to their sensitivity to
-adrenergic
agonists (309).
T-type VDCCs are activated at more negative volt-
ages than L-type VDCCs. The kinetic hallmark of native
and heterogously expressed I
Ca,T
is slow activation and
fast voltage-dependent inactivation. The single-channel
conductance of T-type VDCCs is also smaller than that of
L-type VDCCs (388). Three genes coding for T-type
1-
subunits have been cloned (103, 339, 340, 387) and named
Ca
v
3.1, Ca
v
3.2, and Ca
v
3.3. The Ca
v
3.1 isoform is function-
ally expressed in neonatal rat atrial myocytes (282), in the
AT-1 cell line (438), in the developing mouse heart (103),
and in mouse SAN and AVN (314). In the adult SAN,
expression of the Ca
v
3.1 isoform is higher than that of
Ca
v
3.2 (45, 314). The Ca
v
3.2 isoform has been cloned from
a human heart library (103), and possibly constitutes the
predominant I
Ca,T
isoform in the rat embryonic heart
(146). In contrast, expression of Ca
v
3.1 channels in-
creases during the perinatal period and is maximal in
adulthood (358). To date, expression of the Ca
v
3.3 iso-
form has not been found in the myocardium, the SAN, and
AVN (314, 340). No specific drugs able to discriminate
between Ca
v
3.1 and Ca
v
3.2 channels are presently avail-
able. However, Ca
v
3.2 channels are much more sensitive
to Ni
2
than Ca
v
3.1 channels, the concentration for half-
block (EC
50
) being 5 and 150
M for Ca
v
3.2 and Ca
v
3.1
channels, respectively (265).
C. St-Channels
The I
st
current has been recorded by Guo et al. (166)
in rabbit SAN and AVN (165). This current has also been
found in SAN cells from guinea pig, rat, and mouse (86,
457). I
st
activates at about 70 mV, peaks at about 50
mV, and is positively regulated by
-adrenergic agonists
(166). I
st
is carried by Na
, but is clearly distinct from I
Na
,
since it is insensitive to tetrodotoxin (TTX), blocked by
DHP antagonists, and inhibited by divalent cations such
as Ca
2
,Mg
2
, and Ni
2
(166). Single st-channels have a
unitary conductance similar to that of L-type VDCCs (13
pS) and are facilitated by BAY K 8644 (333). Mitsuye et al.
(334) have reported that I
st
expression is restricted to
spontaneously active SAN and AVN cells, since no I
st
is
recorded in quiescent cells from these two regions.
The molecular determinants of I
st
have not yet been
identified. I
st
could be mediated by a new subtype of
L-type VDCCs (166), or by a still unidentified splice vari-
ant. This hypothesis would be consistent with the phar-
macological sensitivity of I
st
to agonist and antagonist
DHPs.
D. Voltage-Dependent Na
Channels
A significant fraction of rabbit SAN pacemaker cells
expresses I
Na
in culture (348, 351). I
Na
is heterogeneously
expressed in the adult rabbit SAN (204, 250). Honjo et al.
(204) reported that I
Na
expression is higher in large cells
(presumably from the SAN periphery) than in small cells
(from the SAN center). A high percentage of small cells is
devoid of I
Na
(204). Consistently, pacemaking in SAN
periphery is sensitive to TTX, while automaticity in tissue
balls from the SAN center is insensitive to TTX applica-
tion (250, 255). In contrast, SAN cells of newborn rabbits
show high I
Na
expression (24). In the newborn, expres-
sion of I
Na
is maximal during the first 3 wk after birth and
then declines irrespectively of cell size (24). High expres-
sion of I
Na
is present also in adult mouse SAN cells (279,
312, 313). The SAN I
Na
is more sensitive to TTX than I
Na
of the working myocardium. TTX-sensitive I
Na
has been
reported in rabbit neonatal (25) and in adult mouse SAN
cells (279). TTX in the nanomolar range reduces the beat-
ing rate of isolated mouse hearts (304) and SAN pace-
maker cells (279).
Na
channels are related to a large family of genes
coding for 10
-subunit isoforms (see Ref. 534 for review).
The electrophysiological properties of the
-subunit can
be modulated by accessory
1- and
2-subunits (534). By
using in situ hybridization, Baruscotti et al. (25) have
shown that the newborn SAN expresses the neuronal
TTX-sensitive Na
v
1.1
-subunit isoform. Expression of
both the TTX-sensitive Na
v
1.1 and TTX-resistant “cardiac”
Nav1. 5 isoforms has been reported in the mouse SAN at
the protein level (279).
E. Voltage-Dependent K
Channels
Voltage-dependent K
channels (VDKCs) control the
action potential repolarization phase in spontaneously
active cells (58) and in the working myocardium (353,
422). In SAN and AVN, three VDKC-mediated currents
have been recorded: the fast (I
Kr
) and slow (I
Ks
) delayed
rectifiers and the transient outward current (I
to
).
I
Kr
is activated upon depolarization from a threshold
of 50 mV. In voltage-clamp experiments, I
Kr
fully acti-
vates at about 20 mV and displays strong inward recti-
fication (93, 218, 496). At positive voltages, inactivation of
932
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
I
Kr
counterbalances activation, thereby generating a re-
gion of negative slope conductance (93). When the mem-
brane voltage is switched back to negative potentials, I
Kr
slowly deactivates to generate tail currents, which contrib-
ute to set the maximum diastolic potential (see sect.
VI). Tail
currents are due to channel reopening followed by time-
dependent closure at negative membrane potential (452).
Shibasaki (452) has reported that the single-channel con-
ductance of I
Kr
channels in 150 mM K
is 11 pS, giving an
estimate of 1,000 channels in a typical rabbit SAN cell.
I
Kr
is sensitive to class III methanesulfonanilide com-
pounds E-4031 and dofetilide (UK-68798), as well as to a
plethora of unrelated compounds used in medical prac-
tice (473).
I
Kr
channels are coded by the ERG gene family,
which includes three members named ERG1-ERG3. The
ERG1 gene is expressed in the heart (398). Mutations in
the human ERG1 gene or in its accessory subunit KCNE2
can impair ventricular repolarization and lead to long-QT
syndrome (92, 236). In the mouse SAN, mRNAs of the
three known ERG1 splicing variants 1a, 1a, and 1b have
been found (93). However, the composition of native I
Kr
channels in SAN, AVN, and Purkinje fibers has not yet
been elucidated. Lees-Miller et al. (273) have proposed
that both ERG1a and -1b are able to form recombinant
channels with properties similar to those of native I
Kr
, but
to date, only ERG1a immunoreactivity has been identified
in the ventricle (397).
The I
Ks
current is kinetically and pharmacologically
distinct from I
Kr
. I
Ks
has slower activation and faster
deactivation kinetics than I
Kr
. I
Ks
is sensitive to the 293B
channel blocker, which is currently employed for study-
ing the physiological role of this current (275, 278). Ex-
pression of I
Kr
and I
Ks
in the rabbit SAN is heterogeneous.
Indeed, high I
Kr
and I
Ks
densities have been recorded in
large SAN cells, while smaller cells seem to express only
I
Kr
(278). Species-dependent differences in I
Kr
and I
Ks
expression seem to exist. I
Kr
has not been recorded in
isolated pig SAN cells, which express only I
Ks
(372). I
Ks
has not yet been directly recorded in mouse SAN cells
(Mangoni, unpublished observations). However, Temple
et al. (478) have reported expression in the mouse SAN of
the KCNQ1 accessory subunit KCNE1. On the other hand,
mice lacking KCNE1 have susceptibility to spontaneous
atrial fibrillation, but do not show alterations in SAN
pacemaker activity (478).
Three genes named KCNQ1-KCNQ3 encode for I
Ks
,
but only KCNQ1-related expression is found in the heart
(352). The sensitivity of I
Kr
and I
Ks
to
-adrenergic ago-
nists in pacemaker cells would need further investigation.
Voltage-clamp studies on intact SAN tissue preparations
have reported stimulation of I
K
by epinephrine (62).
These observations have been confirmed in isolated rab-
bit SAN cells (277). Sensitivity of I
Kr
to
-adrenergic
agonists has been reported in ventricular myocytes (194),
but this issue has not been addressed in SAN cells.
The I
to
current is characterized by fast activation and
inactivation kinetics (see, for example, Ref. 277). Pharma-
cologically, I
to
can be identified thanks to its sensitivity to
4-aminopyridine (4-AP) (422). In working myocytes, at
least two components of I
to
have been identified and
named I
to,f
and I
to,s
, according to their fast and slow
inactivation kinetics, respectively (353). Expression of I
to
in rabbit SAN cells is heterogeneous (60, 206, 277). In-
deed, pacemaker activity of SAN tissue balls isolated from
the SAN periphery is more sensitive to block of I
to
by 4-AP
than that from the center of the SAN (60). In rabbit SAN
cells, I
to
density is positively correlated with the cell size
(206, 277).
I
to
channels are coded by the K
v
1 and K
v
4 gene family
(354). The composition of native I
to
channels in the SAN
and conduction system is not yet known. In the mouse
heart, inactivation of Kv1.4 channels abolishes I
to,s
(291).
The channel complex mediating the I
to,f
component is
formed by heteromultimers of K
v
4.2 and K
v
4.3 channel
subunits (168). The KChIP2 protein is also associated
with the I
to
channel complex (8). KChIP2 regulates cur-
rent inactivation and channel targeting to the cell mem-
brane (8). Ventricular cells isolated from mice lacking
KChiP2 have no I
to
and display prolonged action potential
duration. Episodes of ventricular tachyarrhythmias are
recorded in KChiP2 knockout mice (260).
In conclusion, I
to
appears to play a role in the action
potential repolarization phase of peripheral SAN cells, but
is probably less important in the SAN center. However,
mice lacking both I
to,s
and I
to,f
have no apparent dysfunc-
tion of the SAN rhythm (290). Further studies will be
needed to elucidate the subunit composition and the
physiological role of I
to
in spontaneously active cells.
F. G Protein-Activated, ATP-Dependent,
and Inward Rectifier K
Channels
The acetylcholine-activated current (I
KACh
)is
strongly expressed in the SAN, atria, and AVN (124, 159,
367). In these tissues, I
KACh
is activated by muscarinic and
adenosine receptors by direct binding of G protein
␤␥
-
subunits to the I
KACh
channel complex (for review, see
Refs. 520, 521). Four genes named Kir3.1-Kir3.4 underlie
I
KACh
channels (520). Kir3.1 and Kir3.4 are expressed in
the heart (521). Cardiac I
KACh
channels are tetrameric
complexes containing Kir3.1 and Kir3.4 channels (521).
However, mice lacking Kir3.4 channels have no cardiac
I
KACh
(522), because Kir3.1 channels require Kir3.4 chan-
nels to be targeted at the cell membrane (238).
ATP-dependent K
channels have been described in
rabbit SAN cells (184). These channels underlie I
K,ATP
and
are open when the intracellular concentration of ATP is
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 933
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
lowered. Van Wagoner (493) has shown that I
K,ATP
chan-
nels are activated by stretch. Cell swelling by a hypotonic
solution was used in this work to demonstrate that I
K,ATP
can generate a significant whole cell current under me-
chanical stress. Activation of I
K,ATP
by stretch has been
reported also in ventricular cells (251), but these obser-
vations have not yet been confirmed in pacemaker tissue.
In rabbit SAN cells, I
K,ATP
slows pacemaker activity and
hyperpolarizes the cell maximum diastolic potential
(184). Han et al. (184) have proposed that the slowing of
heart rate induced by I
K,ATP
can be important for protect-
ing the myocardium from ischemic damage or stunning.
This interesting hypothesis has not yet been experimen-
tally verified.
Rabbit SAN and AVN cells express very low densities
of inward-rectifier K
current I
K1
(62, 215, 350, 363, 366).
In contrast, I
K1
is present in the working myocardium
(422) and in Purkinje fibers (118). I
K1
is responsible for
the negative diastolic potential of PFCs. However, I
K1
can
be moderately expressed in rat (457) and mouse (86) SAN
cells, even if at a much lower density than the working
myocardium. In the heart, I
K1
channels are coded by the
Kir2.1 and Kir2.2 channel subunits (352).
G. Cl
Channels, Volume-Activated Channels,
and Stretch-Activated Cationic Channels
The first evidence for the presence of chloride cur-
rent (I
Cl
) in spontaneously active cells came from Pur-
kinje fibers. In this preparation, De Mello (106) reported
that I
Cl
represents a substantial fraction of the total mem-
brane conductance during the action potential. I
Cl
has
been reported by Seyama who estimated that Cl
are
responsible for 9% of the total membrane conductance
and that I
Cl
underlies a significant part of the membrane
inward-going rectification in the SAN (449). Hagiwara
et al. (174) have characterized a stretch-activated back-
ground I
Cl
in rabbit SAN cells. This current is distinct
from the voltage- and Ca
2
-dependent, cAMP-sensitive
[I
Cl(Ca)
] present in ventricular myocytes (16, 471), and
probably belongs to the family of volume-activated anion-
selective currents (VAC
Cl
) (27), because the current was
activated by cell inflation (174). Hagiwara et al. (174)
reported that SAN background I
Cl
is insensitive to intra-
cellular Ca
2
depletion and cAMP. Bescond et al. (37)
have reported that angiotensin II (ANG II) activates back-
ground I
Cl
in SAN cells via a protein kinase C (PKC)-
dependent signaling pathway. Verkerk et al. (504) have
tested the possibility that I
Cl(Ca)
is present in SAN cells
and its possible role in the action potential repolarization.
They recorded I
Cl(Ca)
in about one-third of SAN cells
tested. I
Cl(Ca)
activated from about 20 mV, peaked be-
tween 30 and 40 mV, and was augmented by norepi-
nephrine. The kinetic behavior of SAN I
Cl(Ca)
(504) was
similar to I
Cl(Ca)
in atrial, Purkinje, and ventricular myo-
cytes (210). Verkerk et al. (504) have tested the role of
I
Cl(Ca)
in SAN automaticity, by employing action potential
clamp experiments and numerical modeling (504). They
found that I
Cl(Ca)
activates late during the action potential
upstroke phase and gives moderate contribution to the
repolarization phase. No contribution of I
Cl(Ca)
to the
maximum diastolic potential or diastolic depolarization
rate was observed (504).
VAC
Cl
are activated by an increase in cell volume or
by agents that alter membrane tension and mechanical
stretch (27). VAC
Cl
open with a quite slow response to cell
volume change (464). I(VAC
Cl
) is time independent, out-
wardly rectifying, and reverses between the cell resting
potentials and action potential plateau phase. Conse-
quently, VAC
Cl
shortens the action potential duration and
depolarizes the membrane resting potential, thereby act-
ing to decrease cell volume (27). In the heart, VAC
Cl
play
a role in the ischemic response and seem to be overex-
pressed in the hypertrophied myocardium (27). The exis-
tence of mechanosensitive VAC
Cl
in SAN has been di-
rectly proposed firstly by Arai et al. (14), who showed that
the positive chronotropic response induced by strong
stretch stimuli in SAN tissue is inhibited by Cl
channel
blockers. However, Cooper and Kohl (99) have failed to
observe an effect of a Cl
channel blocker under condi-
tions of moderate mechanical load.
SAC are cationic nonselective or K
selective. Com-
pared with VAC
Cl
, SAC activate fast upon mechanical
stimulation. Kohl et al. (251) have suggested that fast
activation of SAC can be involved in the beat-by-beat
regulation of heart rate. SAC are inhibited by Gd
3
(383),
streptomycin (151), and the spider toxin GsMTx-4 (42).
Cooper and Kohl (99) reported that the chronotropic
response of isolated pacemaker cells is inhibited by
GsMTx-4, thus indicating functional expression of SAC in
SAN cells.
IV. PUMPS AND EXCHANGE CURRENTS
A. The Na
-K
Pump Current I
p
The electrogenic role of Na
_
K
pump is well estab-
lished. Under physiological ionic concentrations, three
Na
are extruded and two K
are transported in the
intracellular milieu for each pump cycle. Consequently,
the Na
_
K
pump generates a net outward current that
influences cellular pacemaking. In Purkinje fibers, I
p
stim-
ulation hyperpolarizes the membrane potential and slows
spontaneous activity (149). Noma and Irisawa (362) have
recorded I
p
-mediated membrane hyperpolarization in rab-
bit SAN multicellular preparations, by rapidly switching
perfusion from K
-free Tyrode’s solution to one contain-
ing 5.4 mM K
(362). A similar hyperpolarization of the
934
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
membrane potential has been observed in atrioventricular
cells (261). I
p
properties can be studied by varying the
intracellular K
concentration after block of time- and
voltage-dependent currents, according to the protocol de-
scribed by Gadsby et al. for ventricular cells (150). I
p
can
also be isolated pharmacologically by employing ouabain
or strophantidine (261, 362, 431). Sakai et al. (431) have
studied the properties of I
p
in isolated rabbit SAN cells.
They reported that I
p
carries a significant steady-state
outward current in physiological extracellular K
and 50
mM intracellular Na
. Similarly to I
p
recorded in ventric-
ular cells (150), SAN I
p
displays voltage-dependent behav-
ior (431). I
p
voltage dependence is evident in the pace-
maker range with a relative half activation voltage of
about 50 mV. Furthermore, I
p
is very sensitive to
changes in intracellular Na
so that the current doubles
when changing Na
from about 15 to 25 mM Na
(431).
In SAN cells, I
p
contributes to the setting of the
maximum diastolic potential in the range of 60 mV
(362). In this respect, Sakai et al. (431) have estimated
that the net outward I
p
-mediated current in intact SAN
cells is 14 pA in the experiments by Noma and Irisawa
(362) and 20 pA in their recordings. However, since
automaticity is essentially a periodical process, it is also
possible that I
p
varies cyclically during the pacemaker
cycle according to its intrinsic voltage and Na
depen-
dency. Furthermore, one may wonder if I
p
can respond to
transient changes in intracellular Na
induced by activa-
tion of ion channels such as I
Na
and I
f
. Choi et al. (90)
have measured changes in intracellular Na
activity (a
i
Na
)
in multicellular and isolated SAN cell preparations under
application of isoproterenol, carbachol, and I
f
blockers.
These authors reported that isoproterenol-induced accel-
eration of the pacing rate was accompanied by an eleva-
tion of a
i
Na
. Application of Cs
and ZD-7288 reduced (but
did not completely abolished) the rise in a
i
Na
, thus sug-
gesting a link between I
f
stimulation by isoproterenol and
a
i
Na
. In contrast, both carbachol and ZD-7288 reduced a
i
Na
in spontaneously beating myocytes but had no effect on
quiescent cells. These data are strongly suggestive of a
significant influence of I
f
on intracellular a
i
Na
. It is thus
tempting to speculate that f-channels can be functionally
coupled to I
p
in SAN pacemaker cells. This hypothesis can
be particularly relevant when considering that f-channels
are concentrated in membrane lipid rafts. It is possible
that opening of f-channels induces a highly localized rise
in intracellular Na
concentration. The presence of the
Na
_
K
pump near f-channels is important for maintain-
ing the ionic homeostasis.
B. The Na
-Ca
2
Exchanger Current I
NCX
and the
Na
-H
Exchanger
The NCX is a major actor intervening in cardiac cell
Ca
2
homeostasis. Indeed, Ca
2
entry during the diastole
and action potential upstroke increases intracellular Ca
2
concentration ([Ca
2
]
i
) stimulating NCX activity. Stoichi-
ometry of Na
-Ca
2
transport through NCX is electro-
genic, one Ca
2
is extruded for three Na
entering the cell
(454). The result is a net inward current (I
NCX
) which
depolarizes the membrane voltage. In the working myo-
cardium, NCX is responsible for Ca
2
efflux during action
potential repolarization (35, 142). In the rabbit myocar-
dium, I
NCX
extrudes 30% of the total Ca
2
required to
activate contraction, while the sarcoplasmic reticulum
Ca
2
-ATPase (SERCA) actively removes Ca
2
from the
cytoplasm to replenish the SR (142). Regardless of the
species considered, Ca
2
extrusion by NCX quantitatively
matches Ca
2
influx triggering Ca
2
-induced Ca
2
release
(CICR), while SERCA transport must match ryanodine
receptor (RyR)-dependent Ca
2
release. The ventricular
I
NCX
contributes to action potential plateau as well as to
contractile relaxation (35). Positive regulation of NCX
activity by
-adrenergic receptor activation is linked to
positive inotropism in ventricular and Purkinje myo-
cytes (35).
Early evidence of the presence of I
NCX
in pacemaker
tissue was obtained by Brown et al. (65) using multicel-
lular SAN preparations. These authors showed that the
slow inward current had a second late component that
could be attributed to Na
-Ca
2
exchange. Similarly,
Zhou and Lipsius (547) recorded I
NCX
in cat latent atrial
pacemaker cells by applying depolarizing voltage steps
eliciting I
Ca,L
which, in turn, activated I
NCX
. NCX activity
closely follows changes in [Ca
2
]
i
. For example, in cane
toad (Bufo marinus) sinus venosus cells, an increase in
I
NCX
precedes the measured systolic Ca
2
transient, thus
suggesting that I
NCX
is significantly stimulated by an in-
crease in subsarcolemmal [Ca
2
]
i
(228). Furthermore, in-
hibition of NCX activity by Ni
2
slows the decline of the
caffeine-induced release, indicating that NCX is a major
controller of [Ca
2
]
i
in toad sinus venosus cells (228). The
physiological role of I
NCX
in pacemaking can be tested by
fast removal of extracellular Na
or substitution with Li
.
In these conditions, [Ca
2
]
i
quickly rises and pacemaker
activity stops (44, 228, 433). Intact NCX activity is thus
necessary for maintaining automaticity of pacemaker
cells in both amphibians (228) and mammals (44, 433).
The importance of NCX in the generation of cardiac au-
tomaticity will be discussed in a following section of this
review.
The Na
-H
exchanger (NHE) is a pH-regulatory
protein present in the plasma membrane of all cardiac
myocytes. In response to intracellular acidosis, the pro-
tein removes one intracellular H
in exchange for an
extracellular Na
(305). NHE is regulated by several hu-
moral factors including ANG II and endothelin (305).
NHE1 is the predominant isoform expressed in cardiac
myocytes (305). In isolated ferret hearts, NHE1 is respon-
sible for 50% of the total H
efflux (163). NHE1 has been
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 935
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
implicated in cardiac pathological states including ische-
mia (see Refs. 198, 305 for review). Indeed, strong activa-
tion of NHE1 following intracellular acidosis leads
to increased intracellular Na
concentration and conse-
quent [Ca
2
]
i
build up through activation of I
NCX
(198).
We can thus expect that NHE and, more generally, pro-
teins involved in H
handling (198) can be involved in the
SAN chronotropic response to ischemia and acidosis (see
sects.
IXE and XB). However, we currently lack a detailed
characterization of NHE-mediated current in SAN.
V. PATTERNS OF GENE EXPRESSION IN
ADULT PACEMAKER TISSUE
In this short section, we review evidence on the
differential expression pattern of ion channels in the
heart’s rhythmogenic centers and in working myocar-
dium. Han et al. (179) have studied the expression of ion
channels potentially involved in automaticity in canine
Purkinje fibers and compared it with ventricular myocar-
dium. This study combined RT-PCR with Western blotting
of membrane proteins. Results have shown that HCN4
is highly expressed in canine Purkinje fibers, while it
seemed undetectable in ventricular muscle. HCN2 expres-
sion was comparable in Purkinje fibers and the ventricle.
Expression of Ca
v
1.2 and NCX was higher in the ventricle
than in Purkinje fibers. Genes coding for I
Ca,T
(Ca
v
3.1,
Ca
v
3.2, and Ca
v
3.1) were all more expressed in Purkinje
fibers. Similar peculiarities in the pattern of expression of
Purkinje fibers have been recently reported by Gaborit
et al. (148) in nondiseased human heart.
Marionneau et al. (317) have employed a high-
throughput RT-PCR approach (112) to assess the ion
channel expression pattern in the mouse SAN and AVN.
Indeed, real-time RT-PCR array technology allows han-
dling of small tissue samples that contain low quantities
of mRNA. These authors have compared the expression
of 71 channels and related genes in the SAN, AVN, right
atrium, and left ventricle (Fig. 9A). Transcripts coding for
HCN1 and HCN4 show the highest expression level in the
SAN. The SAN and AVN are distinguished by high expres-
sion of Ca
v
1.3 and Ca
v
3.1 mRNAs. HCN2 mRNA is present
in the SAN, but at lower levels than HCN1 and HCN4
mRNA. The AVN is characterized by expression of Nav1.7,
Kv1.6, as well as the K
-subunit Kv
1. These transcripts
appear to be specific of the AVN (317). Significant expres-
sion of Nav1.1 has been found in the AVN. The Ca
v
2
2
may constitute a potential accessory subunit for VDCCs in
the SAN and AVN, because mRNA encoding for this sub-
unit is highly expressed in the nodes and in Purkinje
fibers. Mouse pacemaker tissues show predominant ex-
pression of the Na
channel
-subunits Nav
1 and Nav
3.
It is possible that the association of Nav
1 and Nav
3to
Nav
-subunits speeds up inactivation of I
Na
. Finally,
FIG.9.A: a combined quantitative PCR and two-way hierarchical
clustering approach is used to compare the expression pattern in six
mouse SAN (SAN1– 6), AVN (AVN1– 6), atria (A1– 6), and ventricles
(V1–6). Seventy-one mRNAs encoding ion channel subunits and Ca
2
handling proteins are considered. The expression of each gene is rep-
resented by a colored row, and each column defines a tissue sample. A
false color scale starting from dark green (lowest expression) to bright
red is applied. Four gene clusters (A–D) are shown at the left of the
panel. In the figure, cluster B identifies genes that are highly expressed
in working myocardium, while cluster C identifies groups of genes that
are highly expressed in the SAN and AVN. [From Marionneau et al. (317),
with permission from Wiley-Blackwell.] B: quantitative RT-PCR analysis
of 10 representative gene transcripts in rabbit central SAN (blue), pe-
ripheral SAN (green), right atrium (red), and SAN ring bundle (RSARB,
purple). Symbol (⫹⫹) stands for abundant expression, () stands for
expression, (/) indicates expression in some cells, and () indicates
absence of detectable expression. [From Tellez et al. (477).]
936 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
Kv1.1, Kv1.4, Kv1.6, Kv
1, and Kv
3 are highly expressed
in the SAN and AVN (317). Electrophysiological investi-
gation of mouse SAN and AVN cells will be needed to
elucidate the functional role of these K
channels in
pacemaking and conduction.
By combining quantitative RT-PCR and immunohis-
tochemistry, Tellez et al. (477) have investigated the re-
gional expression of genes involved in automaticity and
conduction in the rabbit central and peripheral SAN. The
atrial muscle was used as a reference of nonpacing myo-
cardium (Fig. 9B). In this work, a quantitative switch
between channel isoforms has been highlighted: moving
from the atrium to the SAN center, Ca
v
1.2 is substituted
by Ca
v
1.3, Nav1.5 by Nav1.1, and Kv1.4 by Kv4.2. As to
RyRs, RyR2 is gradually downregulated, while expression
of RyRs augmented in the SAN center. The SAN periphery
shows an intermediate expression pattern between atrial
muscle and central SAN (477).
In conclusion, cells endowed with automaticity are
characterized by enhanced expression of some channels
including HCN4, Ca
v
1.3, and Ca
v
3.1. Furthermore, the
central rabbit and mouse SAN express the TTX-sensitive
Na
v
1.1 subunit. Differences in the expression pattern of
ion channels in spontaneously active tissues versus the
working myocardium are essentially quantitative, rather
than qualitative. Numerical modeling work has suggested
that quantitative changes in ionic channels can induce
pacemaking (264). Overexpression of HCN (143) channels
or silencing of I
K1
(327) has been shown to induce auto-
maticity in myocardial cells. However, pacemaker cells
have also a peculiar morphological phenotype. It is thus
likely that factors other than ion channels contribute to
the specific phenotype of primary SAN pacemaker cells.
VI. GENESIS OF CARDIAC AUTOMATICITY:
MECHANISMS OF PACEMAKING
A. Concepts
In the following sections, we will review the current
knowledge on the ionic and intracellular signaling mech-
anisms underlying the generation of pacemaker activity.
We aim to discuss published evidence on how different
families of ion channels, pumps, exchangers, and intra-
cellular Ca
2
signaling generate and regulate pacemaking.
Many research groups have focused on primary SAN
pacemaker activity so that the bulk of evidence discussed
here is obtained from isolated pacemaker cells or intact
SAN tissue. There is currently no general agreement as to
which mechanism is necessary for initiating automaticity
or, in other words, which ionic and/or cellular mechanism
specifically confers automaticity to pacemaker cells.
Different groups have proposed alternative views of
the mechanisms underlying pacemaker activity and em-
phasized the importance of in situ catecholamines pro-
duction (136, 395, 396), ionic currents such as I
f
(120,
121), I
Ca
(172, 309), I
Kr
(93, 371), I
st
(334), or spontaneous
diastolic Ca
2
release (307).
From a conceptual point of view, searching for pace-
maker mechanisms requires the description of the elec-
trophysiological, signaling, and transcriptional processes
that are active (or specifically inactive) in pacemaker
cells. These mechanisms underlie the morphological and
functional differences between spontaneously active cells
and contractile myocytes and/or would be responsible for
the commitment of cardiac precursors toward the pace-
maker function in the adult SAN and in the cardiac con-
duction system. Knowledge of the pacemaker mecha-
nisms would help confer automaticity to target regions of
the working myocardium or to undifferentiated cardiac
precursors. Such a concept has now been applied with the
aim to create “biological pacemakers” in stem cells or in
the cardiac conduction system by in situ gene transfer
(see sect.
XI). We will first discuss evidence involving ion
channels in the generation of automaticity and then re-
view recent work showing that pacemaking also involves
the activity of NCX and spontaneous diastolic Ca
2
re-
lease. I
f
is considered as a key ion channel underlying
automaticity. The relevance of I
f
in the genesis of pace-
maker activity is supported by different lines of electro-
physiological, pharmacological, and genetic evidence.
The importance of spontaneous Ca
2
release in the gen-
eration of pacemaker activity is indicated in experiments
involving pharmacological inhibition of RyRs in rabbit
SAN cells. When reviewing the current knowledge of the
I
f
-based pacemaking and that mediated by spontaneous
Ca
2
release, we will try to highlight differences as well as
similarities between these two proposals. Particularly,
these views are not mutually exclusive and, more impor-
tantly, experiments performed in other laboratories (in-
cluding ours) from genetically modified mouse strains
lacking Ca
v
1.3, Ca
v
3.1, and HCN channels indicate that
the generation of automaticity requires the intervention of
more than one individual mechanism. Also, it has been
shown that the influence of a given ion channel or RyRs
on pacemaker activity can vary regionally in the SAN.
B. Ion Channels and Cardiac Automaticity:
General Considerations
Automaticity in pacemaker cells of the adult heart
stops when the cell membrane potential is depolarized by
addition of KCl in the extracellular solution. Automaticity
in adult pacemaker cells thus differs from that of early
embryonic myocytes which maintain spontaneous con-
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 937
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
tractions even at depolarized membrane potentials (326,
505). This observation demonstrates that mature pace-
maker cells require voltage-dependent ion channels to
generate automaticity. From a theoretical point of view,
pacemaker activity can be considered as an oscillator
generated by a time-varying outward current with a con-
stant or voltage-dependent inward current activated dur-
ing the repolarization phase. The generation of the dia-
stolic depolarization requires the flow of a net inward
current at the end of the repolarization phase. Several ion
channels with distinct biophysical and pharmacological
properties are potentially involved in the diastolic depo-
larization phase. I
Kr
and I
Ks
are activated during the up-
stroke phase of the action potential and then deactivate
quite slowly during the end of the repolarization and
diastolic depolarization. As I
Kr
and I
Ks
drive outward
currents, a net inward current must overcome outward
components to initiate membrane depolarization. As to
the relevance of membrane ion channels, we will base our
discussion on the working hypothesis that, in isolated
pacemaker cells, such an inward current is generated by
the association of I
f
with I
st
,I
Ca,T
, I
Ca,L
, and I
Na
. The
importance of I
NCX
will be reviewed in a separate section.
A substantial amount of data on ion channels in SAN
pacemaker activity comes from rabbit and guinea pig
SAN. However, new insights have been obtained from
genetically altered mouse strains.
C. Role of I
Kr
and I
Ks
in Automaticity
In rabbit SAN pacemaker cells, the cardiac I
Kr
cur-
rent sets the position of the maximum diastolic potential
and controls action potential repolarization (371, 496).
The physiological significance of I
Kr
in rabbit SAN auto-
maticity has been studied in spontaneously beating cells
using E-4031. In spontaneously active cells, partial inhibi-
tion of I
Kr
by nanomolar concentrations of E-4031 posi-
tively shifts the maximum diastolic potential and de-
creases the action potential amplitude and rate of rise.
E-4031 also prolongs the action potential repolarization
phase (371, 496). The overall physiological effect is a
slowing of the pacing rate in isolated hearts and pace-
maker cells (Fig. 10, A and B). Slowing of pacemaker
activity is due to a reduction in the recruitment of other
currents contributing to the diastolic depolarization, such
as I
f
, I
Ca,L
, and I
Ca,T
. Indeed, the positive shift of the
maximum diastolic potential will reduce I
f
activation and
increase voltage-dependent inactivation of I
Ca,L
and I
Ca,T
during the diastolic depolarization (93). In isolated SAN
cells of rabbit (371) and guinea pig (320), complete block
of I
Kr
by micromolar concentrations of E-4031 quickly
terminates automaticity, and the cell resting potential
settles between 30 and 40 mV. Similar effects of I
Kr
block by E-4031 have been reported in mouse pacemaker
cells by Clark et al. (93). Consistent with this pharmaco-
FIG. 10. A: I
Kr
inhibition by 5
M E-4031 slows the pacing rate of isolated Langendorff-perfused mouse hearts. [From Clark et al. (93).]
B: atrio-sinus electrical interactions influence SAN pacemaking and action potential repolarization phase. Effect of 1
M E-4031 on automaticity
recorded from rabbit SAN tissue strips. Two independent experiments are shown. Asterisks indicate location of the leading pacemaking site. I
Kr
inhibition depolarizes the maximum diastolic potential and slows pacemaking in all tissue strips connected to the crista terminalis. However, E-4031
abolishes pacemaker activity in strip B after that it has been cut off the crista terminalis. Filled circles indicate the site of action potential recording.
[From Verheijck et al. (500).] C: currents from 20 consecutive action potential cycles are averaged before and after application of 2.5
M E-4031.
E-4031-sensitive difference current is obtained by subtraction. Cell has been voltage-clamped with a simulated ideal SAN action potential waveform
(b). c: Current-to-voltage plot of mean E-4031-sensitive current from six different SAN cells voltage-clamped with the action potential waveform in
b. Arrows show direction of current trajectory. The ideal diastolic interval (DD) of simulated action potential is indicated by shadowed grey vertical
box at 65 and 55 mV. [From Clark et al. (93).]
938 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
logical evidence is the observation that mice lacking the
ERG1b-mediated channels have no fast I
Kr
component
and are prone to develop episodes of bradycardia (272). It
is interesting to note that I
Kr
blockers induce negative
chronotropism also in the zebrafish heart (329). This ob-
servation indicates that the function of ERG1/KCNH2
channels in pacemaking may be shared between fishes
and mammals. Clark et al. (93) have studied I
Kr
density
and kinetics during pacemaking by pharmacological sub-
traction of the E-4031-sensitive current from the total
systolic and diastolic current. These recordings have
shown that I
Kr
reaches a maximum during the repolariza-
tion phase at about 25 mV and then progressively de-
clines throughout the diastolic interval to reach almost
the zero level before the following action potential thresh-
old (Fig. 10C). However, Zaza et al. (541) reported that I
Kr
in rabbit SAN cells does not completely deactivate during
the pacemaker cycle and can be present as a sustained
current component throughout the diastolic depolariza-
tion phase.
The role of I
Kr
in rabbit pacemaking has also been
studied by a numerical modeling approach (369, 371).
Calculation results are consistent with the view that I
Kr
is
the key repolarizing current in rabbit SAN cells. I
Kr
in-
ward rectification results in a decrease in the membrane
resistance during the repolarization phase, relative to pla-
teau phase (369). I
Kr
deactivation contributes to the net
current change during the early diastolic depolarization
phase (369). Consequently, I
Kr
activation and kinetics
during the pacemaker cycle are critical factors for deter-
mining activation of ion channels during diastolic depo-
larization. It has also been suggested that I
Kr
dependency
from the extracellular K
concentration ([K
]
o
)iniso-
lated pacemaker cells may account for the sensitivity of
pacing rate to elevated [K
]
o
in the intact SAN (369).
I
Kr
thus plays an obligatory role in action potential
repolarization of isolated rabbit and mouse (93) SAN
cells. However, block of I
Kr
by E-4031 significantly slows,
but does not completely suppress, pacemaking in isolated
Langendorff-perfused mouse hearts (93, 371) (Fig. 10A)or
in intact rabbit right atrial preparations (500) (Fig. 10B).
These observations may be explained by the presence of
the electrotonic load imposed to the SAN by atrial tissue,
which has a more negative diastolic potential. Atrial elec-
trical influence partially compensates for I
Kr
block to
drive SAN repolarization (500). In conclusion, I
Kr
also
plays an important role in the way in which SAN automa-
ticity is coupled to the surrounding atrial tissue. The
regional distribution of I
Kr
expression needs to be taken
into consideration for understanding the physiological
relevance of this current. As previously indicated, I
Kr
is
expressed regionally in the rabbit SAN (249). Small cells
display a lower I
Kr
density than larger cells (278). It has
been proposed that low I
Kr
expression is responsible for
the less negative maximum diastolic potential in central
SAN cells and could explain the enhanced sensitivity of
pacemaking in the SAN center to E-4031 (278).
If I
Kr
appears to be the predominant delayed rectifier
current in the rabbit and rodents, primary SAN automa-
ticity in pigs has been reported to depend from I
Ks
(372).
In this preparation, block of I
Ks
by 293B stops automatic-
ity the same way E-4031 does in rabbit SAN cells. As no
detectable I
Kr
has been found in pig SAN cells, we should
conclude that I
Ks
effectively replaces I
Kr
for pacemaking
in the pig. The presence of I
Ks
rather than I
Kr
in pig SAN
can be a form of adaptation to a slower heart rate in large
mammals compared with rodents (372). Inhibition of I
Ks
has a significant effect on pacemaker activity in guinea pig
(320) and peripheral rabbit (275) SAN cells. As these cells
express both I
Kr
and I
Ks
, I
Kr
can still drive pacemaking
under inhibition of I
Ks
.
D. Role of I
f
in Automaticity
The physiological relevance of the I
f
current in pace-
making has been a matter of debate for many years.
However, a consistent amount of data now shows that I
f
plays a key role in the generation of adult SAN pacemaker
activity. For completeness, we will briefly summarize
some useful aspects of the debate on the role of I
f
in SAN
pacemaking, because this point is still discussed in some
recent reviews (2, 94, 121, 307).
In the SAN, I
f
is the only voltage-dependent current
to be activated upon membrane hyperpolarization. By the
biophysical point of view, f-channels can open during the
late phase of repolarization close to the maximum dia-
stolic potential. According to the current view, I
f
would
initiate the first part of the diastolic depolarization until
the activation threshold of T- and L-type channels is
reached (94, 117). It has been questioned whether I
f
size
and kinetics at diastolic voltages would be compatible
with the amount of inward current needed to initiate a
depolarizing phase in the presence of an outward current
(113, 173). Arguments for this were based on the observed
variability of the I
f
activation threshold in isolated SAN
cells and on the relatively slow I
f
activation kinetics at
positive voltages (114). These kinds of uncertainties have
led Denyer and Brown (114) and Hagiwara et al. (173) to
propose an alternative view of the initiation of the dia-
stolic depolarization in SAN. These authors have mea-
sured a voltage-independent, time-independent “back-
ground” current (I
b
) carried by Na
(173). The size of this
current would be theoretically sufficient to drive the dia-
stolic depolarization upon I
Kr
decay (173). According to
this view, the diastolic depolarization is initiated by the
decay of an outward current (I
Kr
) which unmasks a volt-
age-independent constant depolarizing current (I
b
) (173).
These results have been directly challenged by Di-
Francesco (116) who showed that, at least in some beat-
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 939
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
ing SAN cells, the instantaneous time-independent cur-
rent on hyperpolarization flows in the outward direction
(rather than inward) in the full range of the diastolic
depolarization. This work has also shown that the I
f
size
at weak hyperpolarizing potentials can be substantially
underestimated in patch-clamp experiments due to cur-
rent run-down and wash-out phenomena of the cytosolic
environment (116). However, the reversal potential of I
b
can vary in preparations of isolated SAN cells as reported
by different authors (114, 116, 173, 492, 496). Noble et al.
(359) have used a numerical modeling approach to infer
the effects of varying I
b
and I
f
densities on rabbit SAN
pacemaking. These authors have suggested that I
b
and I
f
may play a reciprocal role in pacemaking, because an
increase in one of these current induces a reduction in the
amount of the other current. As for I
b
, the size of I
f
at
voltages positive to the maximum diastolic potential
would be sufficient to drive the diastolic depolarization.
In this respect, Van Ginneken and Giles (492) and Zaza
et al. (541) have also reported that the size of I
f
can
increase significantly in the diastolic depolarization range.
Particularly, Zaza et al. (541) have shown that the net
inward Cs
-sensitive current during the diastolic depolar-
ization is almost fourfold that theoretically necessary to
drive this phase.
In conclusion, available electrophysiological data
show that I
f
is in fact activated during the diastolic depo-
larization. Both time-independent background currents
and I
f
may show variability between preparations and
recording conditions employed (113, 116, 173, 496). At
least part of SAN background currents can be attributed
to either I
Cl
(37) or TRPM4-mediated I
b
described by
Demion et al. (109) However, their relevance to pacemak-
ing is not yet firmly established. Discussions on the role of I
f
in pacemaking are now focused on whether this current is
necessary for automaticity and on its relative contribution to
the
-adrenergic modulation of heart rate (121, 307).
The relevance of I
f
to pacemaker activity is now
supported by pharmacological and genetic evidence. In-
deed, pharmacological inhibition of f-channels slows
pacemaker activity in vitro in isolated pacemaker cells
and SAN tissue, as well as in vivo in experimental animal
models.
Sensitivity of f-channels to Cs
has been used to infer
the quantitative contribution of f-channels to SAN pace-
maker activity (113, 365). When tested on spontaneously
beating cells, 2–5 mM Cs
significantly slows the pacing
rate (113). Cs
induces negative chronotropism even in
the absence of any evident additional effect on I
Kr
or I
Ca,L
(113). The fact that Cs
slows but does not stop pace-
maker activity has been interpreted as an indication that
I
f
contributes to pacemaking without being an absolute
prerequisite for automaticity (113, 365). This interpreta-
tion has been questioned by DiFrancesco (120), who high-
lighted that Cs
block would be partially relieved at pos-
itive voltages. According to DiFrancesco (120), unblocked
f-channels would still be able to drive the diastolic depo-
larization at rest and to mediate control of pacemaking by
autonomic agonists (420). In conclusion, the action of
Cs
on pacemaker activity demonstrates that I
f
partici-
pates to the generation of pacemaker activity. However,
no definitive quantitative insights about the importance of
f-channels can be obtained by using this ion.
The negative chronotropic action of organic f-chan-
nel blockers such UL-FS 49 (480, 488), ZD-7228 (53, 318),
and ivabradine (480) provides further evidence on the
relevance of I
f
in the control of the diastolic depolariza-
tion rate. These inhibitors of native cardiac I
f
are all open-
and use-dependent f-channel blockers (see Ref. 21 for
review). The mechanism of channel block is essentially
similar for all these drugs. Indeed, the drug binds to the
channel from the intracellular side and probably stays in
its binding site when the channel shuts off during deacti-
vation. When the channel opens during the following
pacemaker cycle, current flow in the inward direction will
promote channel unblock (66, 162, 489, 490).
The most specific organic blocker of I
f
is ivabradine
(21). When tested at concentrations that are selective for
I
f
, ivabradine slows pacemaker activity by specifically
reducing the slope of the linear part of the diastolic
depolarization (68, 479) (Fig. 11A). This drug blocks the
native rabbit SAN I
f
with an EC
50
of 2–3
M (48). At
these concentrations, ivabradine slows automaticity in
isolated rabbit pacemaker cells by 20% (479). Consis-
tently, ivabradine reduces the heart rate in vivo in con-
scious dogs (461) and mice (132, 280) (Fig. 11B). The
observation that ivabradine leads to heart rate slowing,
but does not block pacemaking, suggests that pharmaco-
logical inhibition of f-channels in these conditions can
regulate pacemaking without impairing automaticity per
se. However, the relative fraction of I
f
blocked during
pacemaker activity in these experiments is difficult to
estimate since, as previously discussed, the blocking ac-
tion of ivabradine on f-channels is both use and current
dependent (66). Consequently, current block will be more
pronounced at positive potentials such as during action
potential repolarization and late diastolic depolarization.
The fraction of I
f
blocked for a given concentration of
ivabradine measured at potentials negative to the maxi-
mum diastolic potential is thus likely to represent an
underestimation of the real current block during pace-
maker activity. On the other hand, slowing of pacing rate
induced by ivabradine will favor unblock of the channel.
Reduction of heart rate by ivabradine at a given rate is
thus the equilibrium between f-channel block and un-
block aided by the reduced firing rate. Further investiga-
tion is needed to clarify the exact quantitative relationship
between I
f
block by heart rate reducing agents (as ivabra-
dine) and the slowing of pacemaker activity. The hetero-
geneous distribution of I
f
will also influence the sensitiv-
940
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
ity of pacemaker activity to different f-channel blockers in
the center versus the periphery of the SAN. Indeed, the
fractional slowing of pacemaker activity by Cs
, UL-FS
49, and ZD-7228 is larger in the SAN periphery than in the
center (355).
In addition to electrophysiological and pharmacolog-
ical data, there is now substantial genetic evidence of the
relevance of f-channels in the generation of automaticity.
In animal models, insights are coming from a spontaneous
zebrafish mutant (18) as well as from genetically modified
mice lacking HCN2 and HCN4 channels in the heart (196,
294, 466). The first genetic evidence associating f-chan-
nels with the generation of cardiac pacemaking came
from a large-scale random mutagenesis study aimed to
identify new genes involved in zebrafish cardiovascular
development (Fig. 12A). The mutant line slow mo is as-
sociated with a reduced heart rate in embryos (18) and
adult (513) zebrafish. Isolated heart cells from slow mo
embryos have no fast-activating I
f
component and show
strongly reduced I
f
density (18) (Fig. 12A). Similarly, adult
slow mo fishes have reduced heart rate and low I
f
expres-
sion in the atrium and ventricle (513). The association
between low I
f
expression and bradycardia in slow mo
mutants constitute strong evidence of the importance of
f-channels for automaticity in zebrafish.
Inactivation of HCN2 channels in the mouse heart
induces SAN dysrhythmia and a 30% reduction of I
f
in
isolated SAN cells (294) (Fig. 12B). HCN2 gene inactiva-
tion slows the kinetics of I
f
activation, suggesting that
HCN2 channels may contribute to the fast kinetic compo-
nent of I
f
. Interestingly, maximal I
f
current stimulation by
cAMP is not changed in SAN cells from HCN2 knockout
mice (294) (Fig. 12B). HCN2 knockout mice do not show
reduction of the mean heart rate (294). However, SAN
dysrhythmia of HCN2 knockout mice indicates that these
channels play a role in stabilizing the heart rate. Global or
heart-specific inactivation of HCN4 channels provokes
lethality in mouse embryos. Embryos lacking HCN4 chan-
nels die between day 9 and 12 post coitum (466). Younger
embryos lacking HCN4 channels have slow heart rate and
show almost complete suppression of I
f
. Furthermore,
cAMP regulation of heart rate is abolished in developing
hearts (466). Stieber et al. (466) have also reported an
absence of pacemaker cells having “mature” action poten-
tial characteristics in HCN4-deficient embryonic hearts.
Indeed, only cells characterized by early “embryonic”
pacemaker activity (308) are found in HCN4-deficient
hearts. To overcome the problem of embryonic lethality
in mice lacking HCN4 channels, Hermann et al. (196) have
developed a temporally inducible deletion of Hcn4
(HCN4
C
mouse). SAN cells from HCN4
C
mice show a 75%
reduction of I
f
with a depressed response to isoprotere-
nol. Most HCN4
C
cells are quiescent, but normal automa-
ticity can be restored to normal pacing rates by superfu-
sion of saturating doses of isoproterenol (196). In vivo,
HCN4
C
mice are characterized by the presence of SAN
pauses, yet the maximal heart rate measured under exer-
cise, or after administration of isoproterenol, does not
differ from that of wild-type counterparts (196). Hermann
et al. (196) reported that the frequency of SAN pauses is
significantly increased during transitions for elevated to
low heart rates. These authors have thus proposed that
HCN4 channels are not required for acceleration of heart
rate, but rather constitute a backup mechanism that is
FIG. 11. A: pharmacological inhibition of I
f
by 0.3
M ivabradine
slows pacemaker activity of a rabbit SAN pacemaker cell (top panel).
Note that I
f
inhibition reduces the cell pacing rate by selectively reduc-
ing the slope of the linear part of the diastolic depolarization, without
affecting any other action potential parameter. Ivabradine block of I
f
current is shown in the bottom panel. [From DiFrancesco (121).]
B: in vivo heart rate reduction in mice treated by an intraperitoneal bolus
of increasing doses of ivabradine. (AP) indicate coinjection with iv-
abradine of atropine and propranolol to inhibit pharmacologically, the
autonomic input (see sect. VIIA). [Original data from Marger et al. (316).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
941
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
important for stimulating and stabilizing pacemaker initi-
ation in conditions of lowered heart rate.
In summary, results obtained in mice lacking HCN2
and HCN4 channels suggest a predominant role of HCN4
channels in mediating the adrenergic stimulation of I
f
.
However, results obtained from HCN4
C
mice (196) indi-
cate that HCN4 channels are not the only mechanism
involved in the
-adrenergic regulation of heart rate, at
least in the adult mouse.
E. Role of I
Ca,L
in Automaticity
L-type Ca
2
channels are important contributors to
the upstroke phase of the pacemaker action potential and
play a role in the generation of the SAN diastolic depo-
larization (for a recent review, see Ref. 310). Block of I
Ca,L
by nifedipine stops pacemaker activity in primary central
pacemaker cells (250). However, application of nifedipine
to SAN tissue strips from the periphery of the node slows
pacemaking but does not block automaticity. These ob-
servations have been accounted for by the presence in the
SAN periphery of I
Na
driving the action potential upstroke
under I
Ca,L
blockade (250). The differential effect of I
Ca,L
inhibition in the center and in the periphery of the rabbit
SAN underlines the problem of separating the possible
contribution of I
Ca,L
to the diastolic depolarization from
that of the upstroke phase of the action potential. This
problem is particularly difficult to solve, since pharmaco-
logical block of I
Ca,L
during the action potential upstroke
can have “knock on” effects on the recruitment of ionic
currents during repolarization as well as the following
diastolic depolarization. Furthermore, I
Ca,L
is regulated by
PKA and CaMKII (390, 509). Consequently, the contribu-
tion of I
Ca,L
to automaticity can be influenced by the
phosphorylation state of the channel. CaMKII activity is
required for pacemaking, since CaMKII inhibitors termi-
nate automaticity in isolated cells (509). The necessity of
CaMKII in automaticity can be explained, in part, by the
regulation of I
Ca,L
activation and reactivation kinetics by
CaMKII (509). However, CaMKII can also have other un-
known targets in pacemaker cells. Thus sensitivity of
pacemaking to inhibition of CaMKII should not be taken
as evidence of the importance of I
Ca,L
in the genesis of
automaticity per se.
There is now substantial genetic and pharmacolog-
ical data supporting the view that I
Ca,L
contributes to
both the action potential upstroke and diastolic depo-
larization. Indeed, Ca
v
1.2 channels seem to be involved
in the action potential upstroke phase, while Ca
v
1.3
channels substantially contribute to the diastolic depo-
larization (309, 310).
Verheijck et al. (497) have been the first to propose
that the I
Ca,L
activation threshold is more negative in SAN
than in the ventricle (497). These authors recorded the
FIG. 12. A: Slow mo zebrafish mutant embryos have reduced heart
rate and lack the fast kinetic component of I
f
. Heart rate slowing is
consistently found throughout the embryo development and is indepen-
dent from the water temperature (top panel). I
f
current density is
strongly downregulated in cardiac myocytes isolated from slow mo
embryos. Residual I
f
in Slow mo hearts displays slower activation kinet-
ics than in wild-type (WT) counterparts (bottom panel). [From Baker
et al. (18), copyright National Academy of Science, USA.] B: telemetric
electrocardiograms in freely-moving HCN2 knockout (HCN2
/
) mice
show SAN arrhythmia. Note that the basal heart rate is similar in WT and
HCN2
/
mice, but the interbeat interval is more variable in knockout
animals. The density of I
f
is reduced by 30% in HCN2
/
SAN cells, but
current responsiveness to cAMP is unaltered. [From Ludwig et al. (294),
with permission from Macmillan Publisher Ltd.]
942 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
DHP-sensitive current during the diastolic depolarization
and showed that I
Ca,L
is activated as early as in the linear
part of the pacemaker potential. I
Ca,L
current density then
increases at the action potential threshold and becomes
fully activated during the upstroke. This work demon-
strated that the kinetics of SAN I
Ca,L
do not behave as its
ventricular counterpart but is activated at more negative
voltages to support the pacemaker potential.
Involvement of L-type channels in pacemaking in vivo
is also strongly suggested by in vivo ECG recordings
showing that DHPs induce bradycardia in anesthetized
mice (268). The unexpected observation that mice lacking
L-type Ca
v
1.3 channels had pronounced bradycardia and
SAN arrhythmia was the first genetic indication of the
importance of these channels in pacemaker activity (393).
In these mice, bradycardia persists after pharmacological
block of the autonomic nervous system, and intact atria
from Ca
v
1.3
/
mice have slower pacing rate than wild-
type counterparts (393) (Fig. 13). Two studies have dem-
onstrated that Ca
v
1.3 channels play a major role in auto-
maticity of isolated SAN (546) (Fig. 14A) and in pace-
maker cells (309) (Fig. 14B). Indeed, pacemaker cells
from Ca
v
1.3 knock-out mice have erratic and intermittent
pacemaking, leading to an overall lower degree of au-
tomaticity than that of wild-type cells (309) (Fig. 14B).
These studies have also shown that Ca
v
1.3 channels
generate I
Ca,L
with different properties than that of
Ca
v
1.2-mediated I
Ca,L
. Indeed, native Ca
v
1.3 channels
activate at negative potentials from about 50 mV (see
Fig. 15A). Inactivation of Ca
v
1.3 channels shifted the
activation of I
Ca,L
to more positive potentials, thereby
abolishing I
Ca,L
in the voltage range corresponding to
FIG. 13. Electrocardiograms of freely moving Ca
v
1.3
/
mice show
bradycardia and slowing of atrioventricular conduction. Note that in
Ca
v
1.3
/
mice, the interbeat interval is irregular and significantly longer
than in WT mice (SAN arrhythmia, top panel). Association of bradycar-
dia and arrhythmia is typical of resting Ca
v
1.3
/
mice with unaltered
autonomic regulation of heart rate (ANS). When the input of the
autonomic nervous system is blocked postsynaptically by combined
administration of atropine and propranolol (ANS), arrhythmia is in
part compensated, but bradycardia and prolongation of the PR interval
are still evident. (Mangoni and Nargeot, unpublished recordings.)
FIG. 14. A: intracellular microelectrode recordings of automaticity in isolated SANs from WT, Ca
v
1.3
/
, and Ca
v
1.3
/
mice. Note slowing of
automaticity and pauses in a SAN from Ca
v
1.3
/
mouse. [Data from Zhang et al. (546).] B: isolated SAN pacemaker cells from Ca
v
1.3
/
mice have
slower pacemaker activity than WT cells (top panel). Slow pacemaking in cells lacking Ca
v
1.3 channels is associated with irregular interbeat interval
(bottom panel). [Data from Mangoni et al. (309).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
943
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
that of the diastolic depolarization. Inactivation of
Ca
v
1.3 channels reduces I
Ca,L
density by 70% in pace-
maker cells. Residual I
Ca,L
is attributable to Ca
v
1.2-
mediated I
Ca,L
. Consistent with this hypothesis, I
Ca,L
in
pacemaker cells from Ca
v
1.3 knockout mice is more
sensitive to DHPs (309) and has faster inactivation
kinetics (546). Activation of
-adrenergic receptors by
norepinephrine shifts negatively the threshold for acti-
vation of Ca
v
1.3-mediated I
Ca,L,
to about 55 mV (309).
Interestingly, this threshold is comparable to that ob-
served for the nifedipine-sensitive I
Ca,L
measured in
spontaneously beating rabbit SAN cells (497). This sug-
gests the existence of Ca
v
1.3 channels also in rabbit
SAN cells. The maximal pacing rate in Ca
v
1.3 knockout
hearts in the presence of isoproterenol is slightly
slower than that of wild-type hearts (322). This obser-
vation can be explained by taking into consideration
that stimulation of I
Ca,L
by saturating concentrations of
norepinephrine cannot compensate for the lack of I
Ca,L
in the diastolic depolarization range (309).
In conclusion, insights gained from mice lacking
Ca
v
1.3 channels are indicative of a distinction in the func-
FIG. 15. A: voltage-dependent Ca
2
cur-
rents in pacemaker SAN cells from WT and
Ca
v
1.3
/
mice. Currents are evoked from a
holding potential of 80 mV at the indicated
test potentials. Current stimulation by BAK K
8644 is used to detect activation of I
Ca,L
at a
given test voltage. In Ca
v
1.3
/
SAN cells, BAY
K 8644 has no effect at test potentials negative
to 30 mV. This observation indicates that
inactivation of Ca
v
1.3 channels abolishes a
I
Ca,L
component which has intermediate acti-
vation threshold between I
Ca,T
, and the “clas-
sical” I
Ca,L
.Ca
v
1.3-mediated I
Ca,L
is activated in
the diastolic depolarization range (DD, shaded
gray box). [Data from Mangoni et al. (309).]
B: inactivation of Ca
v
3.1 channels suppresses
I
Ca,T
in spontaneously active SAN and AVN
cells isolated from Ca
v
3.1
/
mice. Note reduc-
tion of the total Ca
2
current density and the
slowing of the total I
Ca
inactivation. The resid-
ual Ca
2
current is attributed to Ca
v
1.3- and
Ca
v
1.2-mediated I
Ca,L
. [Data from Mangoni
et al. (314).] C: current-to-voltage relationships
(left) and steady-state inactivation (right)of
native SAN Ca
v
3.1, Ca
v
1.3, and Ca
v
1.2 chan-
nels. [Adapted from Mangoni et al. (310).]
944 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
tional role of Ca
v
1.3 channels contributing to automaticity
and Ca
v
1.2 channels triggering myocardial contraction.
Indeed, suppression of Ca
v
1.3-mediated I
Ca,L
impacts on
cardiac automaticity and conduction, but has no effect on
myocardial contractile performance (322). The differen-
tial roles of Ca
v
1.3 and Ca
v
1.2 channels in the heartbeat
has also been shown pharmacologically by employing a
knock-in mouse strain in which Ca
v
1.2 channels are in-
sensitive to DHPs (Ca
v
1.2DHP
/
mouse) (462). The in
vivo bradycardic effect induced by DHPs is not changed
in Ca
v
1.2DHP
/
mice, indicating that the dominant L-
type VDCC isoform participating to automaticity is in fact
Ca
v
1.3 (462).
As stated above, Ca
v
1.3 knockout mice also display a
slowing of atrioventricular conduction (see Fig. 13) (322,
393, 546). Particularly, Ca
v
1.3 knockout mice show I and
II degree atrioventricular blocks, an observation which
has been found in freely moving (393) and anesthetized
mice (546), as well as in isolated Ca
v
1.3 knockout hearts
(322). It will be interesting to test the hypothesis that
Ca
v
1.3 channels are also involved in automaticity of the
AVN.
Inactivation of the Ca
v
1.2
1-subunit induces lethal-
ity of the late mouse embryo. Heart development in early
Ca
v
1.2 knockout embryos is paralleled by overexpression
of the Ca
v
1.3
1-subunit (529), as well as by upregulation
of a distinct L-type Ca
2
current that is also present in
Ca
v
1.3 knockout embryonic hearts (445). Overexpres-
sion of Ca
v
1.3 channels can be a compensatory mecha-
nism for the loss of Ca
v
1.2 channels, even if Ca
v
1.3 chan-
nels cannot ensure viability of the late embryo. Embry-
onic hearts from a spontaneous zebrafish mutant strain
lacking the Ca
v
1.2
1 subunit (Isl mutant) show de-
pressed and erratic atrial beating (425). Consistent with
the role of Ca
v
1.2 in mediating cardiac contraction, hearts
from Isl mutants have no ventricular contraction. It is not
known whether Ca
v
1.2 channels can contribute to auto-
maticity in embryonic fish atria. Jones et al. (222) re-
ported progressive loss of Ca
v
1.2 protein in SAN tissue
from aging guinea pigs, concomitantly with a reduction in
automaticity and an augmentation of SAN rate sensitivity
to DHPs. These data are suggestive of a participation of
Ca
v
1.2 channels in pacemaking, at least during the aging
process.
F. Role of I
Ca,T
in Automaticity
I
Ca,T
is expressed in the SAN (145, 172), AVN (140),
and Purkinje fibers (199, 484). T-channels can be found in
the automatic mouse and zebrafish embryonic myocar-
dium (18, 104). I
Ca,T
has also been described in the am-
phibian sinus venosus (49).
The role of I
Ca,T
in cardiac automaticity has long
been uncertain. Native I
Ca,T
has low steady-state availabil-
ity at diastolic membrane voltages typical of leading SAN
pacemaker cells. However, it has long been known that
I
Ca,T
inhibitors such as Ni
2
and tetrametrine slow pace-
maker activity of SAN cells (172, 439).
The role of T-channels in automaticity has been re-
cently investigated by targeted inactivation of the Ca
v
3.2
(84) and Ca
v
3.1 (314)
-subunit isoforms. Mice lacking
Ca
v
3.2 channels have no ECG alterations (84), indicating
that the lack of this isoforms either has no impact on the
generation and conduction of the cardiac impulse, or that
a compensatory mechanism is established during heart
development. It is not still possible to discriminate be-
tween these hypotheses, because SAN ionic currents and
automaticity have not yet been studied in Ca
v
3.2
/
mice.
A striking finding coming from Ca
v
3.1
/
mice is that I
Ca,T
cannot be detected in SAN and AVN cells (Fig. 15B).
Indeed, as both Ca
v
3.1 and Ca
v
3.2 mRNA are expressed in
mouse SAN, one would expect to find residual I
Ca,T
in
Ca
v
3.1
/
pacemaker cells. The absence of Ca
v
3.2-medi-
ated I
Ca,T
in Ca
v
3.1
/
SAN cells has not yet been ex-
plained, but recent studies have reported that only Ca
v
3.1-
mediated I
Ca,T
is recorded in rat atria after birth (146,
358). These findings are suggestive of a developmental
switch between Ca
v
3.2 and Ca
v
3.1 channels in the heart.
According to this hypothesis, Ca
v
3.2 channels are func-
tionally expressed in the developing myocardium, while
Ca
v
3.1 channels are predominant in the adult heart. It
would be important to investigate the differential role of
Ca
v
3.1 and Ca
v
3.2 channels to automaticity during the
cardiac development. Inactivation of Ca
v
3.1 channels in-
duces moderate bradycardia and slowing of atrioventric-
ular conduction in Ca
v
3.1
/
mice (314) (Fig. 16). Moder-
ate bradycardia is observed in Ca
v
3.1
/
mice even after
pharmacological block of the autonomic nervous system,
indicating slowing of SAN automaticity (Fig. 16C). Ac-
cordingly, spontaneous activity in isolated SAN pace-
maker cells is slowed by 30% (Fig. 16D).
Despite pharmacological and genetic data showing
the involvement of I
Ca,T
in pacemaking, we presently lack
a precise description of how T-channel activity contrib-
utes to the diastolic depolarization. The existence of a
I
Ca,T
-mediated “window” current component in the dia-
stolic depolarization range has been proposed for rabbit
SAN cells (401). However, this is not a consistent finding
between different authors (see, for example, Ref. 172).
The relatively negative threshold for activation of Ca
v
1.3-
mediated I
Ca,L
can possibly interfere with measurement of
I
Ca,T
, leading to the false impression of residually avail-
able I
Ca,T
at about 50 mV (M. Mangoni, unpublished
observations; see also Fig. 15, A and C). As the absolute
density of I
Ca,T
available at pacemaker potentials is low,
one may wonder whether contribution of T-channels to
pacemaking can be due to coupling with intracellular
Ca
2
signaling. Such a mechanism has been proposed by
Huser et al. (211) in cat latent atrial pacemaker cells.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 945
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
These authors have recorded intracellular Ca
2
release
during pacemaking and found that 40
MNi
2
reduced
Ca
2
release from the sarcoplasmic reticulum (SR) and
the slope of the late phase of the diastolic depolarization.
Ni
2
did not affect Ca
2
release during the action poten-
tial upstroke (systolic Ca
2
), but only Ca
2
released dur-
ing the diastolic depolarization phase (diastolic Ca
2
). In
latent pacemaker cells, I
Ca,T
activation during diastolic
depolarization triggers local Ca
2
release, generating an
inward current through stimulation of I
NCX
(211). Accord-
ing to Lipsius et al. (288), there exists a local control of
Ca
2
release involving T-type channels, RyRs of the SR,
and NCX (288). Such a functional coupling between T-
type channels and SR could explain earlier observations
indicating that prevention of SR Ca
2
release with ryan-
odine reduces I
Ca,T
(285). No direct evidence for I
Ca,T
-
induced SR Ca
2
release in primary SAN pacemaker cells
has yet been reported. Vinogradova et al. (506) have re-
ported that 50
MNi
2
failed to inhibit subsarcolemmal
diastolic Ca
2
release in spontaneously beating rabbit
SAN cells, as well as spontaneous Ca
2
release in arrested
pacemaker cells (510). However, these data do not ex-
clude the possibility that Ca
v
3.1-mediated T-channels can
contribute to SR Ca
2
release since native SAN Ca
v
3.1
channels are resistant to this concentration of Ni
2
(314).
The more negative diastolic potential in latent pacemaker
cells can favor T-channel opening during the diastolic
depolarization. The role of I
Ca,T
in automaticity of the
conduction system is still unknown. Ca
v
3.1
/
mice have
slowed atrioventricular conduction (314) (Fig. 16). Mea-
surements of atrioventricular conduction times indicate
that the AVN is the site where impulse propagation is
slowed (314). It will be interesting to test if automaticity
of AVN cells is affected by inactivation of Ca
v
3.1 channels.
Ca
v
3.1 channels do not appear to be involved in conduc-
tion in mouse Purkinje fibers. Indeed, neither the intra-
ventricular conduction time nor the QRS are changed in
Ca
v
3.1
/
mice (314).
G. Role of N- and R-type Channels in
Heartbeat Regulation
Beside L- and T-type channels, an involvement of
VDCCs belonging to the Ca
v
2 gene family in the regulation
of heart rhythm and rate has been recently proposed.
Mouse lines lacking N-type Ca
v
2.2 and R-type Ca
v
2.3 chan-
nels have been developed (214, 293, 516). No expression
of Ca
v
2.2 channels in the heart has been reported. How-
ever, these channels contribute to the activity of sympa-
thetic nerve terminals (214). Ca
v
2.2
/
mice have in-
creased heart rate and mean blood pressure (214). These
effects are possibly due to reduced activity of the sympa-
thetic nerve terminals (214). Parasympathetic input is not
affected in these mice (214).
Expression of Ca
v
2.3 isoforms in the mouse heart has
been reported in atrial tissue by in situ hybridization
(332), and in the heart at the protein level (293). Enhanced
sympathetic tonus, ventricular arrhythmias, and dysfunc-
tion in ventricular conduction have been found in mice
lacking Ca
v
2.3 channels (293, 516). Disturbances of atrial
activation and alterations of the QRS complex are ob-
FIG. 16. Ca
v
3.1
/
mice have moderate heart rate reduction and prolongation of the PR interval. A: telemetric electrocardiograms of WT and
Ca
v
3.1
/
mice. B: circadian variability of heart rate in WT and Ca
v
3.1
/
mice. Continuous recordings lasting 24 h provide evidence that the mean
heart rate is lowered in Ca
v
3.1
/
mice (dotted lines). C:Ca
v
3.1
/
mice have intrinsically slower heart rate than WT mice. D: pacemaker activity
of isolated SAN cells is slower in Ca
v
3.1 than in WT mice. [Data from Mangoni et al. (314).]
946 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
served even after administration of atropine and propran-
olol (516). Overexpression of Ca
v
3.1 mRNA has been
reported in adult Ca
v
2.3
/
hearts (516), but direct evi-
dence supporting functional expression of Ca
v
2.3 chan-
nels in the mouse SAN and conduction system is lacking.
H. Role of I
Na
in Automaticity
Primary pacemaker cells are characterized by a rel-
atively slow action potential upstroke. The concept that
the action potential of pacemaker cells is mainly driven by
Ca
2
rather than Na
channels is still valid, although new
studies on the rabbit SAN have indicated that a distinction
has to be made between small pacemaker cells from the
SAN center and larger cells from the periphery of the
node. Two different I
Na
components have been function-
ally identified in SAN, a TTX-sensitive “neuronal” I
Na
pre-
dominantly coded by the Na
v
1.1 (24, 25, 279, 304) isoform,
and a TTX-resistant I
Na
, coded by the Na
v
1.5 isoform (276,
279). TTX-sensitive I
Na
contributes to pacemaking in new-
born rabbits (24) (Fig. 17A) and in the adult mouse (279,
304). Na
v
1.5-mediated TTX-resistant I
Na
is involved in
SAN intranodal conduction (276) (Fig. 17, B and C).
The contribution of TTX-sensitive I
Na
to automaticity
has been studied in neonatal rabbit SAN by Baruscotti
et al. (24). In newborn SAN cells, application of 3
M TTX
slows pacemaker activity by 63%. Baruscotti et al. (24)
have reported slowing by TTX of the late phase of the
diastolic depolarization, reduction of the action potential
threshold, and overshoot. These observations indicate
that TTX-sensitive I
Na
contributes to SAN pacemaking
essentially by quickening the late diastolic depolarization
phase and by shifting the action potential threshold neg-
atively to that of VDCCs. The mechanism of contribution
of TTX-sensitive I
Na
to the diastolic depolarization in
neonatal rabbit SAN cells has also been investigated by
these authors in two distinct studies (23, 24). In the first
study (24), they reported the presence of a significant I
Na
“window” component that declines with increasing age of
the animal. In a follow-up (23) study, these authors have
shown that the amount of TTX-sensitive current recorded
by applying ramp depolarizations mainly depends on the
ramp slope. This observation suggests that incomplete
inactivation of TTX-sensitive I
Na
during the upstroke
phase underlies persistent I
Na
during the diastolic depo-
larization (23). The expression of the Na
v
1.1 isoform is
downregulated in the adult (Fig. 17A) (23). The physio-
logical significance of the expression of Na
v
1.1-mediated
I
Na
in newborn SAN has not been completely elucidated,
but it is likely that TTX-sensitive I
Na
constitutes a mech-
FIG. 17. A: differential effect of TTX on pacemaker activity in SAN pacemaker cells from newborn (top, left) and adult (top, right) rabbits.
Inhibition of I
Na
by TTX blocks pacemaking in newborn cells, while it has no effect on automaticity of adult cells. A voltage-clamp ramp protocol
shows the presence of a TTX-sensitive current component in pacemaker cells from newborn rabbits (bottom, left). Such a component is not recorded
in cells from adult rabbits, indicating a developmental regulation of the expression of TTX-sensitive I
Na
. [From Baruscotti et al. (23).] B:a
TTX-sensitive I
Na
contributes to the exponential phase of the diastolic depolarization in the mouse SAN. The figure depicts an “action potential
clamp experiment” in which a mouse SAN cell has been clamped with its own action potential (top, left) so that the net clamp current equals zero.
The cell has then been exposed to TTX and the current necessary to maintain the membrane voltage recorded. This net TTX-sensitive I
Na
is plotted
versus the time (bottom, left) and versus the membrane voltage (right). From this experiment, it is apparent that TTX-sensitive I
Na
activates in the
exponential phase of the diastolic depolarization between 45 and 40 mV, quickly reaches its peak at 20 mV, and switches off at about 20 mV.
[From Lei et al. (279).] C: electrical mapping of SAN intranodal conduction in WT and Scn5a
/
mice. The change in the SAN activation sequence
in Scn5a
/
mice indicates that TTX-resistant Scn5a-mediated I
Na
is important in impulse conduction within the SAN and from the SAN to the right
atrium. [From Lei et al. (276), with permission from Wiley-Blackwell.]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
947
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
anism for increasing the basal heart rate in newborn
animals (24).
During the last years, it has become apparent that I
Na
significantly contributes to mouse pacemaking. Pharma-
cological inhibition of I
Na
by lidocaine reduces the heart
rate of adult mice (268). The pacing rate of Laghendorff-
perfused mouse hearts is slowed by low doses of TTX
(50–100 nM) (304). Adult mouse pacemaker cells show
constitutive expression of I
Na
(86, 312). Both TTX-sensi-
tive and -resistant I
Na
have been functionally identified in
mouse SAN cells (279). Similar to newborn rabbits, ex-
pression of Na
v
1.1
-subunit has been reported in SAN
sections from adult mice (279, 304). These in vivo and in
vitro experiments have demonstrated the involvement of
TTX-sensitive, neuronal I
Na
in mouse cardiac pacemak-
ing. The action of TTX on cellular pacemaking has been
directly investigated in spontaneously beating mouse SAN
cells by Lei et al. (279). These authors have combined
voltage-clamp recordings of I
Na
with staining of SAN tis-
sue with antibodies directed against Na
v
subunits. TTX-
sensitive and TTX-resistant I
Na
have been recorded. TTX-
sensitive I
Na
is associated with expression of the Na
v
1.1
-subunit, while TTX-resistant I
Na
is associated with
Na
v
1.5 expression. Block of TTX-sensitive I
Na
by 50 nM
TTX slows automaticity of intact SAN and isolated pace-
maker cells (279). TTX-sensitive I
Na
has been recorded in
action potential clamp experiments and shown to be
present during the late phase of the diastolic depolariza-
tion and upstroke (Fig. 17B) (279). Na
v
1.5 channels un-
derlie TTX-resistant I
Na
(276, 279). The role of Na
v
1.5
channels in SAN pacemaking has been described in het-
erozygous Scn5a
/
mice (276). Scn5a
/
mice display
major age-dependent dysfunction in atrioventricular con-
duction and a moderate reduction of the mean heart rate
(381). Intact atrial-SAN preparations from Scn5a
/
mice
have shown normal pacemaking in the SAN center, but
demonstrated slower intranodal SAN conduction and exit
block (276) (Fig. 17C). Pacemaking in small leading SAN
cells from Scn5a
/
mice is not different from that re-
corded in wild-type mice. In conclusion, available evi-
dence indicates that Na
v
1.5-mediated I
Na
does not partic-
ipate in the generation of automaticity per se (in the
central SAN), but can influence heart rate by contributing
to impulse propagation within the SAN and from the SAN
to the atrium. (276).
I. Role of I
st
in Automaticity
Limited knowledge exists on the role of I
st
in the
generation of cardiac automaticity. Indeed, we presently
lack selective pharmacological agents targeting the I
st
current. Genetic manipulation of st-channels is also im-
possible, since their molecular basis is still unknown.
Consequently, it is difficult to directly investigate the
significance of I
st
in pacemaking. A significant contribu-
tion of I
st
in controlling the rate of the diastolic depolar-
ization has been proposed by Shinagawa et al. (457) on
the basis of numerical simulations of pacemaking in the
rat SAN. Modeling suggests the possibility that I
st
contrib-
utes to pacemaking by virtue of its low threshold of
activation and slow inactivation rate. These properties
would allow I
st
to be present throughout the pacemaker
cycle (457). The potential contribution of I
st
compared
with that of other currents involved in pacemaking has
also been investigated by Zhang et al. (544). This study
suggests that I
st
will affect pacemaking depending on its
relative size compared with other currents contributing to
pacemaking such as I
f
or I
Na
. Consequently, in rabbit
central SAN pacemaker cells, a significant contribution of
I
st
to pacemaking is predicted. According to Zhang et al.
(544), pacemaking at the periphery of the SAN would be
less sensitive to I
st
, because of the larger density of I
Na
and I
f
(544). Elucidation of the molecular nature of I
st
is
needed to gain new insights into the physiological role of
this current.
L. SR Ca
2
Release and Automaticity
Two studies by Rubenstein and Lipsius (428) and Li
et al. (285) supplied initial evidence that Ca
2
release and
I
NCX
were involved in the generation of pacemaker activ-
ity. Rubenstein and Lipsius (428) observed that ryanodine
slowed automaticity in atrial subsidiary pacemaker cells
by reducing the slope of the late diastolic depolarization
(the exponential phase), while Cs
(presumably by block-
ing I
f
) reduced the first fraction of the diastolic depolar-
ization. They concluded that multiple mechanisms were
involved in automaticity of subsidiary cells and that Ca
2
release from the SR was important in the generation of
the late phase of the diastolic depolarization (428). Slow-
ing of automaticity by ryanodine was also reported by
Rigg and Terrar in guinea pig SAN preparations (417) and
by Li et al. (285) in cultured rabbit SAN cells. The latter
report also described a reduction of intracellular Ca
2
transients (417) and abolition of I
NCX
by ryanodine and
BAPTA-AM (285). These studies were indicative of a role
of intracellular Ca
2
signaling in maintaining automatic-
ity, yet they did not establish a link between a specific
Ca
2
signal and pacemaking. Particularly, we can wonder
if spontaneously active cells posses an intracellular Ca
2
signal coupled to the diastolic depolarization phase. Ju
and Allen (226) have described distinct types of Ca
2
signals in toad sinus venosus cells: a cytosolic transient
(predominantly driven by the action potential), a delayed
Ca
2
signal linked to a change in nuclear Ca
2
content,
and a third signal possibly due to subsarcolemmal local
Ca
2
-induced Ca
2
release (LCICR).
948
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
During the last five years, the group led by E. Lakatta
has extensively investigated the cellular mechanism gen-
erating LCICR and emphasized its relevance in determin-
ing the chronotropic state of rabbit SAN cells. Bogdanov
et al. (44) have studied intracellular Ca
2
release in beat-
ing SAN cells and documented “preaction potential” Ca
2
signals due to LCICR. These signals precede the Ca
2
transient evoked by the action potential upstroke phase
(Fig. 18, A–C). These signals are initiated at the cell edge
and are due to Ca
2
release from the SR, since they are
abolished by ryanodine (Fig. 18D). During spontaneous
activity, preaction potential signals are generated by
LCICR. LCICR stimulates I
NCX
activity (44). Using ramp
protocols mimicking the diastolic depolarization, Bog-
danov et al. (44) have estimated LCICR-mediated I
NCX
density to be 1.7 pA/pF (Fig. 18F). The size of I
NCX
would thus be sufficient to quicken the exponential frac-
tion of the diastolic depolarization. LCICRs observed in
rabbit SAN cells are similar to diastolic Ca
2
release
described in cat latent pacemaker cells (211). Vino-
gradova et al. (510) have supplied evidence indicating that
LCICRs in rabbit primary SAN can be generated in the
absence of a change in the membrane voltage. Indeed,
LCICRs can still be observed in quiescent permeabilized
SAN cells. LCICR size depends on extracellular Ca
2
concentration. (Fig. 19, A and B). Furthermore, if sponta-
neously beating SAN cells are arrested by voltage clamp-
ing at the maximum diastolic potential or at a positive
voltage (10 mV), LCICRs do not disappear, but persist
for some seconds in the absence of action potentials (Fig.
19C) (510). Under these conditions, LCICRs display sto-
chastic “roughly periodical” (510) behavior and maintain
similar periodicity as during spontaneous activity. These
results support the view that LCICRs are not directly
linked to the activity of membrane ion channels but re-
flect the existence of an independent intracellular “Ca
2
clock” mediated by spontaneous Ca
2
release from SR.
LCICRs eventually disappear in arrested cells possibly
after SR depletion. We can expect that during pacemak-
ing, Ca
2
stores of the SR are cyclically refilled by open-
ing of VDCCs (510) or by store-operated Ca
2
influx as
described by Ju et al. (229).
Cellular mechanisms underlying this internal “Ca
2
clock” are not understood, but Vinogradova et al. (508) have
recently shown that LCICRs are abolished by inhibition of
PKA activity and stimulated by cAMP (Figs. 20 and 21). They
also reported that basal PKA activity is almost 10 times
higher in SAN cells than in atrial and ventricular myocytes
and that high PKA activity seems to be a prerequisite for
pacemaking. Indeed, PKA inhibitors can stop cellular auto-
maticity (Fig. 20) (508).
In conclusion, SR-mediated LCICR is involved in the
generation of the exponential fraction of the diastolic
depolarization. The two key elements of this mechanism
are pre-action potential RyR-mediated LCICR and NCX,
FIG. 18. Characteristics of LCICR in rabbit SAN pacemaker cells. A: line scan image of Ca
2
release (bottom panel) and corresponding
normalized fluorescence (top panel) as a function of time and position within the scan line. The scan line lays perpendicularly to the cell long axis
(inset cell drawing). Colored arrows indicate areas where fluorescence is averaged. B: pacemaker activity (black line) and normalized fluorescence
at the cell edge (red line) and middle (green line). Arrows indicate the time interval corresponding to the 3-dimensional plot in A. Asterisks indicate
a peak of Ca
2
release at the cell edge. This Ca
2
release occurs locally, during the exponential part of the diastolic depolarization. C: here, the line
scan image is oriented parallel to long cell axis and visualizes the cell edge where LCICR is triggered. White lines indicate the propagation of
the [Ca
2
]
i
wave. D:3
M ryanodine (ryan) slows pacemaker activity and abolishes preaction potential LCICR in rabbit SAN cells. E: dose-response
curve of pacemaker activity inhibition by ryanodine. F:Ca
2
release (top panel) and total membrane current (middle panel) during voltage clamping
of a SAN cell with a waveform simulating the SAN pacemaker cycle (bottom panel). In each panel, the black line indicates recording in control
conditions, and the red line is plotted after application of 3
M ryanodine. Ryanodine blocks diastolic LCICR and reduces the inward current
recorded during the simulated diastolic depolarization (see enlarged middle inset). Asterisks indicate LCICR; the symbol (#) indicates the residual
inward current in presence of ryanodine. [From Bogdanov et al. (44).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
949
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
which converts subsarcolemmal Ca
2
release to an inward
current contributing to diastolic depolarization. In a recent
work, Lyashkov et al. (299) have reported colocalization of
NCX and RyRs in rabbit SAN cells and proposed that the
proximity between NCX and RyRs permits quick conversion
of LCICR into oscillations of the membrane voltage.
FIG. 19. LCICR in intact (A) and permeabilized (B) rabbit SAN pacemaker cells. A: simultaneous recordings of pacemaker activity and confocal
line scan images of LCICR as in Fig. 18C. B: images of Ca
2
release in a quiescent permebilized SAN cell under different extracellular Ca
2
concentrations as indicated. LCICR is present even in quiescent SAN cells, and its amplitude is a function of extracellular Ca
2
concentration. C:
voltage clamping to the cell maximum diastolic potential does not immediately block LCICR in SAN cells. Note LCICR events (wither arrows) in
the absence of changes in membrane voltage. Membrane voltage oscillations, possibly due to I
NCX
activated by LCICR, are recorded. These
observations indicate LCICR are a spontaneous voltage-independent phenomenon. [From Vinogradova et al. (510).]
FIG. 20. Basal PKA-dependent phosphorylation is necessary for maintaining pacemaker activity in rabbit SAN pacemaker cells. A: superfusion
with 15
M of protein kinase inhibitor (PKI) reversibly suppresses pacemaking. B: a line scan image shows strong reduction of spontaneous LCICR
by PKI in permebilized SAN cells. C and D: simultaneous recordings of pacemaker activity (top panel), line scan images (middle panel), and
normalized Ca
2
release averaged over the line scan image (bottom panel). In C, control conditions are established by voltage clamping a
spontaneously beating SAN cell to 30 mV to record spontaneous Ca
2
release. In D, note the reduction in both frequency and size of spontaneous
Ca
2
release by PKI. [From Vinogradova et al. (510).]
950 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
VII. AUTONOMIC REGULATION OF
PACEMAKER ACTIVITY
A. Principles
In this section, we review the current knowledge on
the cellular mechanisms underlying the autonomic regu-
lation of cardiac automaticity.
The autonomic nervous system is the major extracar-
diac determinant of the heart rate. Sympathovagal control
of cardiac automaticity is a complex phenomenon. Multi-
ple interactions occur between sympathetic and parasym-
pathetic centers in the central nervous system, and pre-
synaptic peripheral interactions exist also (223, 284, 375).
In the adult heart, the sympathetic branch of the auto-
nomic nervous system accelerates heart rate, while the
parasympathetic branch slows it. Exceptions to this gen-
eral rule can be observed experimentally, but a detailed
discussion of these mechanisms is beyond the scope of
this review. The terminal part of the cardiac autonomic
nervous system is constituted by the intrinsic cardiac
neuronal plexus (ICNP), which plays a pivotal role in
regulating the heart rate, conduction, and contractile
force (15). Autonomic fiber projections to the heart rhyth-
mogenic centers are abundant. The canine SAN is densely
innervated by postganglionic fibers of the ICNP, which is
formed by nerve fibers entering the epicardium and form-
ing 400 ganglia around the junction between the right
atrium and the superior vena cava (384). Pauza et al. (384)
have estimated that the canine SAN can be innervated by
more than 54,000 intracardiac neurons residing in the INP.
A similar organization of SAN innervation by the INP has
been found in humans (385). The rat AVN and common
bundle are innervated by a dense network of thin fibers
projecting from a neuronal cluster adjacent to the inter-
atrial septum and the right pulmonary sinus (26). The
distribution of sympathetic and parasympathetic fibers in
the SAN is heterogeneous (375) and can slightly vary
between individuals (385), as well as on an age-dependent
way (26, 385). Vagal and sympathetic activation induce a
shift of the leading pacemaker site (54). Compared with
the atrial myocardium, the SAN is enriched in adrenergic
and muscarinic receptors (29). Similar to SAN ion chan-
nels, adrenergic and muscarinic receptor densities vary
regionally and may contribute to pacemaker shift during
autonomic stimulation (300, 375). SAN nerves contain
between 5 and 15 axons and terminate as naked terminals
with varicosities containing the neurotransmitter release
site (419). Adrenergic and nonadrenergic release sites are
in close vicinity to each other (375, 419). Choate et al. (88)
reported that synaptic varicosities at cholinergic termi-
nals in the guinea pig SAN are in close proximity with the
membranes of SAN myocytes. According to Hirst et al.
(200), cholinergic terminals form specialized “neuromus-
cular” junctions with SAN myocytes.
Autonomic control of pacemaker activity in vivo is
based on concomitant input from sympathetic and para-
sympathetic limbs. However, the ratio between vagal and
sympathetic input varies in a species-dependent way
(375). The impact of sympathovagal balance on the heart
rate can be appreciated by comparing basal rates in dif-
ferent mammalian species in the presence and in the
absence of autonomic input. In mammals, basal heart
rates are inversely correlated with the body weight.
Smaller mammals such as mice and bats have fast heart
FIG. 21. Stimulation of the PKA-dependent signaling pathway by
cAMP or 0.1
M isoproterenol (ISO) stimulates intracellular Ca
2
re-
lease. A: this line scan image shows the increase in the frequency of
Ca
2
release in a permebilized rabbit SAN cell superfused with 10
M
cAMP. B: recordings of pacemaker activity (top panel) line scan image
(middle panel) and fluorescence (bottom panel, same protocol as in Fig.
20) in control conditions (B) and during superfusion with ISO (C). [From
Vinogradova et al. (510).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
951
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
rate, ranging from 500 beats/min in mice during daytime
(157) to 800–1,000 beats/min in flying bats (269). In con-
trast, the heart rate in medium-sized and larger mammals
can vary between 60 and 70 beats/min in humans and 20
beats/min in whales (see Ref. 375 for review). Isolated
hearts and SAN pacemaker cells also show a wide range
of basal rates and maintain the same rate-to-animal
weight ratio as in the presence of an autonomic input
(375). The intrinsic properties of pacemaker cells in dif-
ferent species are one of the bases of the variability of
heart rate in mammals. However, the species-dependent
balance between sympathetic and parasympathetic input
also contributes to this wide range of pacemaking fre-
quencies (375). The two branches of the autonomic ner-
vous system interact to generate an adaptable equilibrium
so that SAN automaticity can be under dominance of the
sympathetic or parasympathetic limb. Sympathovagal
dominance is generally assessed in vivo by pharmacolog-
ical inhibition of the autonomic input by combined injec-
tion of atropine and propranolol. Even if propranolol does
not block
-adrenergic receptors, the heart rate measured
under these conditions constitutes a reliable estimate of
the intrinsic pacing rate of a “denervated” heart (29).
In animals, the beating frequency of the isolated SAN
is also an index of the intrinsic heart rate. In the mouse,
the intrinsic SAN rate is significantly lower than the
in vivo heart rate, indicating the existence of a significant
sympathetic tone in this species (157). In contrast, it can
be shown that dogs and humans are under prominent
vagal tone, since pharmacological block of the autonomic
input significantly accelerates the basal heart rate (223,
375). However, adrenergic dominance does not demon-
strate the absence of a vagal tone. For example, the
presence of vagal tone in small rodents can be easily
demonstrated by injection of atropine in freely moving
mice (157).
The effect of parasympathetic input on pacemaking
is greater when the SAN is under sympathetic tone. This
phenomenon has been named “accentuated antagonism”
by Levy (284). At the organ level, presynaptic regulation
of catecholaminergic terminals by the vagus nerve and
pacemaker shift contribute to accentuated antagonism
(300). In this respect, the functional inhomogeneity of
SAN tissue can be an important factor underlying accen-
tuated antagonisms (300). In SAN and AVN cells of the
rabbit, accentuated antagonism has been explained by the
regulation of cAMP levels by nitric oxide (NO) (see be-
low). Beside the basal regulation of the heart rate exerted
by the autonomic balance, cardiac automaticity is modu-
lated on a beat-by-beat basis so that pacemaking is
quickly adapted to the physiological state of the organism.
In this respect, the high-frequency (HF) and low-fre-
quency (LF) spectra of the heart rate variability (HRV)
have been used to study the dynamic regulation of pace-
making (157). However, it is becoming clear that intrinsic
factors are also involved in the beat-by-beat regulation of
heart rate. Indeed, it has been shown that “nonautonomic
mechanisms” may contribute to HF spectra of HRV in
some physiological conditions (34, 76, 463). Bernardi et al.
(34) reported that, at peak exercise, in both healthy hu-
man subjects and transplanted patients the HF spectra of
the HRV is almost completely generated by a “nonauto-
nomic” mechanism that is synchronized with ventilation.
These results were then confirmed by Casadei et al. (76),
who reported that under exercise, the “nonneuronal”
component of heart rate regulation increases by 35%.
Slovut et al. (463) have suggested that this phenomenon
is due to SAN stretch during diastolic atrial filling (see
sect. IXD).
B. Sympathetic Regulation of Pacemaker Activity
Activation of the
-adrenergic receptor underlies the
positive chronotropic effect induced by catecholamines
on automaticity. Catecholamines enhance the activity of
ion channels as well as intracellular Ca
2
release. The
relative importance of sarcolemmal ion channels and
LCICR in the
-adrenergic regulation of pacemaker activ-
ity is still debated. In this section, we will present and
compare evidence linking these mechanisms to stimula-
tion of pacemaker activity by catecholamines.
I
f
and I
Ca,L
have been proposed to constitute impor-
tant mechanisms in heart rate acceleration by cat-
echolamines (64, 67, 120, 364). The open probability of
f-channels increases even for a small augmentation of
intracellular cAMP. A rise in cAMP positively shifts the I
f
activation curve, thereby supplying more inward current
during the linear part of diastolic depolarization (120)
(Fig. 22). DiFrancesco (120) has proposed that I
f
is the
predominant mechanism for increasing the slope of the
diastolic depolarization at low adrenergic tone. This is
based on the observation that low doses of the
-adren-
ergic agonist isoproterenol (which is supposed to mimic
weak adrenergic activity) increase the slope of the dia-
stolic depolarization without affecting the action poten-
tial waveform (120). Remarkably, specific regulation of
the diastolic depolarization slope is a common property of
low doses of autonomic agonists and selective I
f
blockers
(21). Another strong line of experimental evidence for the
capability of f-channels to stimulate pacemaker activity is
the acceleration by cAMP analogs of pacing rate in SAN
cells. Bucchi et al. (67) have shown that isoproterenol and
Rp-cAMP similarly increase the diastolic depolarization
slope in rabbit SAN cells. Stieber et al. (466) also gave
strong evidence for the ability of f-channels to accelerate
embryonic heart rate. These authors have shown that
cAMP cannot accelerate the heartbeat of HCN4
/
em-
bryos. Overall, different lines of direct and indirect evi-
dence are in favor of an important contribution of f-
channels in sympathetic regulation of pacemaker activity.
952
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
The functional role of I
Ca,L
in adrenergic regulation of
heart rate is still unresolved. Three lines of indirect evi-
dence are suggestive of a role of I
Ca,L
in the sympathetic
regulation of heart rate. First, catecholamines robustly
enhance I
Ca,L
in the same concentration range as pace-
maker activity (542). Zaza et al. (542) have shown that I
f
and I
Ca,L
have similar sensitivity to isoproterenol in rabbit
SAN cells. Second, Choate and Feldman (87) have re-
ported that DHPs reduce the positive chronotropic re-
sponse of mouse atria to stimulation of the stellate gan-
glion. Third, the contribution of Ca
v
1.3 channels in the
generation of the diastolic depolarization in mouse SAN
cells suggests that I
Ca,L
can constitute an important mech-
anism for accelerating the diastolic depolarization slope
upon activation of
-adrenergic receptors (309, 310). Con-
sistent with this hypothesis, the positive chronotropic
response to isoproterenol of isolated Ca
v
1.3
/
hearts is
moderately reduced (321). However, the in vivo heart rate
of mice lacking Ca
v
1.3 channels is comparable to that of
wild-type mice (393). The maximal heart rates of wild-
type and Ca
v
1.3
/
mice are significantly reduced by se-
lective I
f
block by ivabradine (Mangoni et al., unpublished
observations). Taken together, these observations are
suggestive of a distinct role of f-channels and Ca
v
1.3
channels in the autonomic regulation of heart rate (see
also Ref. 196).
The relevance of TTX-sensitive and TTX-resistant I
Na
in the sympathetic regulation of heart rate is of interest,
yet no reports have specifically studied the autonomic
regulation of I
Na
in the SAN. The cardiac TTX-resistant
Scn5A-mediated I
Na
is sensitive to phosphorylation by
PKA and PKC (286, 407). Taking into consideration the
relevance of Scn5A channels in SAN conduction, we can
speculate that sympathetic stimulation of TTX-resistant
I
Na
can accelerate heart rate by accelerating impulse con-
duction within the SAN as well as from the SAN to the
atrium. In contrast, the effects of autonomic agonists on
TTX-sensitive I
Na
are not easily predictable. Maier et al.
(304) have hypothesized that TTX-sensitive Na
channels
can be negatively regulated by the sympathetic nervous
system as in neurons. If this hypothesis is valid, the
contribution of TTX-sensitive I
Na
to the diastolic depolar-
ization may be higher in basal conditions than under
strong adrenergic activation.
In pacemaker cells of amphibians (227) and mam-
mals,
-adrenergic receptor activation stimulates SR Ca
2
release and NCX activity (415, 506). In spontaneously
active rabbit SAN cells, isoproterenol robustly enhances
the amplitude and frequency of RyR-dependent LCICR
(415, 506). Vinogradova et al. (506) have reported that
ryanodine abolishes the augmentation of subsarcolemmal
Ca
2
release and LCICRs. This effect is accompanied by a
strong reduction in the isoproterenol-induced positive
chronotropic effect, particularly at low agonist doses.
Vinogradova et al. (506) have thus proposed that SR Ca
2
release is the major effector of the positive chronotropic
effect of the
-adrenergic receptor pathway. Partial inhi-
bition of the isoproterenol chronotropic effect in rabbit
SAN has been qualitatively confirmed by other authors
(67, 267, 415), although it shows quantitative variability.
FIG. 22. A: low doses of isoproterenol and ACh selectively modulate the slope of the diastolic depolarization in rabbit SAN cells (a) and shift
the I
f
voltage dependence (b, c). These experiments have led DiFrancesco (see Ref. 121 and sect. VII for discussion) to propose that the predominant
mechanism by which the autonomic nervous system controls the heart rate is I
f
.[a from DiFrancesco (120), with permission from the Annual Review
of Physiology; b and c are from Accili and DiFrancesco (4).] B: a cAMP analog (R
p
-cAMPs) stimulates pacemaker activity by increasing the slope
of the diastolic depolarization under control conditions and in the presence of ryanodine (Ry). This observation suggests that impairment of Ca
2
release from the SR does not prevent regulation of pacemaker activity by direct binding of cAMP analogs to f-channels. [From Bucchi et al. (67),
with permission from Elsevier.]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
953
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
Indeed, in a previous study using intact SAN, ryanodine
reduced
-adrenergic stimulation of pacemaking by 40%
(415). Pacemaking in the intact SAN is also less sensitive
to ryanodine than in isolated pacemaker cells. Musa et al.
(349) have indicated that such a difference can be due to
heterogeneous expression of proteins regulating Ca
2
handling in the center versus the periphery of the SAN.
Functionally, the central leading pacemaking site is less
sensitive to inhibition of SR Ca
2
release (267). However,
Lyashkov et al. (299) have reported that, in isolated rabbit
SAN cells, the effects of ryanodine, BAPTA, and Li
on
pacemaker activity are similar irrespective of the cell size.
Discrepancy between results obtained in intact SAN tis-
sue and isolated cells may be attributed to the network
organization of pacemaker cells in the SAN.
In Figure 23, we attempt to summarize the down-
stream targets of the
-adrenergic receptor-dependent
signaling pathway in SAN pacemaker cells. This model is
based on evidence from rabbit SAN cells and is comple-
mented by recent insights from genetically modified
mouse strains. For our discussion, we can distinguish two
distinct pacemaker mechanisms: an “ion channel clock”
formed by voltage-dependent ion channels and the intra-
cellular SR-dependent “Ca
2
clock” (307). Activation of
-adrenergic receptors stimulates adenylyl cyclase (AC)
activity, which converts ATP in cAMP. Elevated cAMP
promotes voltage-dependent opening of f-channels and
activates PKA (126). The catalytic subunit of PKA en-
hances the activity of different ion channels of the mem-
brane-delimited pacemaker mechanism by channel phos-
phorylation. These include Ca
v
1.3, Ca
v
1.2 channels (309),
as well as st-channel complexes (334). Vinogradova et al.
(508) have reported that basal PKA activity is higher in
rabbit SAN cells than in working myocytes. High phos-
phorylation of PLB and RyRs generates periodical LCICR
causing I
NCX
-mediated membrane voltage oscillations,
which contribute to the control of the chronotropic state
of the cell (43, 508).
-Adrenergic receptor activation
further elevates PKA activity, thereby increasing the fre-
quency and number of LCICRs and stimulating I
NCX
.
The I
f
-based” pacemaker mechanism (121) and that
of the spontaneous SR-dependent “Ca
2
clock” (307) have
been placed as opposed to each other. However, they can
be qualitatively reconciled in a general framework. Both
models have a common major messenger that is cAMP.
cAMP can control the chronotropic state of the cell by at
least three distinct effectors: 1) f-channels by direct chan-
nel opening, 2)Ca
v
1.3 channels, and 3) RyRs by channel
FIG. 23. This cartoon summarizes the ionic mechanisms contributing to the diastolic depolarization in a SAN pacemaker cell. Voltage-dependent
ion channels as well as the proposed “Ca
2
clock” (307) are represented together. Possible interactions between these mechanisms are indicated.
In pacemaker cells, high basal cAMP-mediated PKA-dependent phosphorylation stimulates a perpetual “free running” Ca
2
cycling by pumping Ca
2
into the SR via SERCA2 and LCICR via RyRs. PKA also stimulates Ca
2
entry through Ca
v
1.3-mediated I
Ca,L
. The thick red line indicates the
persistent spontaneous Ca
2
cycling. The possibility that Ca
v
3.1-mediated I
Ca,T
can contribute to replenishment of SR Ca
2
stores is suggested.
Spontaneous LCICR from SR is linked to the diastolic depolarization via Ca
2
activation of inward I
NCX
current. Direct cAMP-dependent activation
of HCN channels or cAMP-mediated, PKA-dependent phosphorylation of Ca
v
1.3 channels and I
st
(dashed lines) strongly stimulates the pacemaker
cycle driven by the “membrane ion channels clock” (MCC). It is conceivable that the MCC can entrain the intracellular Ca
2
clock because of the
dependency of SR Ca
2
content from VDCCs. On the other hand, the Ca
2
clock may trigger oscillations of the membrane voltage and initiate normal
pacemaking via the MCC. It is thus likely that under physiological conditions the MCC and the Ca
2
clock mutually entrain one another. [Adapted
from Vinogradova et al. (508).]
954 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
phosphorylation. Experimentally, maneuvers altering in-
tracellular cAMP levels will affect the activity of all these
effectors. One may wonder why the “ion channels clock”
and the “free-running Ca
2
clock” (307) share a common
messenger. The answer probably resides in the necessity
to synchronize the replenishment of SR Ca
2
stores to
Ca
2
release in the late phase of the diastolic depolariza-
tion. Spontaneous LCICR ceases in arrested cells, since
Ca
v
1.3 channels are probably the predominant source of
Ca
2
entry during the action potential. Consequently, as
much as the “Ca
2
clock” accelerates under the action of
catecholamines, the sarcolemmal pacemaker mechanism
mediated by f- and Ca
v
1.3 channels will also speed up and
shorten diastolic depolarization.
However, as highlighted by Maltsev et al. (307), the
“Ca
2
clock” hypothesis of the genesis of SAN automatic-
ity assumes that spontaneous LCICR is the predominant
mechanism that controls the chronotropic state of SAN
cells in both basal conditions and under
-adrenergic
receptor stimulation. These authors have proposed that I
f
plays only a minor role in mediating the effect of cat-
echolamines on heart rate. This view is contradicted by a
substantial set of experimental results, such as the obser-
vation that ryanodine does not prevent acceleration of
pacemaker activity in rabbit SAN cells by cAMP analogs
(67, 68) and that specific f-channel inhibition by ivabra-
dine can reduce the heart rate of freely moving mice by up
to 26% (Mangoni et al., unpublished observations). Fur-
thermore, the importance of f-channels in the autonomic
regulation of heart rate is strongly supported by inherited
dysfunction of pacemaking in humans harboring muta-
tions in the HCN4 gene (see sect. XA) and related murine
models (see sect. VID).
C. Parasympathetic Regulation of Pacemaking
The parasympathetic regulation of cardiac automa-
ticity is mediated by the activation of muscarinic recep-
tors following release of acetylcholine (ACh) from vagal
nerve endings. Cholinergic agonists induce a potent neg-
ative chronotropic effect on cardiac automaticity and
atrioventricular conduction in vivo as well as in isolated
heart preparations (215, 325). Signaling and ionic mecha-
nisms underlying the muscarinic regulation of heart rate
have been studied for more than 30 years, yet it is not
completely understood which mechanisms determine re-
sponses in heart rate under various physiological condi-
tions, such as high sympathetic tone or parasympathetic
discharge.
Hutter and Trautwein (212) observed that vagal stim-
ulation stops pacemaking in the frog sinus venosus. As
suppression of automaticity was due to an increase in
membrane K
conductance, Hutter and Trautwein (212,
213) concluded that the vagal control of the heart rate was
due to activation of a K
current. Early observations by
Hutter and Trautwein are now accounted for by the acti-
vation of I
KACh
channels via vagally released ACh (432).
However, other authors have failed to observe hyperpo-
larization of the membrane potential after vagal stimula-
tion. For instance, studies by Toda and West (481, 519)
reported that vagal stimulation slowed automaticity by
reducing the slope of diastolic depolarization in atrial and
AVN preparations. This has been confirmed by Shibata et
al. (453) on isolated rabbit SAN preparations. Both hyper-
polarization of the maximum diastolic potential and a
decrease in the slope of diastolic depolarization are ob-
served when the muscarinic regulation of pacemaker ac-
tivity is studied using isolated spontaneously active cells.
Indeed, in isolated rabbit SAN cells, high ACh doses (1
M) hyperpolarize the maximum diastolic potential and
eventually stop pacemaking due to maximal I
KACh
activa-
tion (124, 507). In contrast, low ACh doses (1–10 nM) slow
pacemaker activity by decreasing the slope of diastolic
depolarization in the absence of other changes in action
potential parameters and maximum diastolic potential
(Fig. 22A). This is because binding of ACh to muscarinic
receptors activates different signaling pathways in pace-
maker cells. These pathways involve direct activation of
I
KACh
channels, negative regulation of cAMP production,
and positive regulation of cAMP hydrolysis.
The muscarinic M
2
receptor activates the inhibitory
G protein
-subunit (
i
) which negatively couples to AC
activity (283, 440). Downregulation of cAMP can affect
the activity of voltage-dependent ion channels involved in
pacemaker activity and reverse the signaling processes
involved in sympathetic stimulation of heart rate (Fig. 24).
Moreover,
␤␥
-subunits of
i
directly open K
ACh
channels
(520, 523). The role of I
KACh
, I
f
, and I
Ca,L
in mediating the
negative chronotropic effect of ACh has been studied in
isolated SAN pacemaker cells. Recently, insight into the
importance of I
KACh
in the regulation of heart rate has
been obtained using genetically modified mice lacking
K
ir
3.4 channels (522). Wickmann et al. (522) have shown
that K
ir
3.4
/
mice lack I
KACh
in the atrium and show
prominent reduction autonomic regulation of heart rate in
the HF and LF spectrum of the HRV. Consistent with
these observations, Gehrmann et al. (158) have shown
that transgenic mice expressing low amounts of
␤␥
have
a reduced negative chronotropic response to cholinergic
agonists.
DiFrancesco et al. (124) have studied the relative
sensitivity to ACh of I
f
and I
KACh
in rabbit SAN cells.
These authors have compared the ACh dose dependency
of the shift in the I
f
activation curve, the I
KACh
density,
and the slowing of pacemaker activity. In this study, low
ACh doses significantly shifted the I
f
activation curve in
the negative direction and slowed pacemaking in a con-
centration range that did not activate I
KACh
. Using a sim-
ilar approach, Zaza et al. (542) have compared the sensi-
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 955
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
tivity of I
Ca,L
and I
f
to ACh. In this study, I
f
inhibition was
observed at lower doses of ACh than that required to
inhibit “basal” I
Ca,L
(542). These results have led Di-
Francesco (119) to propose that I
f
is the predominant ion
channel which underlies the muscarinic control of heart
rate at low vagal tone. This view is consistent with recent
evidence obtained by Yamada in isolated beating mouse
hearts (530). In these experiments, heart rate reduction
induced by low ACh doses was reported to be insensitive
to I
KACh
inhibition by tertiapin. Caution should be used
when transposing in vitro observations from isolated
pacemaker cells and denervated hearts to in situ heart
rate regulation. Due to the accentuated antagonism phe-
nomenon, different ionic mechanisms can participate to
muscarinic regulation of pacemaker activity, according to
their relative sensitivity in cholinergic agonists as well as
to intracellular levels of cAMP. The cardiac I
Ca,L
has been
reported to be remarkably “insensitive” to muscarinic
regulation in basal conditions, because relatively high
ACh doses are required to inhibit this current (190, 323).
However, Petit-Jacques et al. (390) have shown that mus-
carinic regulation of SAN I
Ca,L
depends on the previous
-adrenergic stimulation and cAMP levels. Indeed, mod-
erate doses of ACh can significantly inhibit I
Ca,L
if previ-
ously stimulated by
-adrenergic agonists (390). PKA in-
hibition or cell dialysis with a nonhydrolyzable cAMP
analog abolishes I
Ca,L
regulation by ACh (390). ACh can
thus act as an “antiadrenergic” agent. Interestingly, accen-
tuated antagonism of ACh after
-adrenergic receptor
activation has been observed also on I
f
in canine Purkinje
fibers (81). It is thus important to distinguish between
“direct” and “indirect” cholinergic regulation of I
Ca,L
,to
distinguish between basal current inhibition and that ob-
served after adrenergic stimulation. Thus it cannot be
excluded that indirect (or antiadrenergic) inhibition of
I
Ca,L
can be relevant in vagal regulation of heart rate in
conditions of tonic sympathetic activation, when the ac-
centuated antagonism phenomenon is present. The regu-
lation of Ca
v
1.3 channels by ACh has not yet been inves-
tigated. In this respect, the relatively positive holding
potentials (about 30 mV) employed thus far to study the
regulation of I
Ca,L
by autonomic agonists (186, 542) would
inactivate most of Ca
v
1.3 channels (309).
Han et al. (186) have demonstrated that an NO sig-
naling pathway is involved in the muscarinic regulation of
I
Ca,L
in conditions of accentuated antagonism, when in-
FIG. 24. Summary of the signaling pathways involved in the muscarinic regulation of pacemaker activity. In the SAN and AVN, ACh and Ado
share the same pathways for signal transduction. ACh binds to the muscarinic M
2
receptor, which is coupled to a “direct” G protein-dependent
pathway activating Kir3.1/Kir3.4 (GIRK1/GIRK4) channel complexes. Beside this direct pathway, two other “indirect” channel regulation pathways
lead to downregulation of intracellular cAMP. In the first cascade of events, the
G
i
protein subunit inhibits AC activity, thereby reducing the
synthesis of cAMP. Inhibition of AC synthesis is viewed as the predominant pathway by which ACh regulates the voltage dependence of f-channels
(HCN). The second muscarinic pathway is initiated by the stimulation of the NOS activity and production of NO. NOS activates the enzyme guanylate
cyclase (GC), which converts GTP in cGMP. Elevation of cGMP concentration stimulates PDE II-dependent hydrolysis of cAMP, thereby reducing
PKA activity. This NOS-dependent pathway for regulation of PKA activity has been proposed to be the major pathway for the muscarinic regulation
of I
Ca,L
during pacemaking. In the figure, negative NOS-dependent regulation of Ca
v
1.3-mediated I
Ca,L
and I
st
is suggested. It is not known if the SAN
I
Kr
can also be negatively regulated by the NOS-dependent pathway.
956 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
tracellular cAMP is increased by stimulation of
-adren-
ergic receptors. In cardiac myocytes, muscarinic receptor
activation is coupled to NO synthesis, which stimulates
guanylyl cyclase activity. Elevated cGMP production pro-
motes phosphodiesterase activity that inhibits I
Ca,L
by
cAMP breakdown (323). Han et al. (186) have provided
evidence that the NO signaling pathway plays an obliga-
tory role in the muscarinic regulation of rabbit SAN I
Ca,L
(186). In these cells, block of NO synthesis prevents I
Ca,L
inhibition by cholinergic agonists such as carbamylcho-
line (185). This signaling pathway regulating I
Ca,L
is
present also in the AVN (170). The NO signaling pathway
in rabbit SAN seems to be primarily mediated by a con-
stitutive endothelial NO oxide synthase (eNOS) isoform
(182). Also, specific pharmacological inhibition of pos-
phodiesterase II (PDEII) blocks NO-dependent inhibition
of I
Ca,L
, indicating that this PDE isoform plays a dominant
role in downstream regulation of cAMP level (182). In
contrast, the role of intracellular [Ca
2
]
i
in mediating
NO-dependent inhibition of I
Ca,L
would need further in-
vestigation. Han et al. (186) have shown that buffering
[Ca
2
]
i
with BAPTA blocks eNOS activity. However, in
the same study, depletion of SR Ca
2
stores by ryanodine
and thapsigargin did not affect regulation of I
Ca,L
, suggest-
ing that Ca
2
involved in modulation of eNOS activity
comes from an independent intracellular compartment.
The relevance of the eNOS/PDEII signaling pathway in
mediating accentuated antagonism in the canine heart has
been shown by Sasaki et al. (437). These authors have
also shown that NO can partially modulate accentuated
antagonism but does not play an important role in the
absence of previous
-adrenergic receptor stimulation.
Beside the eNOS/PDEII signaling pathway, protein
phosphatases can also regulate the activity of ion chan-
nels under basal conditions or under the action of auto-
nomic agonists. Ke et al. (235) have reported that the
p21-activated protein kinase (Pak1) is functionally ex-
pressed in guinea pig SAN cells. In pacemaker cells, when
cellular Pak1 activity is enhanced by expressing a consti-
tutively active form of the protein, I
Ca,L
and I
K
densities
are reduced and the positive chronotropic response of
cells to isoproterenol is blunted. The regulatory role of
active Pak1 on I
Ca,L
and I
K
is likely attributable to in-
creased activity of the protein phosphatase PP2A (235). It
remains to be established if Pak1 can be functionally
associated with the vagal regulation of heart rate.
The importance of NO signaling pathway in the vagal
regulation of heart rate in mice has not been clearly
demonstrated. Mice lacking eNOS or the G
o
protein
isoforms lack muscarinic regulation of I
Ca,L
(183, 487).
However, Vandecasteele et al. (494) failed to observe
reduction of both adrenergic and muscarinic regulation of
I
Ca,L
in another eNOS-deficient mouse line. Consistent
with these results, Mori et al. (343) have reported that
inhibition of eNOS does not change accentuated antago-
nism in isolated mouse atria.
VIII. CARDIAC AUTOMATICITY AS AN
INTEGRATED MECHANISM: NUMERICAL
MODELING OF PACEMAKER ACTIVITY
A. General Models of Automaticity
Modeling of cardiac automaticity integrates the be-
havior of channels and ionic homeostasis in a calculating
environment to predict pacemaking under different con-
ditions. Numerical models are used to gain insight into the
roles of ion channels and/or Ca
2
handling proteins, when
biophysical or pharmacological approaches are limited or
impossible. The growing knowledge on the cardiac pace-
maker mechanism is reflected by the increasing complex-
ity of numerical models of automaticity. A detailed com-
parison of all the numerical models of automaticity that
have been developed in the past is beyond the scope of
this review. For further discussion on the historical and
technical aspects of this topic, the reader can consult the
recently published manuscript by Wilders (525). Here, we
will discuss some predictions of numerical models on
the importance of ion channels and Ca
2
signaling in the
generation and regulation of automaticity.
Recent cellular models of SAN automaticity have
been developed from the DiFrancesco and Noble model
of Purkinje fibers automaticity (127). In this model, auto-
maticity is based on I
f
. Suppression of I
f
abolishes auto-
maticity (127). The DiFrancesco and Noble model in-
cluded calculations of intracellular systolic Ca
2
tran-
sients and I
NCX
activity. Noble and Noble (360) have
adapted the DiFrancesco and Noble model to the central
and peripheral SAN automaticity, by scaling current den-
sities to approximate that of SAN cells. Noble et al. (359)
have used this model to predict the possible roles of I
f
and
that of Na
background current [I
b(Na)
] and proposed that
I
f
is a major mechanism for stabilizing automaticity and to
protect SAN pacemaking from hyperpolarizations.
Wilders et al. (526) have also assessed the role of I
f
and
I
b(Na)
in the generation of the diastolic depolarization.
These authors have predicted that the association of I
f
and I
b(Na)
is necessary for pacemaking of isolated rabbit
SAN cells, since joint abolition of both currents stops
automaticity (526). Block of I
f
slows pacemaking by spe-
cifically reducing the slope of the diastolic depolarization
(526). Wilders et al. have suggested that their modeling
results are consistent with the experimental work by
Denyer et al. who proposed that I
b
can sustain pacemak-
ing under I
f
block (113, 526) (see sect. VID). Demir et al.
(111) have developed a model of automaticity of periph-
eral SAN cells, by associating voltage-dependent ion chan-
nels activity with intracellular Ca
2
release and handling
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 957
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
by intracellular Ca
2
buffers. In this model, different ionic
currents contribute to the generation of the diastolic de-
polarization. Particularly, block of I
f
and I
Ca,T
almost
equally slow pacemaker activity, while inhibition of I
Ca,L
mainly affects the late diastolic depolarization and action
potential upstroke (111). Predictions obtained by the
Demir et al. model are generally consistent with many
physiological and pharmacological results on several
ionic currents involved in automaticity, including I
Ca,T
,
I
Ca,L
, I
Kr
, and I
f
(111). The Demir et al. model can suc-
cessfully reproduce certain aspects of the parasympa-
thetic regulation of pacemaker activity. Particularly, the
negative regulation of pacemaking by specific reduction
of the slope of the diastolic depolarization, and the phase-
resetting SAN response upon bursts of vagal activity can
be computed by this model (110). More generally, predic-
tions obtained by the Wilders et al. and Demir et al.
models are qualitatively consistent with experimental re-
sults in rabbit SAN cells. However, they have some limi-
tations when inferring the role of I
Ca,L
and intracellular
Ca
2
release. These limitations are justified as far as these
models have been developed when experimental data on
Ca
v
1.3 channels and LCICR were not available. Kurata
et al. (262) have been the first to develop a model of
SAN automaticity in which the subsarcolemmal Ca
2
([Ca
2
]
sub
) and the SR Ca
2
content were included in
calculations. Furthermore, buffering of [Ca
2
]
i
by tropo-
nin, calsequestrin, and calmodulin were also included
(262). The Kurata et al. model (262) thus aims to combine
the advantages of calculating Ca
2
buffering and Ca
2
exchange between the subsarcolemmal space and the SR.
Subsarcolemmal Ca
2
controls Ca
2
-dependent inactiva-
tion of I
Ca,L
(262). In the Kurata et al. model, block of SR
Ca
2
release has only minor effects on pacemaking, es-
sentially because Ca
2
release is limited to the action
potential upstroke phase.
B. Dedicated Models of Automaticity
Recently developed models of SAN pacemaking con-
sider specific SAN properties, such as the regional varia-
tions of ion channel expression (545), regulation of auto-
maticity by LCICR (306), and automaticity in genetically
modified mice (314). These models will be referred to
here as “dedicated.”
Zhang et al. (545) have modeled pacemaker activity
in the periphery and center of the SAN. This model is
based on experimental results on the heterogeneity of cell
size and ionic current densities observed in the rabbit
SAN (58) (see sects.
II and III). Numerical simulations by
Zhang et al. (545) show that it is possible to reproduce the
slower pacemaker activity, action potential upstroke ve-
locity, and the more positive maximum diastolic potential
in the SAN center by setting lower densities of I
Ca,L
, I
f
, I
Kr
,
and I
Na
in the central than in the peripheral SAN model. In
the central model, I
Na
is set to zero so that the action
potential upstroke phase is entirely controlled by I
Ca,L
.In
the peripheral model, the diastolic depolarization is car-
ried by I
Ca,T
and I
f
, while in the central model I
Ca,T
pre-
dominates on I
f
. The model predicts the higher effect of I
f
blockers on automaticity in the SAN periphery versus the
center. Suppression of I
Ca,L
in the Zhang et al. model stops
pacemaking in the center and sets the membrane poten-
tial to 35 mV (545). In the peripheral model, block of
I
Ca,L
moderately accelerates automaticity possibly be-
cause of the shortening of the action potential duration.
I
Kr
suppression stops automaticity in both the peripheral
and central SAN model. However, it is possible to set a
20% block of I
Kr
without stopping pacemaking. According
to Zhang et al. (545), the moderate slowing of pacemaking
induced by partial inhibition of I
Kr
is consistent with what
is observed in peripheral SAN tissue balls (249). In the
peripheral SAN model, I
to
has higher density than in the
central model. Consistent with experimental results (60),
suppression of I
to
in the peripheral model prolongs the
action potential duration, while has limited effects on the
central SAN model (545). Garny et al. (152) have adapted
the Zhang et al. model in an attempt to reproduce atrio-
sinus interactions and calculate the predicted position of
the SAN leading pacemaker site in the presence and ab-
sence of the sorrounding atrium (Fig. 25). In this model, a
gradual transition between cells matched with the central
and the peripheral model was established via a distribu-
tion function. A variable ratio of intercellular conductivity
was also implemented (see legend of Fig. 25). This cellu-
lar array was connected to cells matched with a rabbit
(197) or human (368) atrial model. The position of the
leading pacemaker site settled in the SAN center when a
greater intercellular coupling between cells in the SAN
periphery than in the center was established. The leading
pacemaker site shifted to the periphery when intercellular
conduction was uniform in the SAN center and periphery.
Finally, the leading pacemaker site was located in the
periphery when the SAN cellular array was electrically
disconneted from the atrium (Fig. 25). In conclusion, this
atriosinus model successfully reproduced the electro-
tonic inhibition of atrial cells on SAN pacemaking. Garny
et al. (152) have interpreted the electrotonic suppression
of pacemaking in the SAN periphery in terms of an “op-
posing depolarization,” rather than a simple hyperpolar-
izing inhibition (152).
In its first version, the model by Zhang et al. did not
include calculations of intracellular Ca
2
handling. Simi-
larly [Na
]
i
and [K
]
i
were kept constant (58). In a follow
up manuscript, these authors have included calculations
of Ca
2
handling in both the central and peripheral SAN
model (56). Available experimental data on Ca
2
-depen-
dent effects on I
Ca,T
, I
Ca,L
, I
Kr
, and I
f
were included.
Boyett et al. (56) have then compared the effects of
958
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
blocking Ca
2
transients in this model and that by Noble
and Noble (360), Wilders et al. (526), and Demir et al.
(111). In all models tested, abolition of Ca
2
transients
variably slowed pacemaking, from a minimum of 9% in the
Wilders et al. model and a maximum of 41% in the Demir
et al. model. Calculations by Boyett et al. suggested that
this slowing of automaticity was predominantly due to
changes in the Ca
2
-dependent activity of ion channels
rather then I
NCX
. These calculations are very useful for
interpreting the possible physiological consequences of
the Ca
2
dependency of ion channel activities, but do not
include LCICR and its functional coupling with I
NCX
(see
sect.
VIL).
Maltsev et al. (306) have modified the Kurata et al.
(262) model to include spontaneous preaction potential
LCICR and activation of I
NCX
. These authors have inter-
preted spontaneous LCICR in terms of a periodical global
cellular release phenomenon of a variable phase (and
amplitude), occurring during the diastolic interval. In this
model, a global LCICR eventually induces the exponential
part of the diastolic depolarization. Shortening of the
interval between successive LCICRs is predicted to aug-
ment the frequency of pacemaking of rabbit SAN cells, by
accelerating the onset of the exponential late phase of the
diastolic depolarization. Vinogradova et al. (508) and Bog-
danov et al. (43) further developed this model by dividing
the global LCICR into the sum of finite release elements of
which the individual phase of a given release event is set
by a random generator (Fig. 26A). In this model, a release
event can occur at any time comprised between the mean
and the standard deviation of the phase of LCICRs ob-
served experimentally. This modification allows one to take
into account the “roughly periodical” nature of LCICR in
basal conditions and to simulate synchronization of Ca
2
release units under activation of the
-adrenergic recep-
tor (508). Bogdanov et al. (43) have used this model to
calculate fluctuations of I
NCX
under basal conditions and
stimulation of pacemaking by isoproterenol (Fig. 26, A
and B). In conclusion, numerical modeling predicts that
subsarcolemmal Ca
2
sigaling may have little effect on
pacemaking when only systolic SR Ca
2
release is con-
sidered (262). However, models including diastolic preac-
tion potential Ca
2
release show that LCICR significantly
accelerates pacemaking (43, 306, 508). At present, no
model including both LCICR and Ca
2
dependency of ion
channels is available.
Mangoni et al. (314) have adapted the Zhang et al.
(545) central model to predict the activation of Ca
v
1.3
and Ca
v
3.1 channels during the diastolic depolarization
in mouse SAN pacemaker cells (Fig. 26C). Simulations
obtained using this model suggest that, in the mouse
SAN, I
Kr
controls the action potential repolarization
phase and is present as a relatively large outward com-
ponent throughout the diastolic depolarization (see
sect.
VIC). I
f
is the first time-dependent current to be
FIG. 25. Numerical modeling of SAN-atrium interactions using a
modified version of the Zhang et al. model of rabbit central and periph-
eral SAN automaticity (152). The membrane voltage of the leading
pacemaker site is depicted in green, the “follower” SAN membrane
voltage in red, and the atrial membrane voltage is depicted in blue. In A,
the central SAN coupling conductance is set 10 times lower than the
peripheral conductance (37.5 vs. 375 nS, see Ref. 152). Note that the
leading pacemaker site locates in the SAN center. In B, a uniformly low
coupling conductance has been set throughout the SAN model (37.5 nS).
In these conditions, the leading pacemaker site shifts to the SAN pe-
riphery. C: when the SAN and atrium are disconnected (no SAN-driven
action potential in the atrium), the leading pacemaker site shifts to the
SAN periphery, where the intrinsic cell beating rate is faster than in the
SAN center. D: simulation parameters has been set as in A, but a
numerical model of a human atrial cell has been employed. [From Garny
et al. (153), with permission from Elsevier.]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
959
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
activated at the end of the repolarization phase. Ca
v
3.1
channels activate upon depolarization before Ca
v
1.3
and TTX-sensitive Na
channels. According to this
model, Ca
v
3.1 channels can contribute to the linear part
of the diastolic depolarization. Ca
v
1.3 channels activate
close to the exponential fraction of the diastolic depo-
larization and may constitute the predominant voltage-
dependent mechanism contributing to this phase. This
prediction is consistent with experimental observations
indicating that DHPs mainly affect the late phase of the
diastolic depolarization (319). Ca
v
1.3 channels also
contribute to the upstroke phase of the action poten-
tial. Consistent with experimental observations in
mouse SAN cells (279), TTX-sensitive I
Na
appears to
contribute to the exponential fraction of the diastolic
depolarization, as observed by Lei et al. (279). I
st
is
present throughout the pacemaker cycle as a sustained
component.
IX. ADDITIONAL REGULATORS OF
CARDIAC AUTOMATICITY
A. Neuropeptides
Neuropeptides are released with ACh or norepineph-
rine at autonomic nerve terminals (297, 468). The vasoac-
tive intestinal polypeptide (VIP) is coreleased with ACh
and stimulates the synthesis of cAMP (91). VIP acceler-
ates heart rate in different species including monkeys
(107), dogs (412), and rats (418). It has been proposed
that VIP can contribute to postvagal tachycardia, because
at the end of vagal discharge ACh will be removed faster
from the synaptic space than VIP (195). Shvilkin et al.
(458) have reported that in Langendorff-perfused rat
hearts, vagally released VIP moderates the negative chro-
notropic effect induced by high-frequency vagal stimula-
tion. A direct positive chronotropic effect of VIP has been
FIG. 26. A: simulated effects of
-adrenergic receptor activation (ISO) on I
NCX
and membrane voltage noise during the diastolic depolarization.
Families of simulated LCICR events, total I
NCX
(I
NaCa
), and membrane voltage (V
m
) in control conditions (top panel), under ISO (middle panel), and
ryanodine (no LCRs, bottom panel). Predicted beating rates are shown in beats per min (bpm). B: overlapped representative cycles from simulations
as in A, showing changes in cycle length and I
NCX
amplitude and fluctuations. Inset depicts fluctuating I
NCX
at the initiation of the exponential phase
of the diastolic depolarization (indicated as boxes). The symbol
i
indicates initial changes in the NCX current related to individual LCRs. [From
Bogdanov et al. (43)] C: voltage-dependent “ion channels clock” generating automaticity is proposed here. The “clock” has been represented by using
a numerical computation of the mouse SAN automaticity according to Mangoni et al. (314). For each ionic current considered, the underlying ion
channel or channel gene family has been indicated. See the text (sect.
VIIIB) for further discussion.
960 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
reported by Rigel and Lathrop in dogs (413). These au-
thors have observed stimulation of pacemaker activity of
SAN and atrioventricular junctional pacemakers (414). A
positive shift of I
f
activation curve by VIP (5– 6 mV) has
been described by Chang et al. (82) in canine Purkinje
cells, and by Accili and DiFrancesco (3) in rabbit SAN
cells. This shift of the I
f
voltage dependence is probably
the predominant mechanism of action of VIP (3, 82).
B. Adenosine
Adenosine (Ado) is a paracrine and autocrine factor
regulating a plethora of cellular functions in the cardio-
vascular system. These include automaticity, coronary
circulation, cell adhesion and migration, angiogenesis,
and metabolism (346). The cardiovascular effects of Ado
have been intensively studied, due to its cardioprotective
effects in ischemic preconditioning and cardiac stunning
(345). Three distinct Ado receptors (AAR), namely, the
A1, A2, and A3AR are expressed in heart tissue (346).
Transgenic mice overexpressing A1ARs have significant
bradycardia and constitutive slowing of atrioventricular
conduction (242). Indeed, Ado is a potent negative regu-
lator of SAN pacemaker activity (517, 518) and AVN con-
duction (31). This latter property has led to the clinical
use of Ado for termination of reentrant supraventricular
tachycardias (32). In spontaneously active cells, AARs
share the signaling pathway of muscarinic receptors (Fig.
24). Ado mimics downstream effects of muscarinic recep-
tor activation, such as downregulation of cAMP synthesis
and activation of Kir3.1/Kir3.4 channels (I
KAdo
) (30, 543).
Under conditions of accentuated antagonisms, Ado re-
duces automaticity in guinea pig Purkinje fibers (281) and
rabbit SAN cells (30). Belardinelli et al. (30) have studied
the effects of Ado on I
f
and I
Ca,L
in rabbit SAN cells. They
reported that Ado consistently activated I
KAdo
, but inhib-
ited I
Ca,L
and I
f
only after stimulation by isoproterenol
(30). Shimoni et al. (456) have proposed that the indirect
inhibitory effect of Ado on I
Ca,L
is mediated by a NO-
dependent signaling pathway (Fig. 24). However, Zaza,
Rocchetti, and DiFrancesco (543) have reported that Ado
and ACh can shift the activation of “basal” I
f
(without
previous stimulation by isoproterenol) by 5and9mV,
respectively. In this study, low doses (0.03
M) of Ado
slowed automaticity of SAN cells without hyperpolarizing
the maximum diastolic potential. Reasons for the discrep-
ancy between these two studies are not known. Zaza et al.
(543) have emphasized that Ado can slow the heart rate
by specific regulation of I
f
(543).
C. Hormones
Some hormones can influence pacemaking by regu-
lating ion channels involved in automaticity. The SAN and
AVN are enriched with ANG II receptors (430). Habuchi
et al. (169) have shown that ANG II can slow pacemaker
activity in rabbit SAN cells by inhibiting I
Ca,L
. These au-
thors have reported that maximal I
Ca,L
inhibition by ANG
II is 30%. Slowing of pacemaker activity is characterized
by depolarization of the maximum diastolic potential and
a reduction of the action potential velocity. These effects
on the pacemaker cycle are consistent with inhibition by
ANG II of I
Ca,L
(169). Relaxin (RLX), a reproduction hor-
mone, accelerates heart rate in vivo in animal models.
Han et al. (181) have studied the effects of RLX in rabbit
SAN cells. In this study, RLX accelerated pacemaker ac-
tivity of isolated cells and dose-dependently enhanced
I
Ca,L
amplitude. The RLX signaling pathway is dependent
on cAMP and PKA, because cAMP analogs and PKI abol-
ished the effects of RLX (181).
Thyroid hormones (TH) are involved in many physi-
ological processes. The TH triiodothyronine (T
3
) has a
major role in long-term regulation of heart rate, cardiac
output, and tissue lipid content (246). Hyperthyroidism is
associated with tachycardia, arrhythmias, and elevated
cardiac output in animals and humans (28). In contrast,
hypothyroidism is associated with reduced heart rate and
cardiac output (28). THs act through nuclear hormone
receptors (TRs), which are ligand-dependent transcrip-
tion factors (28). Changes in the expression of cardiac ion
channels have been observed in hyperthyroid and hypo-
thyroid animals. Shimoni et al. and Guo et al. have shown
that I
to
density is robustly augmented in ventricles (455)
and neonatal cardiomyocytes (167) isolated from hyper-
thyroid rats. Renaudon et al. (410) have studied the ef-
fects of incubation of isolated rabbit SAN cells with T
3
.In
this study, T
3
significantly increased the density of I
f
in
the absence of an effect on its current-voltage depen-
dence. This observation suggests an augmentation of f-
channels expression in the cell membrane. Renaudon
et al. (410) have proposed that an increase in f-channel
expression underlies the increase in heart rate in hyper-
thyroid states. This hypothesis is partially supported by
three studies measuring the amount of mRNAs coding for
HCN2 channels in the heart of hyperthyroid and hypothy-
roid rats and mice (161, 270, 380). Indeed, Pachucki et al.
(380) have reported a fourfold increase in HCN2 mRNA
when T
3
was administrated to hypothyroid rats. Le Bouter
et al. (270) have found complex remodeling of ion channel
gene expression patterns in ventricles of hypothyroid
mice and reported a significant reduction in HCN2
mRNAs. In mammals, two genes encode for TRs, namely,
the TR
and TR
receptors with their isoforms (28). Mice
lacking the TR
isoform show 20% heart rate slowing
even after stimulation by T
3
. Gloss and co-workers (161,
524) have produced two distinct mouse lines lacking TR
and TR
.TR
/
mice have mild bradycardia and a re-
duction in HCN2 and HCN4 mRNAs in the heart. These
effects were mimicked by hypothyroidism. In contrast,
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 961
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
TR
/
mice were euthyroid and showed no significant
alterations in HCN4 and HCN2 mRNAs. Consequently,
Gloss et al. (161) proposed that TR
is the target for T
3
in
the heart and is responsible for the regulation of heart
automaticity by TH. However, the functional expression
of ion channels in SAN cells of TR
/
mice has not been
studied. It is possible that T
3
affects the expression of
proteins other than HCN in the SAN. For instance, down-
regulation of NCX and concomitant augmentation of Na
-
K
-ATPase in hypothyroid hearts has been reported
(303).
Parathyroid (PTH) hormones accelerate heart rate.
Hara et al. (189) have studied the effects of PTH on
pacemaker activity in rabbit SAN and canine Purkinje
fibers. PTH reversibly accelerated pacing rate in both
preparations and augmented the I
f
maximal conductance
by 68% (189). The effect of PTH was due to an increase in
the slope of diastolic depolarization. In isolated rabbit
SAN cells, Cs
inhibited the effects of PTH much better
than the VDCC blocker verapamil. Hara et al. (189) have
thus proposed that I
f
is the predominant mechanism by
which PTH increases heart rate.
D. Mechanical Load and Atrial Stretch
The cardiac cycle is associated with periodical filling
of the atrial chambers. Changes in the venous return thus
affect the diastolic atrial dimensions. An increase in the
right atrial filling stretches the atrial wall, as well as the
SAN. Atrial stretch due to venous return can have a direct
effect on pacemaker activity. Bainbridge (17) reported
that injection of fluids into the jugular vein of anesthetized
dogs elevated venous return and increased the heart rate.
This phenomenon is still referred to as the “Bainbridge
reflex.” Other investigators repeated Bainbridge’s experi-
ments and observed a positive or negative in vivo chro-
notropic response under various experimental conditions
(see Refs. 177, 251 for review). In vivo, this variability has
been accounted for by a dependence of the Bainbridge
reflex from the initial heart rate (95, 99).
Even if Bainbridge had attributed the positive chro-
notropic response to a reduced vagal tone upon increased
venous return, there is now substantial evidence indicat-
ing that at least part of the Bainbridge reflex is due to an
intrinsic control of SAN pacemaking by the mechanical
load of the atrium. Indeed, it is possible to reproduce a
positive chronotropic response to stretch in SAN tissue
(108), and on Purkinje fibers automaticity (233). These
results strongly suggest that the Bainbridge effect is, at
least in part, an intracardiac phenomenon. In this respect,
Wilson and Bolter have excluded that the Bainbridge
reflex can be due to a stretch effect on intracardiac neu-
rons (527). Furthermore, and consistent with early in vivo
evidence (95), Cooper and Kohl (99) have reported that
the direction of the chronotropic effect (positive or neg-
ative) induced by cell stretch depends on the initial cell
pacing rate. At the species level, the chronotropic effect
will be positive or negative depending on the mean resting
cellular rate of the species considered (99).
Cooper et al. (100) have shown that pacemaker ac-
tivity of SAN cells can be intrinsically mechano-modu-
lated via activation of SACs, because application of the
GsMTx-4 toxin, which blocks SACs, prevents the positive
chronotropic response to stretch in guinea pig SAN tissue
strips (99). The action of stretch on SAN cells can explain,
in part, the differences in pacing rate of isolated pace-
maker cells and in vivo heart rate, because electrophysi-
ological experiments are generally performed on mechan-
ically unloaded myocytes (251). Beside SACs, it is likely
that the mechanosensitivity of other ion channels can also
be involved in the chronotropic response to stretch of
SAN cells. For example, Lin et al. (287) have reported that
HCN channels are mechanosensitive. Regulation of pace-
maker activity by atrial filling can be relevant for the fine
beat-by-beat modulation of heart rate. Cooper et al. (99)
have highlighted that regulation of the diastolic phase by
mechanical load is useful for initiating a new heartbeat
when the venous return is elevated (99). Somewhat con-
flicting evidence exists as to the physiological relevance
of Bainbridge’s reflex in humans. Slovut et al. (463) have
proposed that stretch-induced modulation of SAN auto-
maticity can be responsible of the respiratory sinus ar-
rhythmia in cardiac allograft recipients (see also sect.
VIIA). This suggestion is based on the observation that in
23% of patients studied, the heart rate oscillated with
arterial pulse pressure (463). These authors have also
shown that intrinsic cardiac neurons are not involved in
this phenomenon so that HRV could be related to Bain-
bridge’s reflex. On the other hand, Casadei et al. (76) have
reported that, in normal subjects, the nonneuronal com-
ponent of HRV is negligible at rest and that this compo-
nent augments only in conditions of ganglionic block
during strong exercise. It is thus possible that, in humans,
Bainbridge’s reflex can become relevant under particular
physiological conditions (e.g., under reduced or blocked
autonomic tone), while under basal conditions, the auto-
nomic nervous system dominates the beat-by-beat regu-
lation of heart rate.
E. Electrolytes and Temperature
By affecting SAN pacemaker activity, the extracellu-
lar ionic environment (89, 447, 448) and the body temper-
ature can influence the heart rate.
In conditions of cardiac ischemia or intense exercise,
there is a significant augmentation of the extracellular K
962 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
and of the sympathetic tone (75). In the SAN, an increase
in the extracellular K
concentration shifts the leading
pacemaker site (292) (Fig. 6) and depolarizes the maxi-
mum diastolic potential (240, 248). It has been proposed
that increased extracellular K
will reduce the contribu-
tion of I
f
to the diastolic depolarization (240), but such an
effect would be difficult to demonstrate, because K
will
also enhance the I
f
Na
conductance (89). The negative
chronotropism associated with increased extracellular
K
may thus be attributed to a reduction of I
Kr
current
(240). Choate et al. (89) have reported that in intact SAN
preparations, the positive chronotropic effect observed
after stimulation of the sympathetic ganglion is reduced
when the extracellular K
is raised. These authors have
proposed that such a mechanism may protect the isch-
emic heart from deleterious effects of an excessive sym-
pathetic input (89). Ca
2
and Mg
2
also affect SAN chro-
notropism and can induce pacemaker shift (373, 378).
Ca
2
has a positive chronotropic effect, while Mg
2
in-
duces negative chronotropism (378). Mg
2
also interferes
with autonomic regulation of SAN pacemaking, possibly
by preventing pacemaker shift associated with norepi-
nephrine and ACh (378) (see Fig. 6). Acidosis of the
extracellular environment induces negative chronotro-
pism (399, 469). This mechanism can be responsible for
bradycardia during parturition (469) or under ischemic
conditions (164). Changes in heart rate following alter-
ations in the ionic concetrations of blood plasma have
been reported also in humans. For example, Severi and
co-workers (447, 448) have shown that even moderate
(within the physiological range) changes in the plasma
concentration of Ca
2
,K
, and H
induce significant
changes in the heart rate of human subjects, thus high-
lighting the importance of plasma electrolyte compositi-
tion in regulating cardiac automaticity.
The effects of temperature on pacemaking can be
roughly distinguished between short-term effects (e.g.,
transient changes in pacing rate) or long-term effects that
represent adaptation to low temperatures. In vivo, lower-
ing the temperature induces reduction of the SAN rhythm
and conduction time (247, 271). In the intact SAN, the
effects of temperature on pacemaking can also be asso-
ciated with pacemaker shift (271). Le Heuzey et al. (271)
reported that lowering the temperature in isolated rabbit
SAN preparations induces a caudal shift of the leading
pacemaker site. Hof et al. (202) have shown that a drop in
the temperature below the physiological range can also
interfere with the chronotropic effect of extracellular
Ca
2
, a phenomenon which seems also to be due to
pacemaker shift.
It has been recently demonstrated that pacemaking
can show long-term adaptation to low environmental tem-
peratures (193). Haverinen and Vornanen (193) reported
that rainbow trout bred at low water temperature have
increased heart rate. These authors have shown increased
I
Kr
density in isolated sinus venous cells of low-tempera-
ture bred fish. These results demonstrate that ionic cur-
rents involved in pacemaking can be modulated on a
long-term basis to adapt to environmental constraints.
SAN tissue seems also particularly resistant to long-
term exposure to low temperatures. Nishi et al. (357) have
shown that SAN automaticity can be recovered unscathed
after several days in cold Tyrode’s solution (5°C), while
the surrounding atrial tissue becomes completely unex-
citable. Similarly, Furuse et al. (147) reported that SAN
pacemaking is resistant to cold cardioplegia. Resistance
of SAN pacemaking to low temperatures has been ex-
ploited clinically for cardioprotection during heart sur-
gery, grafting and, more generally, for clinical hypother-
mia (259, 382, 409). Indeed, heart rate slowing induced by
low temperatures increases coronary blood flow, via an
improvement of diastolic atrial filling (10). Clinical hypo-
thermia can be very important in the treatment of ische-
mic heart disease, because elevated body temperature is
associated with worsening of ischemia-induced myocar-
dial necrosis (178).
Increased body temperature can directly affect the
heart rate (51, 98, 188, 239). Core temperature elevation
can significantly increase SAN rate during exercise (51).
Fever also induces a direct positive chronotropic re-
sponse (98, 188, 239). In a clinical study by Kiekkas et al.
(239), it was noted that increase of core temperature
during fever episodes was followed by a significant in-
crease in heart rate and a decrease in arterial blood
pressure. Alterations of heart rate and arterial blood pres-
sure were significantly affected by magnitude of fever
(239).
X. GENETIC AND ACQUIRED DISEASES OF
CARDIAC AUTOMATICITY
A. Inherited Dysfunction of SAN Automaticity
Propagation of the heartbeat requires coordination
between impulse generation and the spread of cardiac
excitation through the conduction system, which delivers
the impulse to the working myocardium. Genetic or ac-
quired dysfunction of cardiac ion channels can lead to
debilitating and life-threatening arrhythmias or heart
block. Arrhythmias are generally linked to desynchroni-
zation of the sequential activation of the heart, or to
defects in myocardial repolarization. Acquired arrhythmo-
genic diseases can be secondary to medical interventions,
including cardiac surgery and/or drug administration. Fur-
thermore, cardiac ischemia and heart failure can favor the
genesis of arrhythmias and sudden death (266, 315). Sev-
eral mutations in genes coding for cardiac ion channels
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 963
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
have been identified and shown to be linked to either
inherited dysfunction of the heartbeat or enhanced sus-
ceptibility to drug-induced arrhythmias (92, 236). During
the last few years, alterations in ion channels contributing
to the genesis and regulation of pacemaker activity have
been found, and some insight into possible physiopatho-
logical mechanisms underlying inherited dysfunction of
automaticity is now available.
SAN dysfunction (SND) accounts for half of the num-
ber of patients requiring implantation of an electronic
pacemaker device (266). SND is characterized by a com-
bination of symptoms including fatigue and syncope. Bra-
dycardia, SAN arrest, or exit block are typical features of
SND (315). In some cases, alternating periods of brady-
cardia and tachyarrhythmias can be observed in SND
patients (36). In a substantial number of cases, SND is
associated with acquired cardiac conditions, such as
heart failure, ischemia, cardiomyopathy, or administra-
tion of antiarrhythmic drugs. However, in a significant
percentage of patients, SND is unrelated to structural
abnormalities of the heart, but shows familial legacy (274,
301, 436).
To date, three mutations underlying SND have been
identified in the human HCN4 gene (hHCN4) (330, 443,
486). In a recent case report, Schultze-Bahr et al. (443)
have identified a hHCN4 mutation that caused sinus bra-
dycardia with intermittent atrial fibrillation and loss of
exercise-dependent increase in heart rate in a patient.
This mutation named (hHCN4–573X) generates a trunca-
tion at the protein COOH terminus upstream to the chan-
nel cAMP-binding domain (CNBD). Recombinant hHCN4
and hHCN4 –573X channels have similar voltage depen-
dence. However, hHCN4 –573X channels are insensitive to
cAMP. Coexpression of hHCN4–573X with normal recom-
binant hHCN4 channels indicates that mutant channels
have a dominant-negative effect so that coexpressed f-
channels cannot be facilitated by cAMP. Insensitivity of
hHCN4–573X channels to cAMP can explain SAN brady-
cardia and autonomic incompetence in patients carrying
the hHCN4–573X mutation (443). It is possible that atrial
fibrillation could be secondary to bradycardia. Ueda et al.
(486) have searched for mutations in ion channel genes in
a series of 25 unrelated patients having SND, conduction
disturbances, and idiopathic ventricular fibrillation. A
missense mutation in hHCN4 (hHCN4-D553N) has been
identified in this survey. This mutation is located in the
linker region between the transmembrane core and the
intracellular COOH terminus of hHCN4. The D553N mu-
tation seems to affect cellular trafficking of hHCN4 chan-
nels. In coexpression experiments, hHCN4-D553N chan-
nels reduced I
f
current mediated by normal HCN4 chan-
nels, suggesting a dominant negative suppression of
HCN4. In a patient carrying the hHCN4-D553N mutation,
severe bradycardia with associated prolongation of the
Q-T interval was present. Also, a long sinus pause, fol-
lowed by polymorphic ventricular tachycardia, has been
recorded (486).
A mutation named hHCN4-S672R has been recently
identified by Milanesi et al. (330) in members of a large
Italian family. Compared with the other two mutations
discussed above, hHCN4-S672R is associated with con-
genital asymptomatic bradycardia and is inherited as an
autosomal dominant allele. Affected individuals have mild
bradycardia (mean heart rate is 52 beats/min) without
other associated arrhythmias. The hHCN4-S672R maps in
the CNBD, yet affected f-channels are normally respon-
sive to cAMP. However, coexpressing normal hHCN4
with hHCN4-S672R channels yields currents that activate
at voltages 10 mV more negative than wild-type hHCN4-
mediated currents (330). We can infer that such a negative
shift in the I
f
activation curve results in diminished I
f
-
mediated inward current in the diastolic depolarization
range, leading to slowed basal heart rate. Milanesi et al.
(330) have emphasized the similarity between the moder-
ate reduction of heart rate in patients carrying hHCN4-
S672R mutation and the effect of low ACh doses (10–30
nM) on pacemaking of isolated rabbit SAN cells. Another
hHCN4 familial mutation, namely, G480R, has been re-
cently identified by Nof et al. (361). The hHCN4-G480R
maps in the pore domain. Affected f-channels activate at
more negative voltages than wild-type counterparts and
have defective regulation of intracellular trafficking. Indi-
viduals carrying hHCN4-G480R have asymptomatic bra-
dycardia (361).
SAN bradycardia has also been shown in association
with congenital heart block (CHB) (209). CHB is charac-
terized by progressive complete atrioventricular block
affecting fetuses and newborns (see Ref. 55 for review).
CHB is generally detected just before or immediately after
birth (512). CHB is due to production of autoantibodies
against intracellular soluble ribonucleoproteins named 48
kDa SSB/La, 52 kDa SSA/Ro, and 60 kDa SSA/Ro (512). Hu
et al. (209) have reported inhibition of I
Ca,L
and I
Ca,T
by
IgGs isolated from mothers having CHB-affected children.
It is thus tempting to speculate that downregulation of
Ca
v
1.3 and Ca
v
3.1 channels by maternal antibodies under-
lies SAN bradycardia in CHB (209). This hypothesis has
been recently supported by Qu et al. (406), who showed
that the Ca
v
1.3 channel protein is expressed in the human
fetal heart and that anti-Ro/La antibodies can effectively
inhibit Ca
v
1.3-mediated I
Ca,L
expressed in tsA201 cells.
Mutations in Na
v
1.5 channels can also cause SND.
Benson et al. (33) have recently described two mutations
in the Na
v
1.5 gene leading to recessive disorders of car-
diac conduction, characterized by bradycardia progress-
ing to atrial unexcitability during infancy. It is unclear
whether bradycardia is due to dysfunction of the impulse
generation in the SAN or to partial block of intranodal
conduction.
964
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
B. Automaticity in Heart Failure and
Cardiac Ischemia
Heart failure (HF) is a major risk factor for life-
threatening ventricular arrhythmias and sudden death. HF
carries a poor prognosis; mortality rates at 5 years for
untreated or poorly treated HF are close to 60% (231).
ECG recordings in HF patients show decreased mean
heart rate and heart rate variability (223, 375). Remodel-
ing of intrinsic SAN function and regulation by the auto-
nomic nervous system has been observed in HF. For
example, Sanders et al. (434) have mapped SAN activation
and conduction reserve in 18 HF patients. They have
shown complex rearrangements of SAN excitability, in-
cluding alteration of intranodal conduction, prolongation
of the SAN recovery time, and a caudal shift of the leading
pacemaker site. Experimental animal models of HF fairly
reproduce clinical observations (see the recent review by
Janse, Ref. 221). Opthof et al. (376) have investigated SAN
function in a rabbit model of HF associated with sudden
death. Isolated right atrial preparations of HF rabbits
(376) show a decrease in the intrinsic SAN rate and an
increase in the sensitivity to ACh. However, in spite of
such an enhancement of negative chronotropic factors,
sudden death is still observed in these animals at accel-
erated heart rates. Opthof et al. (376) have suggested that
the reduction of intrinsic SAN rate constitutes an adapta-
tion mechanism to counteract an augmentation in sympa-
thetic tone and its arrhythmogenic potential, and to opti-
mize myocardial oxygen consumption. Verkerk et al.
(503) have reported that isolated SAN cells from HF rab-
bits have an intrinsically longer cycle length and show a
40% reduction of I
f
(and 20% of I
Ks
) compared with con-
trol cells. The slope of the diastolic depolarization ap-
pears to be specifically affected by HF. Verkerk et al.
(503) have thus proposed that the intrinsically slower
SAN rate observed in HF animal models (and possibly in
humans) is due to I
f
downregulation. Consistent with
these findings, a strong reduction in HCN4 and HCN2
mRNA and protein in the SAN has been reported by Zicha
et al. (548) using a canine model of congestive HF. Dis-
ease seems also to have differential effects on the SAN
and the atria, since HF induced an increase in HCN4
expression in atrial tissue. It is possible that HCN upregu-
lation can contribute to the development of supraventric-
ular arrhythmias associated with HF (548).
Cardiac ischemia can induce SND. Bradycardia is
commonly observed upon resuscitation after cardiac ar-
rest or ventricular fibrillation (379). Severe bradycardia
following cardiac ischemia is a major cause of poor sur-
vival after cardiac arrest (156, 379). SAN artery stenosis or
occlusion is another factor leading to bradycardia in pa-
tients (5) and dogs (85). Heart rate slowing can be ob-
served in Langendorff-perfused hearts upon ischemia-
reperfusion (337). Considerable interest exists therefore
in elucidating the mechanisms of ischemia-induced
bradycardia for managing cardiac ischemic disease and
cardiac arrest. The cellular basis of ischemia-dependent
bradycardia is not entirely understood. Slowing of dia-
stolic depolarization, reduction in action potential ampli-
tude, and depolarization of the maximum diastolic poten-
tial have been consistently observed in isolated rabbit
atria during hypoxia (446) or metabolic inhibition (253).
Similar changes in the pacemaker cycle parameters have
been recently reported in isolated rabbit SAN cells by
Gryshenko et al. (164), using a low pH (6.6) “ischemic”
Tyrode’s solution to mimic acidosis of the extracellular
environment in ischemic conditions (see Ref. 75 for re-
view). These effects were reversible when switching back
to normal Tyrode’s solution. Metabolic inhibition of rabbit
SAN cells attenuates I
Ca,L
, I
Kr
, and I
f
(180) and activates
I
K,ATP
(184). In contrast, Gryshenko et al. (164) have
observed an augmentation of the instantaneous net in-
ward current upon hyperpolarization, an effect attributed
to acidosis-dependent reduction of I
KACh
(even if the in-
duction of an inward background current cannot be ex-
cluded). Furthermore, reduction in I
Ca,T
and I
NCX
has
been reported by Du and Nathan (133). It is a surprising
finding of this work that I
Ca,L
was augmented by ischemic
Tyrode’s solution. However, as suggested by Du and
Nathan (133), such an augmentation may not lead to a
positive chronotropic effect, since the more positive max-
imum diastolic potential, induced by ischemic Tyrode’s
solution, can accentuate I
Ca,L
inactivation. From the
above results, it is difficult to attribute ischemia-induced
slowing of cardiac automaticity to a single ionic mecha-
nism. Furthermore, it is possible that a given ionic current
can be more or less affected depending on the experimen-
tal conditions (e.g., acidosis or direct metabolic inhibi-
tion). Future investigations will further elucidate the cel-
lular mechanisms of ischemia-induced bradycardia and
SND.
XI. HEART AUTOMATICITY AND
CARDIOPROTECTION
In the preceding sections, we have discussed the
mechanisms underlying automaticity and its regulation by
the autonomic nervous system, hormones, and different
physiopathological conditions including cardiac ischemia
and HF. We will now review recent pharmacological and
gene therapy approaches aiming to control pacemaker
activity and heart rate in cardiac disease. These new
therapeutic strategies are currently focused on f-channels
and reduction in I
K1
. Indeed, the capability of I
f
to mod-
ulate the slope of the diastolic depolarization renders
f-channels suitable targets for pharmacological regulation
GENESIS AND REGULATION OF THE HEART AUTOMATICITY 965
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
of heart rate and for creating “biological pacemakers” in
the diseased myocardium.
A. Heart Rate and Cardiac Morbidity
The mean resting heart rate is correlated with all-
cause cardiovascular mortality and morbidity. The Fra-
mingham epidemiological study has highlighted the asso-
ciation between elevated heart rate and all-cause mortal-
ity (134, 160, 232). The mechanistic link between heart
rate and mortality is still unclear. Perski et al. (389)
presented evidence supporting the hypothesis that ele-
vated heart rate contributes to the progression of athero-
sclerosis in postinfarct patients. In subjects with ischemic
cardiac disease such as stable angina pectoris, elevated
heart rate can cause an imbalance between myocardial
oxygen supply and the actual physiological demand.
Heart rate is inversely correlated with the degree of filling
of the left ventricle with oxygen-rich blood (10). A better
irrigation of the coronary system is achieved by a reduc-
tion of the heart rate (10). Andrews et al. (9) have con-
firmed this prediction at the clinical level by showing that
lowering heart rate with propranolol effectively prevents
cardiac ischemic episodes. Development of therapeuti-
cally active molecules, able to reduce heart rate, is thus of
outstanding interest. Targeting of ion channels specifi-
cally involved in the regulation of diastolic depolarization
rate constitutes a natural approach toward this goal.
-Blockers (201) and I
f
inhibitors (123) are effective in
reducing ischemic episodes and cardiac mortality by vir-
tue of their heart rate-reducing action. Specific reduction
of heart rate is now considered as a new therapeutically
effective approach to manage cardiac ischemia (123, 201).
Among I
f
inhibitors, the only molecule that has been
clinically developed for treatment of stable angina pecto-
ris is ivabradine (123). In animal models, ivabradine limits
exercise-induced tachycardia and improves the balance
between myocardial oxygen supply and demand (97). To
date, no inotropic (461) or dromotropic effects (96) have
been reported for ivabradine. A phase II clinical study,
conducted on patients with stable angina, has confirmed
specific heart rate reduction at rest and during exercise
and shown the efficacy of ivabradine as an anti-ischemic
and antianginal agent (52). The in vivo pharmacological
properties of ivabradine are consistent with a specific
action on I
f
in the SAN. Compared with
-blockers, iv-
abradine seems more effective in preserving myocardial
recovery and performance after cardiac stunning (52).
The f-channel is the first ion channel underlying pacemak-
ing to be targeted by a therapeutically active drug. It is
possible that other channel classes, such as Ca
v
1.3 and
Ca
v
3.1 channels, will constitute future pharmacological
targets for the development of new molecules to regulate
heart rate without negative inotropism.
B. Automaticity in Engineered “Biological
Pacemakers”
1. Direct gene transfer of “pacemaker” channels in
cardiac cells
Identification of ion channels involved in cardiac au-
tomaticity harbors the attractive perspective of engineer-
ing “biological” pacemakers to stimulate automaticity in
defined regions of the heart (Fig. 27). Biological pacemak-
ers have been proposed as a possible alternative to elec-
tronic devices (424), even if a biological pacemaker can
also work in combination with an electronic pacemaker
(421). Several up-to-date reviews have been recently pub-
lished on this topic (421, 423, 424). Here, we focus on
some fundamental aspects of biological pacemakers. The
first gene therapy approach to increase pacemaker activ-
ity in the working myocardium came from experiments in
which a cDNA encoding for the
2
-adrenergic receptor
was injected in the right atrium of mouse (137) and pig
(138). Overexpression of
2
-receptors increased the basal
heart rate in both animal models (137, 138).
A gene therapy approach was employed by Miake
et al. (327) to generate pacemaker activity in the ven-
tricular myocardium of guinea pigs. In this work, ad-
enoviral gene transfer of a dominant negative Kir2.1
subunit (Kir2.1AAA) has been employed (327). This con-
struct downregulated I
K1
by 80% in 20% of ventricular
myocytes tested (328). Depolarization of the resting po-
tential to 60 mV generated a diastolic depolarization
phase in ventricular myocytes and triggered premature
ventricular beats in vivo (327). On the basis of these
results, Miake et al. (327) have suggested that the working
myocardium is also endowed with automaticity, but pace-
making in ventricular cells is naturally repressed by I
K1
expression. This proposal challenges the view that auto-
maticity depends on the expression of particular “pace-
maker” ion channel. Silva and Rudy (459) and Kurata et al.
(263) have studied the possible mechanisms of automa-
ticity in ventricular cells expressing Kir2.1AAA by em-
ploying modified versions of the Luo and Rudy model of
ventricular action potential (298). Numerical simulations
suggest that automaticity in these cells can be generated
by I
NCX
(459), or by I
Ca,L
and I
NCX
(263). Particularly,
CICR triggered by I
Ca,L
during the action potential would
stimulate I
NCX
, which depolarizes the cell membrane to
begin a new cycle (459), while I
Ca,L
can be responsible for
a membrane voltage instability which promotes automa-
ticity (263). Silva and Rudy have proposed also that
-ad-
renergic stimulation of the ventricular I
NCX
-driven pace-
maker mechanism could not exceed 25%. Consequently,
the automaticity in Kir2.1AAA-expressing cells probably
lacks the sensitivity to autonomic regulation needed to
sustain everyday life. Viswanathan et al. (511) have fur-
ther assessed, by numerical modeling, the problem of
966
MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
automaticity in directly transfected ventricular cells.
These authors have suggested that to obtain stable auto-
maticity, both downregulation of I
K1
and expression of
HCN channels are required (511).
Overexpression of HCN channels by gene transfer
constitutes another strategy for creating biological pace-
makers (421). This approach is based on the experimental
observation that expression of HCN2 or HCN4 channels in
cultured neonatal ventricular myocytes robustly improves
cellular automaticity (143, 403). Expression of a dominant
negative HCN2 subunit abolished automaticity (143).
These observations indicate that HCN channels are able
to confer automaticity to myocardial cells upon gene
transfer. Qu et al. (405) took advantage of this new con-
cept and employed gene transfer of HCN2 channels
in vivo in the canine left atrium. Qu et al. (405) have
shown that I
f
expression in atrial myocytes generated
viable atrial rhythms upon inhibition of SAN activity.
Plotnikov et al. (394) have employed gene transfer of
HCN2 channels to stimulate automaticity in the left bun-
dle branch of the Purkinje fiber network (Fig. 27). In this
work, adenoviral gene transfer of HCN2 channels im-
proved the rate of ventricular rhythms observed after
inhibition of SAN activity by transient vagal stimulation.
To evaluate the sensitivity of these ventricular rhythms to
autonomic agonists, radiofrequency ablation of the AVN
was performed to dissociate the atrial rhythm from that
originating from the left bundle branch (421). In these
conditions, ventricular rhythms were significantly respon-
sive to epinephrine (421).
Bucchi et al. (69) have employed gene transfer of a
mutant HCN2 channel having a positively shifted activa-
tion curve midpoint (46 versus 66 mV for wild-type
HCN2 channels). Compared with rhythms generated by
wild-type HCN2 channels, expression of mutant HCN2
channel significantly improved the responsiveness of ven-
tricular rhythms to catecholamines. Expression of HCN4
channels in the ventricle also seems to create
-adrener-
gic modulation of ventricular escape rhythms (71). Mod-
ification of HCN voltage dependency to favor channel
opening at positive voltages has been employed also by
Tse et al. (483). These authors have shown that focal
FIG. 27. The primary SAN pacemaker cell is the natural pacemaker (top panel). In SAN cells, high expression of native HCN channels (black
channels) contributes to SAN pacemaker dominance and to the autonomic regulation of heart rate. The SAN pacemaker cycle (top, left) drives the
right atrium at a physiological suitable rate. In case of SAN failure or AVN block, the intrinsic pacing rate of the Purkinje fibers network could be
enhanced by gene transfer of recombinant HCN channels (gray channels) in the bundle branches. The ideal Purkinje cell (middle panel) will thus
drive the ventricular cell at a faster frequency than under normal conditions, thereby generating a physiologically suitable ventricular basal rate. The
sensitivity of HCN channels to cAMP is also potentially important for creating a ventricular pacemaker that can be regulated by the autonomic
nervous system. A stem cell-derived biological pacemaker (bottom panel) is based on overexpression of recombinant HCN channels in unexcitable
stem cells implanted in a target region of the working myocardium. Stem cells connect to ventricular cells by forming gap junctions with ventricular
cells. The positive membrane potential of stem cells will cyclically drive ventricular cells to the action potential threshold. [Adapted from Robinson
et al. (421).]
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
967
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
expression of a mutant form of HCN1 in the left atrium
reduces the dependence on electronic pacemakers of
SAN-ablated pigs (483).
2. Stem-cell based gene transfer
An alternative approach for creating biological pace-
makers consists in overexpressing HCN channels in un-
excitable human mesenchymal stem cells (hMSCs). These
cells have the natural capability of coupling to cardiac
myocytes via gap junctions (400). When unexcitable hM-
SCs connect to myocytes, their membrane potential will
become more negative due to the electrotonic influence of
myocytes. Membrane hyperpolarization will activate HCN
channels, thereby generating inward current that is fed
back to the myocyte to trigger an action potential [Fig. 27,
see Robinson et al. (421) for further discussion]. Action
potential repolarization will then tend to shut off HCN
channels in the hMSC before the beginning of a new cycle.
Potapova et al. (400) have reported that HCN2-mediated
currents expressed in hMSCs are sensitive to isoprotere-
nol, indicating that the signaling pathway necessary for
HCN channel regulation is present in hMSCs. For a more
detailed discussion on the clinical and experimental prob-
lems associated with stem cell transfection and implan-
tation in the heart, the reader is referred to the review by
Rosen et al. (423). Downregulation of I
K1
in ventricular
cells could also be employed to achieve robust stem cell-
mediated pacemaking (230). Gene transfer of stem cells
may not constitute an exclusive approach to create bio-
logical pacemakers based on cell therapy. Indeed, it is
possible that also fibroblasts can be genetically modified
to express ion channels (144). Fibroblasts are spontane-
ously capable of connecting with cardiomyocytes in vitro
and can affect conduction between myocytes (144). In
cell culture, Kizana et al. (243) have obtained functional
electrical coupling between myotubes differentiated from
MyoD-Cx43 coexpressing fibroblasts. Interestingly, these
authors have reported that coexpression of Cx43 with
MyoD was critical for observing electrical coupling and
uniform threshold for excitation between adjacent myo-
tubes (243). Cx43 expression also enhances electrical
coupling between neonatal cardiomyocytes and fibro-
blasts in cell culture (244). Expression and modulation of
appropriate Cxs together with ion channels can thus be an
important step for overcoming the problem of poor elec-
trical coupling between stem cells and myocytes in situ.
This potential approach is supported by a numerical mod-
eling study by Jacquemet (219), who emphasized that
fibroblasts can significantly affect myocardial conduction
only if the degree of coupling between fibroblast and
myocytes is sufficiently high.
In conclusion, establishment of biological pacemak-
ers in defined regions of the heart is a fascinating per-
spective for the cellular therapy of the heartbeat. A bio-
logical pacemaker may also necessitate the coexpression
or modification of more than one type of ion channel to
reach a suitable degree of intrinsic automaticity and au-
tonomic regulation of pacemaking (263, 511). The obser-
vation that the Tbx3 transcription factor can reprogram
nonpacing atrial cells to a phenotype similar to that of
automatic cells may also open new perspectives into the
creation of biological pacemakers (208).
XII. CONCLUDING REMARKS
Our understanding of cardiac automaticity has pro-
gressed considerably during the last 10 years. Indeed,
at the beginning of the 1990s, the study of the genera-
tion of pacemaker activity was somewhat limited to the
electrophysiological description of ion currents in
spontaneously active cells. Molecular cloning of gene
families coding for ion channels has since allowed in-
vestigation of gene expression in cardiac tissue and the
establishment of genetically engineered mouse lines
that show specific dysfunctions of cardiac automatic-
ity. These lines are now yielding precious insights into
the generation and regulation of pacemaking and will
probably constitute new animal models of heart rhythm
pathologies. We can expect that the molecular basis of
many inherited diseases of cardiac automaticity will be
elucidated in the near future and that different ion
channels and proteins regulating Ca
2
handling will be
involved in such diseases.
Several questions on how the heart rhythm is gener-
ated and controlled in physiological and pathological con-
ditions remain unanswered. It is not completely under-
stood, for example, which ion channels are essential for
generating the diastolic depolarization in the SAN, AVN,
and the Purkinje fibers network, and which mechanisms
play a dominant role in the autonomic regulation of au-
tomaticity in humans. The interactions between the mem-
brane “ion channels clock” and the intracellular “Ca
2
clock” are not completely elucidated. Nevertheless, these
interactions are of fundamental importance for under-
standing the integration of pacemaker mechanisms at the
cellular level.
From the above discussion, it also appears that car-
diac automaticity at the organ level is a very complex
phenomenon and that, beside cellular mechanisms, inte-
grative factors are involved in cardiac pacemaking. These
factors include the heterogeneity of SAN and AVN tissue
organization and the regulation of heart rate by mechan-
ical load and by several humoral, environmental, and
pathophysiological states.
Finally, ion channels involved in automaticity, such
as f-channels, are now becoming potential new drug tar-
gets. In this respect, the development of specific heart
rate-reducing agents underlines the importance of pursu-
968 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
ing the effort of obtaining further insight into the cardiac
pacemaker mechanisms. The engineering of “biological
pacemakers” is currently based on expression and manip-
ulation of K
ir
and f-channels in native myocardium or in
stem cells. Other complementary approaches can be the
use of fibroblasts and, possibly, manipulation of Cxs. It
will also be interesting to see if future approaches will
also include coexpression of other ion channels involved
in “native” automaticity.
NOTE ADDED IN PROOF
After acceptance of this manuscript, Li et al. (284a)
have proposed a detailed anatomical model of the rabbit
AVN. Numerical simulations of impulse conduction
through AVN are consistent with the hypothesis that au-
tomaticity originates in the PNE of the AVN.
Harzeim et al. (190a) have shown that abolition of
cAMP sensitivity of HCN4 channels by knock-in of a point
mutation in the channel CNBD prevents
-adrenergic
stimulation of heart rate in mouse embryos and lethality
as in HCN4 knockout embryos.
ACKNOWLEDGMENTS
We thank Patrick Atger for excellent technical skills in creat-
ing and editing the manuscript figures. We are indebted to Elodie
Kupfer and Anne Cohen-Solal for breeding and managing mouse
lines in our group. We thank Peter Kohl and Etienne Verheijck for
helpful discussion. We also thank Halina Dobrzynski, Mark Boyett,
and Dario DiFrancesco for sharing figure originals.
Address for reprint request and other correspondence:
M. E. Mangoni, Institute of Functional Genomics, Dept. of Phys-
iology CNRS UMR5203, INSERM, U661, Univ. of Montpellier I,
Univ. of Montpellier II, Montpellier, F-34094 France (e-mails:
matteo.mangoni@igf.cnrs.fr; joel.nargeot@igf.cnrs.fr).
GRANTS
Work in our laboratory has been supported by grants from
the CNRS, the Action Concerte´ e Incitative in Physiology and
Developmental Biology of the French Ministry of Education, the
INSERM National Program for Cardiovascular Diseases, the
Fondation de France and the Agence Nationale pour la Recher-
che (ANR). We also thank the International Research Institute
Sevier (IRIS) for having supported research activity in our
group.
REFERENCES
1. 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 embry-
onic stem cell-derived cardiomyocytes. J Physiol 523: 377–389,
2000.
2. Accili EA, Proenza C, Baruscotti M, DiFrancesco D. From
funny current to HCN channels: 20 years of excitation. News
Physiol Sci 17: 32–37, 2002.
3. Accili EA, Redaelli G, DiFrancesco D. Activation of the hyper-
polarization-activated current (i
f
) in sino-atrial node myocytes of
the rabbit by vasoactive intestinal peptide. Pflu¨ gers Arch 431: 803–
805, 1996.
4. Accili EA, Robinson RB, DiFrancesco D. Properties and modu-
lation of I
f
in newborn versus adult cardiac SA node. Am J Physiol
Heart Circ Physiol 272: H1549–H1552, 1997.
5. Alboni P, Baggioni GF, Scarfo S, Cappato R, Percoco GF,
Paparella N, Antonioli GE. Role of sinus node artery disease in
sick sinus syndrome in inferior wall acute myocardial infarction.
Am J Cardiol 67: 1180–1184, 1991.
6. Alings AM, Abbas RF, Bouman LN. Age-related changes in struc-
ture and relative collagen content of the human and feline sino-
atrial node. A comparative study. Eur Heart J 16: 1655–1667, 1995.
7. Altomare C, Terragni B, Brioschi C, Milanesi R, Pagliuca C,
Viscomi C, Moroni A, Baruscotti M, DiFrancesco D. Hetero-
meric HCN1-HCN4 channels: a comparison with native pacemaker
channels from the rabbit sinoatrial node. J Physiol 549: 347–359,
2003.
8. An WF, Bowlby MR, Betty M, Cao J, Ling HP, Mendoza G,
Hinson JW, Mattsson KI, Strassle BW, Trimmer JS, Rhodes
KJ. Modulation of A-type potassium channels by a family of cal-
cium sensors. Nature 403: 553–556, 2000.
9. Andrews TC, Fenton T, Toyosaki N, Glasser SP, Young PM,
MacCallum G, Gibson RS, Shook TL, Stone PH. Subsets of
ambulatory myocardial ischemia based on heart rate activity. Cir-
cadian distribution and response to anti-ischemic medication The
Angina and Silent Ischemia Study Group (ASIS). Circulation 88:
92–100, 1993.
10. Anrep GV, Hausler H. The coronary circulation. II. The effect of
changes of temperature and of heart rate. J Physiol 67: 299 –314,
1929.
11. Anumonwo JM, Delmar M, Jalife J. Electrophysiology of single
heart cells from the rabbit tricuspid valve. J Physiol 425: 145–167,
1990.
12. Anumonwo JM, Freeman LC, Kwok WM, Kass RS. Delayed
rectification in single cells isolated from guinea pig sinoatrial node.
Am J Physiol Heart Circ Physiol 262: H921–H925, 1992.
13. Anumonwo JM, Tallini YN, Vetter FJ, Jalife J. Action potential
characteristics and arrhythmogenic properties of the cardiac con-
duction system of the murine heart. Circ Res 89: 329–335, 2001.
14. Arai A, Kodama I, Toyama J. Roles of Cl
channels and Ca
2
mobilization in stretch-induced increase of SA node pacemaker
activity. Am J Physiol Heart Circ Physiol 270: H1726–H1735, 1996.
15. Ardell JL. Structure and function of mammalian intrinsic cardiac
neurons. In: Neurocardiology, edited by Armour JA, Ardell JL. New
York: Oxford Univ. Press, 1994, p. 95–114.
16. Bahinski A, Nairn AC, Greengard P, Gadsby DC. Chloride
conductance regulated by cyclic AMP-dependent protein kinase in
cardiac myocytes. Nature 340: 718–721, 1989.
17. Bainbridge FA. The influence of venous filling upon the rate of the
heart. J Physiol 50: 65–84, 1915.
18. Baker K, Warren KS, Yellen G, Fishman MC. Defective “pace-
maker” current (I
h
) in a zebrafish mutant with a slow heart rate.
Proc Natl Acad Sci USA 94: 4554 4559, 1997.
19. Barbuti A, Gravante B, Riolfo M, Milanesi R, Terragni B,
DiFrancesco D. Localization of pacemaker channels in lipid rafts
regulates channel kinetics. Circ Res 94: 1325–1331, 2004.
20. Barbuti A, Terragni B, Brioschi C, DiFrancesco D. Localization
of f-channels to caveolae mediates specific beta2-adrenergic recep-
tor modulation of rate in sinoatrial myocytes. J Mol Cell Cardiol 42:
71–78, 2007.
21. Baruscotti M, Bucchi A, Difrancesco D. Physiology and phar-
macology of the cardiac pacemaker (“funny”) current. Pharmacol
Ther 107: 59–79, 2005.
22. Baruscotti M, Difrancesco D. Pacemaker channels. Ann NY
Acad Sci 1015: 111–121, 2004.
23. Baruscotti M, DiFrancesco D, Robinson RB. Na
current con-
tribution to the diastolic depolarization in newborn rabbit SA node
cells. Am J Physiol Heart Circ Physiol 279: H2303–H2309, 2000.
24. Baruscotti M, DiFrancesco D, Robinson RB. A TTX-sensitive
inward sodium current contributes to spontaneous activity in new-
born rabbit sino-atrial node cells. J Physiol 492: 21–30, 1996.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
969
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
25. Baruscotti M, Westenbroek R, Catterall WA, DiFrancesco D,
Robinson RB. The newborn rabbit sino-atrial node expresses a
neuronal type I-like Na
channel. J Physiol 498: 641–648, 1997.
26. Batulevicius D, Pauziene N, Pauza DH. Topographic morphol-
ogy and age-related analysis of the neuronal number of the rat
intracardiac nerve plexus. Ann Anat 185: 449 459, 2003.
27. Baumgarten CM, Clemo HF. Swelling-activated chloride chan-
nels in cardiac physiology and pathophysiology. Prog Biophys Mol
Biol 82: 25–42, 2003.
28. Baxter JD, Dillmann WH, West BL, Huber R, Furlow JD,
Fletterick RJ, Webb P, Apriletti JW, Scanlan TS. Selective
modulation of thyroid hormone receptor action. J Steroid Biochem
Mol Biol 76: 31–42, 2001.
29. Beau SL, Hand DE, Schuessler RB, Bromberg BI, Kwon B,
Boineau JP, Saffitz JE. Relative densities of muscarinic cholin-
ergic and beta-adrenergic receptors in the canine sinoatrial node
and their relation to sites of pacemaker activity. Circ Res 77:
957–963, 1995.
30. Belardinelli L, Giles WR, West A. Ionic mechanisms of adeno-
sine actions in pacemaker cells from rabbit heart. J Physiol 405:
615–633, 1988.
31. Belardinelli L, Shryock J, West GA, Clemo HF, DiMarco JP,
Berne RM. Effects of adenosine and adenine nucleotides on the
atrioventricular node of isolated guinea pig hearts. Circulation 70:
1083–1091, 1984.
32. Belhassen B, Pelleg A, Shoshani D, Geva B, Laniado S. Elec-
trophysiologic effects of adenosine-5-triphosphate on atrioventric-
ular reentrant tachycardia. Circulation 68: 827–833, 1983.
33. Benson DW, Wang DW, Dyment M, Knilans TK, Fish FA,
Strieper MJ, Rhodes TH, George AL Jr. Congenital sick sinus
syndrome caused by recessive mutations in the cardiac sodium
channel gene (SCN5A). J Clin Invest 112: 1019–1028, 2003.
34. Bernardi L, Salvucci F, Suardi R, Solda PL, Calciati A, Perlini
S, Falcone C, Ricciardi L. Evidence for an intrinsic mechanism
regulating heart rate variability in the transplanted and the intact
heart during submaximal dynamic exercise? Cardiovasc Res 24:
969–981, 1990.
35. Bers DM. Cardiac excitation-contraction coupling. Nature 415:
198–205, 2002.
36. Bertram H, Paul T, Beyer F, Kallfelz HC. Familial idiopathic
atrial fibrillation with bradyarrhythmia. Eur J Pediatr 155: 7–10,
1996.
37. Bescond J, Bois P, Petit-Jacques J, Lenfant J. Characterization
of an angiotensin-II-activated chloride current in rabbit sino-atrial
cells. J Membr Biol 140: 153–161, 1994.
37a.Bharati S. Anatomic-morphologic relations between AV nodal
structure and function in the normal and diseased heart. In:
Atrial-AV Nodal Electrophysiology: A View From the Millenium,
edited by Mazgalev TN, Tchou PJ. Armonki Futura, 2000, p. 25– 48.
38. Billette J. Atrioventricular nodal activation during periodic pre-
mature stimulation of the atrium. Am J Physiol Heart Circ Physiol
252: H163–H177, 1987.
39. Billette J. What is the atrioventricular node? Some clues in sorting
out its structure-function relationship. J Cardiovasc Electrophysiol
13: 515–518, 2002.
40. Blaschke RJ, Hahurij ND, Kuijper S, Just S, Wisse LJ, Deissler
K, Maxelon T, Anastassiadis K, Spitzer J, Hardt SE, Scholer H,
Feitsma H, Rottbauer W, Blum M, Meijlink F, Rappold G, Git-
tenberger-de Groot AC. Targeted mutation reveals essential func-
tions of the homeodomain transcription factor Shox2 in sinoatrial and
pacemaking development. Circulation 115: 1830 –1838, 2007.
41. Bleeker WK, Mackaay AJ, Masson-Pevet M, Bouman LN,
Becker AE. Functional and morphological organization of the
rabbit sinus node. Circ Res 46: 11–22, 1980.
42. Bode F, Sachs F, Franz MR. Tarantula peptide inhibits atrial
fibrillation. Nature 409: 35–36, 2001.
43. Bogdanov KY, Maltsev VA, Vinogradova TM, Lyashkov AE,
Spurgeon HA, Stern MD, Lakatta EG. Membrane potential fluc-
tuations resulting from submembrane Ca
2
releases in rabbit sino-
atrial nodal cells impart an exponential phase to the late diastolic
depolarization that controls their chronotropic state. Circ Res 99:
979–987, 2006.
44. Bogdanov KY, Vinogradova TM, Lakatta EG. Sinoatrial nodal
cell ryanodine receptor and Na()-Ca(2) exchanger: molecular
partners in pacemaker regulation. Circ Res 88: 1254–1258, 2001.
45. Bohn G, Moosmang S, Conrad H, Ludwig A, Hofmann F,
Klugbauer N. Expression of T- and L-type calcium channel mRNA
in murine sinoatrial node. FEBS Lett 481: 73–76, 2000.
46. Boineau JP, Canavan TE, Schuessler RB, Cain ME, Corr PB,
Cox JL. Demonstration of a widely distributed atrial pacemaker
complex in the human heart. Circulation 77: 1221–1237, 1988.
47. Boineau JP, Schuessler RB, Hackel DB, Miller CB, Brockus
CW, Wylds AC. Widespread distribution and rate differentiation of
the atrial pacemaker complex. Am J Physiol Heart Circ Physiol
239: H406–H415, 1980.
48. 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 118: 1051–1057, 1996.
49. Bois P, Lenfant J. Evidence for two types of calcium currents in
frog cardiac sinus venosus cells. Pflu¨ gers Arch 417: 591–596, 1991.
50. Bois P, Renaudon B, Baruscotti M, Lenfant J, DiFrancesco D.
Activation of f-channels by cAMP analogues in macropatches from
rabbit sino-atrial node myocytes. J Physiol 501: 565–571, 1997.
51. Bolter CP, Atkinson KJ. Influence of temperature and adrenergic
stimulation on rat sinoatrial frequency. Am J Physiol Regul Integr
Comp Physiol 254: R840–R844, 1988.
52. Borer JS, Fox K, Jaillon P, Lerebours G. Antianginal and anti-
ischemic effects of ivabradine, an I
f
inhibitor, in stable angina: a
randomized, double-blind, multicentered, placebo-controlled trial.
Circulation 107: 817–823, 2003.
53. BoSmith RE, Briggs I, Sturgess NC. Inhibitory actions of ZEN-
ECA ZD7288 on whole-cell hyperpolarization activated inward cur-
rent (I
f
) in guinea-pig dissociated sinoatrial node cells. Br J Phar-
macol 110: 343–349, 1993.
54. Bouman LN, Gerlings ED, Biersteker PA, Bonke FI. Pace-
maker shift in the sino-atrial node during vagal stimulation.
Pflu¨ gers Arch 302: 255–267, 1968.
55. Boutjdir M. Molecular and ionic basis of congenital complete
heart block. Trends Cardiovasc Med 10: 114–122, 2000.
56. Boyett M, Zhang H, Garny A, Holden AV. Control of the pace-
maker activity of the sinoatrial node by intracellular Ca
2
. Exper-
iments and modelling. Philos Trans R Soc Lond B Biol Sci 359:
1091–1110, 2001.
57. Boyett MR, Dobrzynski H, Lancaster MK, Jones SA, Honjo H,
Kodama I. Sophisticated architecture is required for the sinoatrial
node to perform its normal pacemaker function. J Cardiovasc
Electrophysiol 14: 104–106, 2003.
58. Boyett MR, Honjo H, Kodama I. The sinoatrial node, a hetero-
geneous pacemaker structure. Cardiovasc Res 47: 658 687, 2000.
59. Boyett MR, Honjo H, Yamamoto M, Nikmaram MR, Niwa R,
Kodama I. Downward gradient in action potential duration along
conduction path in and around the sinoatrial node. Am J Physiol
Heart Circ Physiol 276: H686–H698, 1999.
60. Boyett MR, Honjo H, Yamamoto M, Nikmaram MR, Niwa R,
Kodama I. Regional differences in effects of 4-aminopyridine
within the sinoatrial node. Am J Physiol Heart Circ Physiol 275:
H1158–H1168, 1998.
61. Boyett MR, Inada S, Yoo S, Li J, Liu J, Tellez J, Greener ID,
Honjo H, Billeter R, Lei M, Zhang H, Efimov IR, Dobrzynski H.
Connexins in the sinoatrial and atrioventricular nodes. Adv Cardiol
42: 175–197, 2006.
62. Brown H, Difrancesco D. Voltage-clamp investigations of mem-
brane currents underlying pace-maker activity in rabbit sino-atrial
node. J Physiol 308: 331–351, 1980.
63. Brown HF. Electrophysiology of the sinoatrial node. Physiol Rev
62: 505–530, 1982.
64. Brown HF, DiFrancesco D, Noble SJ. How does adrenaline
accelerate the heart? Nature 280: 235–236, 1979.
65. Brown HF, Kimura J, Noble D, Noble SJ, Taupignon A. The
slow inward current, i
si
, in the rabbit sino-atrial node investigated
by voltage clamp and computer simulation. Proc R Soc Lond B Biol
Sci 222: 305–328, 1984.
66. Bucchi A, Baruscotti M, DiFrancesco D. Current-dependent
block of rabbit sino-atrial node I
f
channels by ivabradine. J Gen
Physiol 120: 1–13, 2002.
970 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
67. Bucchi A, Baruscotti M, Robinson RB, DiFrancesco D. I
f
-
dependent modulation of pacemaker rate mediated by cAMP in the
presence of ryanodine in rabbit sino-atrial node cells. J Mol Cell
Cardiol 35: 905–913, 2003.
68. Bucchi A, Baruscotti M, Robinson RB, Difrancesco D. Modu-
lation of rate by autonomic agonists in SAN cells involves changes
in diastolic depolarization and the pacemaker current. J Mol Cell
Cardiol 43: 39 48, 2007.
69. Bucchi A, Plotnikov AN, Shlapakova I, Danilo P Jr, Kryukova
Y, Qu J, Lu Z, Liu H, Pan Z, Potapova I, KenKnight B, Gir-
ouard S, Cohen IS, Brink PR, Robinson RB, Rosen MR. Wild-
type and mutant HCN channels in a tandem biological-electronic
cardiac pacemaker. Circulation 114: 992–999, 2006.
70. Bukauskas FF, Kreuzberg MM, Rackauskas M, Bukauskiene
A, Bennett MV, Verselis VK, Willecke K. Properties of mouse
connexin 30.2 and human connexin 31 9 hemichannels: implica-
tions for atrioventricular conduction in the heart. Proc Natl Acad
Sci USA 103: 9726–9731, 2006.
71. Cai J, Yi FF, Li YH, Yang XC, Song J, Jiang XJ, Jiang H, Lin
GS, Wang W. Adenoviral gene transfer of HCN4 creates a genetic
pacemaker in pigs with complete atrioventricular block. Life Sci
80: 1746–1753, 2007.
72. Callewaert G, Carmeliet E, Vereecke J. Single cardiac Purkinje
cells: general electrophysiology and voltage-clamp analysis of the
pace-maker current. J Physiol 349: 643–661, 1984.
73. Camelliti P, Green CR, LeGrice I, Kohl P. Fibroblast network in
rabbit sinoatrial node: structural and functional identification of
homogeneous and heterogeneous cell coupling. Circ Res 94: 828
835, 2004.
74. Campbell DL, Rasmusson RL, Strauss HC. Ionic current mech-
anisms generating vertebrate primary cardiac pacemaker activity
at the single cell level: an integrative view. Annu Rev Physiol 54:
279–302, 1992.
75. Carmeliet E. Cardiac ionic currents and acute ischemia: from
channels to arrhythmias. Physiol Rev 79: 917–1017, 1999.
76. Casadei B, Moon J, Johnston J, Caiazza A, Sleight P. Is respi-
ratory sinus arrhythmia a good index of cardiac vagal tone in
exercise? J Appl Physiol 81: 556–564, 1996.
77. Catterall WA. Structure and regulation of voltage-gated Ca
2
channels. Annu Rev Cell Dev Biol 16: 521–555, 2000.
78. Catterall WA, Striessnig J, Snutch TP, Perez-Reyes E. Inter-
national Union of Pharmacology XL. Compendium of voltage-gated
ion channels: calcium channels. Pharmacol Rev 55: 579–581, 2003.
79. Cerbai E, Mugelli A. I
f
in non-pacemaker cells: role and pharma-
cological implications. Pharmacol Res 53: 416 423, 2006.
80. Chang F, Cohen IS, DiFrancesco D, Rosen MR, Tromba C.
Effects of protein kinase inhibitors on canine Purkinje fiber pace-
maker depolarization and the pacemaker current i
f
. J Physiol 440:
367–384, 1991.
81. Chang F, Gao J, Tromba C, Cohen I, DiFrancesco D. Acetyl-
choline reverses effects of beta-agonists on pacemaker current in
canine cardiac Purkinje fibers but has no direct action. A difference
between primary and secondary pacemakers. Circ Res 66: 633–636,
1990.
82. Chang F, Yu H, Cohen IS. Actions of vasoactive intestinal peptide
and neuropeptide Y on the pacemaker current in canine Purkinje
fibers. Circ Res 74: 157–162, 1994.
83. Chang SL, Chen YC, Chen YJ, Wangcharoen W, Lee SH, Lin
CI, Chen SA. Mechanoelectrical feedback regulates the arrhyth-
mogenic activity of pulmonary veins. Heart 93: 82–88, 2007.
84. Chen CC, Lamping KG, Nuno DW, Barresi R, Prouty SJ,
Lavoie JL, Cribbs LL, England SK, Sigmund CD, Weiss RM,
Williamson RA, Hill JA, Campbell KP. Abnormal coronary func-
tion in mice deficient in alpha1H T-type Ca
2
channels. Science 302:
1416–1418, 2003.
85. Chiba S, Simmons TW, Levy MN. Chronotropic responses to
experimental ischemia of the canine sino auricular node. Arch Int
Physiol Biochim 84: 81–88, 1976.
86. Cho HS, Takano M, Noma A. The electrophysiological properties
of spontaneously beating pacemaker cells isolated from mouse
sinoatrial node. J Physiol 550: 169–180, 2003.
87. Choate JK, Feldman R. Neuronal control of heart rate in isolated
mouse atria. Am J Physiol Heart Circ Physiol 285: H1340 –H1346,
2003.
88. Choate JK, Klemm M, Hirst GD. Sympathetic and parasympa-
thetic neuromuscular junctions in the guinea-pig sino-atrial node. J
Auton Nerv Syst 44: 1–15, 1993.
89. Choate JK, Nandhabalan M, Paterson DJ. Raised extracellular
potassium attenuates the sympathetic modulation of sino-atrial
node pacemaking in the isolated guinea-pig atria. Exp Physiol 86:
19–25, 2001.
90. Choi HS, Wang DY, Noble D, Lee CO. Effect of isoprenaline,
carbachol, Cs
on Na
activity and pacemaker potential in rabbit
SA node cells. Am J Physiol Heart Circ Physiol 276: H205–H214,
1999.
91. Christophe J, Waelbroeck M, Chatelain P, Robberecht P.
Heart receptors for VIP, PHI and secretin are able to activate
adenylate cyclase and to mediate inotropic and chronotropic ef-
fects. Species variations and physiopathology. Peptides 5: 341–353,
1984.
92. Clancy CE, Kass RS. Inherited and acquired vulnerability to ven-
tricular arrhythmias: cardiac Na
and K
channels. Physiol Rev 85:
33–47, 2005.
93. Clark RB, Mangoni ME, Lueger A, Couette B, Nargeot J, Giles
WR. A rapidly activating delayed rectifier K
current regulates
pacemaker activity in adult mouse sinoatrial node cells. Am J
Physiol Heart Circ Physiol 286: H1757–H1766, 2004.
94. Cohen IS, Robinson RB. Pacemaker current and automatic
rhythms: toward a molecular understanding. Handbook Exp Phar-
macol 41–71, 2006.
95. Coleridge JC, Linden RJ. The effect of intravenous infusions
upon the heart rate of the anaesthetized dog. J Physiol 128: 310
319, 1955.
96. Colin P, Ghaleh B, Hittinger L, Monnet X, Slama M, Giudicelli
JF, Berdeaux A. Differential effects of heart rate reduction and
beta-blockade on left ventricular relaxation during exercise. Am J
Physiol Heart Circ Physiol 282: H672–H679, 2002.
97. Colin P, Ghaleh B, Monnet X, Su J, Hittinger L, Giudicelli JF,
Berdeaux A. Contributions of heart rate and contractility to myo-
cardial oxygen balance during exercise. Am J Physiol Heart Circ
Physiol 284: H676–H682, 2003.
98. Cooper KE. Some responses of the cardiovascular system to heat
and fever. Can J Cardiol 10: 444 448, 1994.
99. Cooper PJ, Kohl P. Species- and preparation-dependence of
stretch effects on sino-atrial node pacemaking. Ann NY Acad Sci
1047: 324–335, 2005.
100. Cooper PJ, Lei M, Cheng LX, Kohl P. Selected contribution:
axial stretch increases spontaneous pacemaker activity in rabbit
isolated sinoatrial node cells. J Appl Physiol 89: 2099–2104, 2000.
101. Coraboeuf E, Deroubaix E, Coulombe A. Effect of tetrodotoxin
on action potentials of the conducting system in the dog heart.
Am J Physiol Heart Circ Physiol 236: H561–H567, 1979.
102. Crampin EJ, Halstead M, Hunter P, Nielsen P, Noble D, Smith
N, Tawhai M. Computational physiology and the Physiome
Project. Exp Physiol 89: 1–26, 2004.
103. Cribbs LL, Lee JH, Yang J, Satin J, Zhang Y, Daud A, Barclay
J, Williamson MP, Fox M, Rees M, Perez-Reyes E. Cloning and
characterization of alpha1H from human heart, a member of the
T-type Ca
2
channel gene family. Circ Res 83: 103–109, 1998.
104. Cribbs LL, Martin BL, Schroder EA, Keller BB, Delisle BP,
Satin J. Identification of the t-type calcium channel [Ca(v)3.1d] in
developing mouse heart. Circ Res 88: 403–407, 2001.
105. De Maziere AM, van Ginneken AC, Wilders R, Jongsma HJ,
Bouman LN. Spatial and functional relationship between myo-
cytes and fibroblasts in the rabbit sinoatrial node. J Mol Cell
Cardiol 24: 567–578, 1992.
106. De Mello WC. Role of chloride ions in cardiac action and pace-
maker potential. Am J Physiol 205: 567–575, 1963.
107. De Neef P, Robberecht P, Chatelain P, Waelbroeck M,
Christophe J. The in vitro chronotropic and inotropic effects of
vasoactive intestinal peptide (VIP) on the atria and ventricular
papillary muscle from Cynomolgus monkey heart. Regul Pept 8:
237–244, 1984.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
971
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
108. Deck KA. Effects of stretch on the spontaneously beating, isolated
sinus node. Pflu¨ gers Arch 280: 120–130, 1964.
109. Demion M, Bois P, Launay P, Guinamard R. TRPM4, a Ca
2
-
activated nonselective cation channel in mouse sino-atrial node
cells. Cardiovasc Res 73: 531–538, 2007.
110. Demir SS, Clark JW, Giles WR. Parasympathetic modulation of
sinoatrial node pacemaker activity in rabbit heart: a unifying
model. Am J Physiol Heart Circ Physiol 276: H2221–H2244, 1999.
111. Demir SS, Clark JW, Murphey CR, Giles WR. A mathematical
model of a rabbit sinoatrial node cell. Am J Physiol Cell Physiol
266: C832–C852, 1994.
112. Demolombe S, Marionneau C, Le Bouter S, Charpentier F,
Escande D. Functional genomics of cardiac ion channel genes.
Cardiovasc Res 67: 438 447, 2005.
113. Denyer JC, Brown HF. Pacemaking in rabbit isolated sino-atrial
node cells during Cs
block of the hyperpolarization-activated
current i
f
. J Physiol 429: 401–409, 1990.
114. Denyer JC, Brown HF. Rabbit sino-atrial node cells: isolation and
electrophysiological properties. J Physiol 428: 405–424, 1990.
115. DiFrancesco D. Characterization of single pacemaker channels in
cardiac sino-atrial node cells. Nature 324: 470 473, 1986.
116. DiFrancesco D. The contribution of the “pacemaker” current (i
f
)
to generation of spontaneous activity in rabbit sino-atrial node
myocytes. J Physiol 434: 23–40, 1991.
117. DiFrancesco D. Funny channels in the control of cardiac rhythm
and mode of action of selective blockers. Pharmacol Res 53: 399
406, 2006.
118. DiFrancesco D. A new interpretation of the pace-maker current in
calf Purkinje fibers. J Physiol 314: 359–376, 1981.
119. DiFrancesco D. The onset and autonomic regulation of cardiac
pacemaker activity: relevance of the f current. Cardiovasc Res 29:
449 456, 1995.
120. DiFrancesco D. Pacemaker mechanisms in cardiac tissue. Annu
Rev Physiol 55: 455–472, 1993.
121. DiFrancesco D. Serious workings of the funny current. Prog Bio-
phys Mol Biol 90: 13–25, 2006.
122. DiFrancesco D. A study of the ionic nature of the pace-maker
current in calf Purkinje fibers. J Physiol 314: 377–393, 1981.
123. DiFrancesco D, Camm JA. Heart rate lowering by specific and
selective I
f
current inhibition with ivabradine: a new therapeutic
perspective in cardiovascular disease. Drugs 64: 1757–1765, 2004.
124. DiFrancesco D, Ducouret P, Robinson RB. Muscarinic modula-
tion of cardiac rate at low acetylcholine concentrations. Science
243: 669 671, 1989.
125. DiFrancesco D, Ferroni A, Mazzanti M, Tromba C. Properties
of the hyperpolarizing-activated current (i
f
) in cells isolated from
the rabbit sino-atrial node. J Physiol 377: 61–88, 1986.
126. DiFrancesco D, Mangoni M. Modulation of single hyperpolariza-
tion-activated channels [i
f
] by cAMP in the rabbit sino-atrial node.
J Physiol 474: 473–482, 1994.
127. DiFrancesco D, Noble D. A model of cardiac electrical activity
incorporating ionic pumps and concentration changes. Philos
Trans R Soc Lond B Biol Sci 307: 353–398, 1985.
128. DiFrancesco D, Tortora P. Direct activation of cardiac pace-
maker channels by intracellular cyclic AMP. Nature 351: 145–147,
1991.
129. DiFrancesco D, Tromba C. Muscarinic control of the hyperpo-
larization-activated current (i
f
) in rabbit sino-atrial node myocytes.
J Physiol 405: 493–510, 1988.
130. Dobrzynski H, Li J, Tellez J, Greener ID, Nikolski VP, Wright
SE, Parson SH, Jones SA, Lancaster MK, Yamamoto M, Honjo
H, Takagishi Y, Kodama I, Efimov IR, Billeter R, Boyett MR.
Computer three-dimensional reconstruction of the sinoatrial node.
Circulation 111: 846 854, 2005.
131. Dobrzynski H, Nikolski VP, Sambelashvili AT, Greener ID,
Yamamoto M, Boyett MR, Efimov IR. Site of origin and molec-
ular substrate of atrioventricular junctional rhythm in the rabbit
heart. Circ Res 93: 1102–1110, 2003.
132. Du XJ, Feng X, Gao XM, Tan TP, Kiriazis H, Dart AM. I
f
channel inhibitor ivabradine lowers heart rate in mice with en-
hanced sympathoadrenergic activities. Br J Pharmacol 142: 107–
112, 2004.
133. Du YM, Nathan RD. Ionic basis of ischemia-induced bradycardia
in the rabbit sinoatrial node. J Mol Cell Cardiol 42: 315–325. 2007.
134. Dyer AR, Persky V, Stamler J, Paul O, Shekelle RB, Berkson
DM, Lepper M, Schoenberger JA, Lindberg HA. Heart rate as a
prognostic factor for coronary heart disease and mortality: findings
in three Chicago epidemiologic studies. Am J Epidemiol 112: 736
749, 1980.
135. Dzhura I, Wu Y, Colbran RJ, Balser JR, Anderson ME. Cal-
modulin kinase determines calcium-dependent facilitation of L-
type calcium channels. Nat Cell Biol 2: 173–177, 2000.
136. Ebert SN, Taylor DG. Catecholamines and development of car-
diac pacemaking: an intrinsically intimate relationship. Cardiovasc
Res 72: 364–374, 2006.
137. Edelberg JM, Aird WC, Rosenberg RD. Enhancement of murine
cardiac chronotropy by the molecular transfer of the human beta2
adrenergic receptor cDNA. J Clin Invest 101: 337–343, 1998.
138. Edelberg JM, Huang DT, Josephson ME, Rosenberg RD. Mo-
lecular enhancement of porcine cardiac chronotropy. Heart 86:
559–562, 2001.
139. Efimov IR, Mazgalev TN. High-resolution, three-dimensional flu-
orescent imaging reveals multilayer conduction pattern in the atrio-
ventricular node. Circulation 98: 54–57, 1998.
140. Efimov IR, Nikolski VP, Rothenberg F, Greener ID, Li J,
Dobrzynski H, Boyett M. Structure-function relationship in the
AV junction. Anat Rec 280: 952–965, 2004.
141. Ehrlich JR, Cha TJ, Zhang L, Chartier D, Melnyk P, Hohnloser
SH, Nattel S. Cellular electrophysiology of canine pulmonary vein
cardiomyocytes: action potential and ionic current properties.
J Physiol 551: 801–813, 2003.
142. Eisner DA, Choi HS, Diaz ME, O’Neill SC, Trafford AW. Inte-
grative analysis of calcium cycling in cardiac muscle. Circ Res 87:
1087–1094, 2000.
143. Er F, Larbig R, Ludwig A, Biel M, Hofmann F, Beuckelmann
DJ, Hoppe UC. Dominant-negative suppression of HCN channels
markedly reduces the native pacemaker current I
f
and undermines
spontaneous beating of neonatal cardiomyocytes. Circulation 107:
485–489, 2003.
144. Feld Y, Melamed-Frank M, Kehat I, Tal D, Marom S, Gepstein
L. Electrophysiological modulation of cardiomyocytic tissue by
transfected fibroblasts expressing potassium channels: a novel
strategy to manipulate excitability. Circulation 105: 522–529, 2002.
145. Fermini B, Nathan RD. Removal of sialic acid alters both T- and
L-type calcium currents in cardiac myocytes. Am J Physiol Heart
Circ Physiol 260: H735–H743, 1991.
146. Ferron L, Capuano V, Deroubaix E, Coulombe A, Renaud JF.
Functional and molecular characterization of a T-type Ca
2
chan-
nel during fetal and postnatal rat heart development. J Mol Cell
Cardiol 34: 533–546, 2002.
147. Furuse A, Kotsuka Y, Asano K. Sinus node potential during cold
cardioplegia. Jpn J Surg 13: 146–151, 1983.
148. Gaborit N, Le Bouter S, Szuts V, Varro A, Escande D, Nattel
S, Demolombe S. Regional and tissue specific transcript signa-
tures of ion channel genes in the non-diseased human heart.
J Physiol 582: 675–693, 2007.
149. Gadsby DC, Cranefield PF. Direct measurement of changes in
sodium pump current in canine cardiac Purkinje fibers. Proc Natl
Acad Sci USA 76: 1783–1787, 1979.
150. Gadsby DC, Kimura J, Noma A. Voltage dependence of Na/K
pump current in isolated heart cells. Nature 315: 63–65, 1985.
151. Gannier F, White E, Lacampagne A, Garnier D, Le Guennec
JY. Streptomycin reverses a large stretch induced increase in
[Ca
2
]
i
in isolated guinea pig ventricular myocytes. Cardiovasc Res
28: 1193–1198, 1994.
152. Garny A, Kohl P, Hunter PJ, Boyett MR, Noble D. One-dimen-
sional rabbit sinoatrial node models: benefits and limitations.
J Cardiovasc Electrophysiol 14: S121–132, 2003.
153. Garny A, Noble D, Kohl P. Dimensionality in cardiac modelling.
Prog Biophys Mol Biol 87: 47–66, 2005.
154. Gaskell WH. On the innervation of the heart, with special refer-
ence to the heart of the tortoise. J Physiol 4: 43–214, 1883.
155. Gaskell WH. Preliminary observations on the innervation of the
heart of the tortoise. J Physiol 3: 369–379, 1882.
972 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
156. Gaul GB, Gruska M, Titscher G, Blazek G, Havelec L, Marktl
W, Muellner W, Kaff A. Prediction of survival after out-of-hospital
cardiac arrest: results of a community-based study in Vienna. Re-
suscitation 32: 169–176, 1996.
157. Gehrmann J, Hammer PE, Maguire CT, Wakimoto H, Tried-
man JK, Berul CI. Phenotypic screening for heart rate variability
in the mouse. Am J Physiol Heart Circ Physiol 279: H733–H740,
2000.
158. Gehrmann J, Meister M, Maguire CT, Martins DC, Hammer
PE, Neer EJ, Berul CI, Mende U. Impaired parasympathetic
heart rate control in mice with a reduction of functional G protein
betagamma-subunits. Am J Physiol Heart Circ Physiol 282: H445–
H456, 2002.
159. Giles W, Noble SJ. Changes in membrane currents in bullfrog
atrium produced by acetylcholine. J Physiol 261: 103–123, 1976.
160. Gillman MW, Kannel WB, Belanger A, D’Agostino RB. Influ-
ence of heart rate on mortality among persons with hypertension:
the Framingham Study. Am Heart J 125: 1148–1154, 1993.
161. Gloss B, Trost S, Bluhm W, Swanson E, Clark R, Winkfein R,
Janzen K, Giles W, Chassande O, Samarut J, Dillmann W.
Cardiac ion channel expression and contractile function in mice
with deletion of thyroid hormone receptor alpha or beta. Endocri-
nology 142: 544–550, 2001.
162. Goethals M, Raes A, van Bogaert PP. Use-dependent block of
the pacemaker current I
f
in rabbit sinoatrial node cells by zatebra-
dine (UL-FS 49). On the mode of action of sinus node inhibitors.
Circulation 88: 2389–2401, 1993.
163. Grace AA, Kirschenlohr HL, Metcalfe JC, Smith GA, Weissberg
PL, Cragoe EJ Jr, Vandenberg JI. Regulation of intracellular pH in
the perfused heart by external HCO
3
- and Na
-H
exchange. Am J
Physiol Heart Circ Physiol 265: H289–H298, 1993.
164. Gryshchenko O, Qu J, Nathan RD. Ischemia alters the electrical
activity of pacemaker cells isolated from the rabbit sinoatrial node.
Am J Physiol Heart Circ Physiol 282: H2284–H2295, 2002.
165. Guo J, Noma A. Existence of a low-threshold and sustained
inward current in rabbit atrio-ventricular node cells. Jpn J Physiol
47: 355–359, 1997.
166. Guo J, Ono K, Noma A. A sustained inward current activated at
the diastolic potential range in rabbit sino-atrial node cells.
J Physiol 483: 1–13, 1995.
167. Guo W, Kamiya K, Hojo M, Kodama I, Toyama J. Regulation of
Kv4.2 and Kv1.4 K
channel expression by myocardial hypertro-
phic factors in cultured newborn rat ventricular cells. J Mol Cell
Cardiol 30: 1449–1455, 1998.
168. Guo W, Li H, Aimond F, Johns DC, Rhodes KJ, Trimmer JS,
Nerbonne JM. Role of heteromultimers in the generation of myo-
cardial transient outward K
currents. Circ Res 90: 586 –593, 2002.
169. Habuchi Y, Lu LL, Morikawa J, Yoshimura M. Angiotensin II
inhibition of L-type Ca
2
current in sinoatrial node cells of rabbits.
Am J Physiol Heart Circ Physiol 268: H1053–H1060, 1995.
170. Habuchi Y, Nishio M, Tanaka H, Yamamoto T, Lu LL, Yo-
shimura M. Regulation by acetylcholine of Ca
2
current in rabbit
atrioventricular node cells. Am J Physiol Heart Circ Physiol 271:
H2274–H2282, 1996.
171. Hagiwara N, Irisawa H. Modulation by intracellular Ca
2
of the
hyperpolarization-activated inward current in rabbit single sino-
atrial node cells. J Physiol 409: 121–141, 1989.
172. Hagiwara N, Irisawa H, Kameyama M. Contribution of two types
of calcium currents to the pacemaker potentials of rabbit sino-
atrial node cells. J Physiol 395: 233–253, 1988.
173. Hagiwara N, Irisawa H, Kasanuki H, Hosoda S. Background
current in sino-atrial node cells of the rabbit heart. J Physiol 448:
53–72, 1992.
174. Hagiwara N, Masuda H, Shoda M, Irisawa H. Stretch-activated
anion currents of rabbit cardiac myocytes. J Physiol 456: 285–302,
1992.
175. Haissaguerre M, Jais P, Shah DC, Takahashi A, Hocini M,
Quiniou G, Garrigue S, Le Mouroux A, Le Metayer P,
Clementy J. Spontaneous initiation of atrial fibrillation by ectopic
beats originating in the pulmonary veins. N Engl J Med 339: 659
666, 1998.
176. Haissaguerre M, Shah DC, Jais P, Shoda M, Kautzner J,
Arentz T, Kalushe D, Kadish A, Griffith M, Gaita F, Yamane T,
Garrigue S, Hocini M, Clementy J. Role of Purkinje conducting
system in triggering of idiopathic ventricular fibrillation. Lancet
359: 677–678, 2002.
177. Hakumaki MO. Seventy years of the Bainbridge reflex. Acta
Physiol Scand 130: 177–185, 1987.
178. Hale SL, Kloner RA. Elevated body temperature during myocar-
dial ischemia/reperfusion exacerbates necrosis and worsens no-
reflow. Coron Artery Dis 13: 177–181, 2002.
179. Han W, Bao W, Wang Z, Nattel S. Comparison of ion-channel
subunit expression in canine cardiac Purkinje fibers and ventricu-
lar muscle. Circ Res 91: 790–797, 2002.
180. Han X, Habuchi Y, Giles WR. Effects of metabolic inhibition on
action potentials and ionic currents in cardiac pacemaking cells
(Abstract). Circulation 90 Suppl I: 582, 1994.
181. Han X, Habuchi Y, Giles WR. Relaxin increases heart rate by
modulating calcium current in cardiac pacemaker cells. Circ Res
74: 537–541, 1994.
182. Han X, Kobzik L, Severson D, Shimoni Y. Characteristics of
nitric oxide-mediated cholinergic modulation of calcium current in
rabbit sino-atrial node. J Physiol 509: 741–754, 1998.
183. Han X, Kubota I, Feron O, Opel DJ, Arstall MA, Zhao YY,
Huang P, Fishman MC, Michel T, Kelly RA. Muscarinic cholin-
ergic regulation of cardiac myocyte I
CaL
is absent in mice with
targeted disruption of endothelial nitric oxide synthase. Proc Natl
Acad Sci USA 95: 6510 6515, 1998.
184. Han X, Light PE, Giles WR, French RJ. Identification and prop-
erties of an ATP-sensitive K
current in rabbit sino-atrial node
pacemaker cells. J Physiol 490: 337–350, 1996.
185. Han X, Shimoni Y, Giles WR. A cellular mechanism for nitric
oxide-mediated cholinergic control of mammalian heart rate. J Gen
Physiol 106: 45–65, 1995.
186. Han X, Shimoni Y, Giles WR. An obligatory role for nitric oxide
in autonomic control of mammalian heart rate. J Physiol 476:
309–314, 1994.
187. Hancox JC, Levi AJ, Lee CO, Heap P. A method for isolating
rabbit atrioventricular node myocytes which retain normal mor-
phology and function. Am J Physiol Heart Circ Physiol 265: H755–
H766, 1993.
188. Hanna CM, Greenes DS. How much tachycardia in infants can be
attributed to fever? Ann Emerg Med 43: 699–705, 2004.
189. Hara M, Liu YM, Zhen L, Cohen IS, Yu H, Danilo P Jr, Ogino
K, Bilezikian JP, Rosen MR. Positive chronotropic actions of
parathyroid hormone and parathyroid hormone-related peptide are
associated with increases in the current, I
f
, the slope of the pace-
maker potential. Circulation 96: 3704–3709, 1997.
190. Hartzell HC. Regulation of cardiac ion channels by catecholamines,
acetylcholine and second messenger systems. Prog Biophys Mol Biol
52: 165–247, 1988.
190a.Harzheim D, Pfeiffer KH, Fabritz L, Kremmer E, Buch T,
Waisman A, Kirchhof P, Kaupp UB, Seifert R. Cardiac pace-
maker function of HCN4 channels in mice is confined to embryonic
development and requires cyclic AMP. Embo J 27: 692–703, 2008.
191. Haufe V, Cordeiro JM, Zimmer T, Wu YS, Schiccitano S,
Benndorf K, Dumaine R. Contribution of neuronal sodium chan-
nels to the cardiac fast sodium current I
Na
is greater in dog heart
Purkinje fibers than in ventricles. Cardiovasc Res 65: 117–127,
2005.
192. Hauswirth O, Noble D, Tsien RW. Adrenaline: mechanism of
action on the pacemaker potential in cardiac Purkinje fibers. Sci-
ence 162: 916–917, 1968.
193. Haverinen J, Vornanen M. Temperature acclimation modifies
sinoatrial pacemaker mechanism of the rainbow trout heart. Am J
Physiol Regul Integr Comp Physiol 292: R1023–R1032, 2007.
194. Heath BM, Terrar DA. Protein kinase C enhances the rapidly
activating delayed rectifier potassium current, I
Kr
, through a reduc-
tion in C-type inactivation in guinea-pig ventricular myocytes.
J Physiol 522 Pt 3: 391–402, 2000.
195. Henning RJ. Vagal stimulation during muscarinic and beta-adren-
ergic blockade increases atrial contractility and heart rate. J Auton
Nerv Syst 40: 121–129, 1992.
196. 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 26: 4423–4432, 2007.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
973
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
197. Hilgemann DW, Noble D. Excitation-contraction coupling and
extracellular calcium transients in rabbit atrium: reconstruction of
basic cellular mechanisms. Proc R Soc Lond B Biol Sci 230: 163–
205, 1987.
198. Hilgemann DW, Yaradanakul A, Wang Y, Fuster D. Molecular
control of cardiac sodium homeostasis in health and disease.
J Cardiovasc Electrophysiol 17 Suppl 1: S47–S56, 2006.
199. Hirano Y, Fozzard HA, January CT. Characteristics of L- and
T-type Ca
2
currents in canine cardiac Purkinje cells. Am J Physiol
Heart Circ Physiol 256: H1478–H1492, 1989.
200. Hirst GD, Choate JK, Cousins HM, Edwards FR, Klemm MF.
Transmission by post-ganglionic axons of the autonomic nervous
system: the importance of the specialized neuroeffector junction.
Neuroscience 73: 7–23, 1996.
201. Hjalmarson A. Significance of reduction in heart rate in cardio-
vascular disease. Clin Cardiol 21: II3–7, 1998.
202. Hof TO, Mackaay AJ, Bleeker WK, Houtkooper JM, Abels R,
Bouman LN. Dependence of the chronotropic effects of calcium,
magnesium and sodium on temperature and cycle length in isolated
rabbit atria. J Pharmacol Exp Ther 212: 183–189, 1980.
203. Hoffman BF, Cranefield PF. Electrophysiology of the Heart.
Mount Kisco, NY: Futura, 1976.
204. Honjo H, Boyett MR, Kodama I, Toyama J. Correlation between
electrical activity and the size of rabbit sino-atrial node cells.
J Physiol 496: 795–808, 1996.
205. Honjo H, Boyett MR, Niwa R, Inada S, Yamamoto M, Mitsui K,
Horiuchi T, Shibata N, Kamiya K, Kodama I. Pacing-induced
spontaneous activity in myocardial sleeves of pulmonary veins
after treatment with ryanodine. Circulation 107: 1937–1943, 2003.
206. Honjo H, Lei M, Boyett MR, Kodama I. Heterogeneity of 4-ami-
nopyridine-sensitive current in rabbit sinoatrial node cells. Am J
Physiol Heart Circ Physiol 276: H1295–H1304, 1999.
207. Hoogaars WM, Barnett P, Moorman AF, Christoffels VM. T-
box factors determine cardiac design. Cell Mol Life Sci 64: 646
660, 2007.
208. Hoogaars WM, Engel A, Brons JF, Verkerk AO, de Lange FJ,
Wong LY, Bakker ML, Clout DE, Wakker V, Barnett P,
Ravesloot JH, Moorman AF, Verheijck EE, Christoffels VM.
Tbx3 controls the sinoatrial node gene program and imposes pace-
maker function on the atria. Genes Dev 21: 1098–1112, 2007.
209. Hu K, Qu Y, Yue Y, Boutjdir M. Functional basis of sinus brady-
cardia in congenital heart block. Circ Res 94: e32–38, 2004.
210. Hume JR, Duan D, Collier ML, Yamazaki J, Horowitz B. Anion
transport in heart. Physiol Rev 80: 31–81, 2000.
211. Huser J, Blatter LA, Lipsius SL. Intracellular Ca
2
release con-
tributes to automaticity in cat atrial pacemaker cells. J Physiol 524:
415–422, 2000.
212. Hutter OF, Trautwein W. Effect of vagal stimulation on the sinus
venosus of the frog’s heart. Nature 176: 512–513, 1955.
213. Hutter OF, Trautwein W. Vagal and sympathetic effects on the
pacemaker fibers in the sinus venosus of the heart. J Gen Physiol
39: 715–733, 1956.
214. Ino M, Yoshinaga T, Wakamori M, Miyamoto N, Takahashi E,
Sonoda J, Kagaya T, Oki T, Nagasu T, Nishizawa Y, Tanaka I,
Imoto K, Aizawa S, Koch S, Schwartz A, Niidome T, Sawada
K, Mori Y. Functional disorders of the sympathetic nervous system
in mice lacking the alpha 1B subunit (Cav 2.2) of N-type calcium
channels. Proc Natl Acad Sci USA 98: 5323–5328, 2001.
215. Irisawa H, Brown HF, Giles W. Cardiac pacemaking in the sino-
atrial node. Physiol Rev 73: 197–227, 1993.
216. Isenberg G, Klockner U. Calcium tolerant ventricular myocytes
prepared by preincubation in a “KB medium”. Pflu¨ gers Arch 395:
6–18, 1982.
217. 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 274: 12835–12839, 1999.
218. Ito H, Ono K. A rapidly activating delayed rectifier K
channel in
rabbit sinoatrial node cells. Am J Physiol Heart Circ Physiol 269:
H443–H452, 1995.
219. Jacquemet V. Pacemaker activity resulting from the coupling with
nonexcitable cells. Phys Rev E Stat Nonlin Soft Matter Phys 74:
011908, 2006.
220. James TN. Structure and function of the sinus node, AV node and
his bundle of the human heart: part II–function. Prog Cardiovasc
Dis 45: 327–360, 2003.
221. Janse MJ. Electrophysiological changes in heart failure and their
relationship to arrhythmogenesis. Cardiovasc Res 61: 208–217,
2004.
222. Jones SA, Boyett MR, Lancaster MK. Declining into failure: the
age-dependent loss of the L-type calcium channel within the sino-
atrial node. Circulation 115: 1183–1190, 2007.
223. Jose AD, Collison D. The normal range and determinants of the
intrinsic heart rate in man. Cardiovasc Res 4: 160–167, 1970.
224. Joyner RW, Kumar R, Golod DA, Wilders R, Jongsma HJ,
Verheijck EE, Bouman L, Goolsby WN, Van Ginneken AC.
Electrical interactions between a rabbit atrial cell and a nodal cell
model. Am J Physiol Heart Circ Physiol 274: H2152–H2162, 1998.
225. Joyner RW, van Capelle FJ. Propagation through electrically
coupled cells. How a small SA node drives a large atrium. Biophys
J 50: 1157–1164, 1986.
226. Ju YK, Allen DG. The distribution of calcium in toad cardiac
pacemaker cells during spontaneous firing. Pflu¨ gers Arch 441: 219
227, 2000.
227. Ju YK, Allen DG. How does beta-adrenergic stimulation increase
the heart rate? The role of intracellular Ca
2
release in amphibian
pacemaker cells. J Physiol 516: 793–804, 1999.
228. Ju YK, Allen DG. Intracellular calcium and Na
-Ca
2
exchange
current in isolated toad pacemaker cells. J Physiol 508: 153–166,
1998.
229. Ju YK, Chu Y, Chaulet H, Lai D, Gervasio OL, Graham RM,
Cannell MB, Allen DG. Store-operated Ca
2
influx and expres-
sion of TRPC genes in mouse sinoatrial node. Circ Res 100: 1605–
1614, 2007.
230. Kanani S, Pumir A, Krinski V. Genetically engineered cardiac
pacemaker: stem cells transfected with HCN2 gene and a myo-
cyte–a model. Physics Letters A. In press.
231. Kannel WB, Ho K, Thom T. Changing epidemiological features of
cardiac failure. Br Heart J 72: S3–9, 1994.
232. Kannel WB, Kannel C, Paffenbarger RS Jr, Cupples LA. Heart
rate and cardiovascular mortality: the Framingham Study. Am
Heart J 113: 1489–1494, 1987.
233. Kaufmann R, Theophile U. [Autonomously promoted extension
effect in Purkinje fibers, papillary muscles and trabeculae carneae
of rhesus monkeys]. Pflu¨ gers Arch 297: 174–189, 1967.
234. Kaupp UB, Seifert R. Molecular diversity of pacemaker ion chan-
nels. Annu Rev Physiol 63: 235–257, 2001.
235. Ke Y, Lei M, Collins TP, Rakovic S, Mattick PA, Yamasaki M,
Brodie MS, Terrar DA, Solaro RJ. Regulation of L-type calcium
channel and delayed rectifier potassium channel activity by p21-
activated kinase-1 in guinea pig sinoatrial node pacemaker cells.
Circ Res 100: 1317–1327, 2007.
236. Keating MT, Sanguinetti MC. Molecular and cellular mecha-
nisms of cardiac arrhythmias. Cell 104: 569–580, 2001.
237. Keith A, Flack M. The form and nature of the muscular connec-
tions between the primary divisions of the vertebrate heart. J Anat
Physiol 41: 172–189, 1907.
238. Kennedy ME, Nemec J, Corey S, Wickman K, Clapham DE.
GIRK4 confers appropriate processing and cell surface localization
to G-protein-gated potassium channels. J Biol Chem 274: 2571–
2582, 1999.
239. Kiekkas P, Brokalaki H, Manolis E, Askotiri P, Karga M,
Baltopoulos GI. Fever and standard monitoring parameters of
ICU patients: a descriptive study. Intensive Crit Care Nurs 23:
281–288, 2007.
240. Kim EM, Choy Y, Vassalle M. Mechanisms of suppression and
initiation of pacemaker activity in guinea-pig sino-atrial node su-
perfused in high [K
]
o
. J Mol Cell Cardiol 29: 1433–1445, 1997.
241. Kirchhof CJ, Bonke FI, Allessie MA, Lammers WJ. The influ-
ence of the atrial myocardium on impulse formation in the rabbit
sinus node. Pflu¨ gers Arch 410: 198–203, 1987.
242. Kirchhof P, Fabritz L, Fortmuller L, Matherne GP, Lankford
A, Baba HA, Schmitz W, Breithardt G, Neumann J, Boknik P.
Altered sinus nodal and atrioventricular nodal function in freely
moving mice overexpressing the A1 adenosine receptor. Am J
Physiol Heart Circ Physiol 285: H145–H153, 2003.
974 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
243. Kizana E, Ginn SL, Allen DG, Ross DL, Alexander IE. Fibro-
blasts can be genetically modified to produce excitable cells capa-
ble of electrical coupling. Circulation 111: 394–398, 2005.
244. Kizana E, Ginn SL, Smyth CM, Boyd A, Thomas SP, Allen DG,
Ross DL, Alexander IE. Fibroblasts modulate cardiomyocyte
excitability: implications for cardiac gene therapy. Gene Ther 13:
1611–1615, 2006.
245. Kleber AG, Rudy Y. Basic mechanisms of cardiac impulse prop-
agation and associated arrhythmias. Physiol Rev 84: 431– 488, 2004.
246. Klein I. Thyroid hormone and the cardiovascular system. Am J
Med 88: 631–637, 1990.
247. Kobayashi M, Godin D, Nadeau R. Sinus node responses to
perfusion pressure changes, ischaemia and hypothermia in the
isolated blood-perfused dog atrium. Cardiovasc Res 19: 20–26,
1985.
248. Kodama I, Boyett MR. Regional differences in the electrical
activity of the rabbit sinus node. Pflu¨ gers Arch 404: 214 –226, 1985.
249. Kodama I, Boyett MR, Nikmaram MR, Yamamoto M, Honjo H,
Niwa R. Regional differences in effects of E-4031 within the sino-
atrial node. Am J Physiol Heart Circ Physiol 276: H793–H802,
1999.
250. Kodama I, Nikmaram MR, Boyett MR, Suzuki R, Honjo H,
Owen JM. Regional differences in the role of the Ca
2
and Na
currents in pacemaker activity in the sinoatrial node. Am J Physiol
Heart Circ Physiol 272: H2793–H2806, 1997.
251. Kohl P, Bollensdorff C, Garny A. Effects of mechanosensitive
ion channels on ventricular electrophysiology: experimental and
theoretical models. Exp Physiol 91: 307–321, 2006.
252. Kohl P, Kamkin AG, Kiseleva IS, Noble D. Mechanosensitive
fibroblasts in the sino-atrial node region of rat heart: interaction
with cardiomyocytes and possible role. Exp Physiol 79: 943–956,
1994.
253. Kohlhardt M, Mnich Z, Maier G. Alterations of the excitation
process of the sinoatrial pacemaker cell in the presence of anoxia
and metabolic inhibitors. J Mol Cell Cardiol 9: 477–488, 1977.
254. Koschak A, Reimer D, Huber I, Grabner M, Glossmann H,
Engel J, Striessnig J. Alpha 1D (Cav1.3) subunits can form L-type
Ca
2
channels activating at negative voltages. J Biol Chem 276:
22100–22106, 2001.
255. Kreitner D. Electrophysiological study of the two main pace-
maker mechanisms in the rabbit sinus node. Cardiovasc Res 19:
304–318, 1985.
256. Kreuzberg MM, Schrickel JW, Ghanem A, Kim JS, Degen J,
Janssen-Bienhold U, Lewalter T, Tiemann K, Willecke K. Con-
nexin 30.2 containing gap junction channels decelerate impulse
propagation through the atrioventricular node. Proc Natl Acad Sci
USA 103: 5959–5964, 2006.
257. Kreuzberg MM, Sohl G, Kim JS, Verselis VK, Willecke K,
Bukauskas FF. Functional properties of mouse connexin30.2 ex-
pressed in the conduction system of the heart. Circ Res 96: 1169
1177, 2005.
258. Kreuzberg MM, Willecke K, Bukauskas FF. Connexin-mediated
cardiac impulse propagation: connexin 30.2 slows atrioventricular
conduction in mouse heart. Trends Cardiovasc Med 16: 266 –272,
2006.
259. Kuniyoshi Y, Koja K, Miyagi K, Mistuyoshi S, Uezu T, Arakaki
K, Yamashiro S, Mabuni K, Haneji S. Management of the heart
rate during coronary artery bypass grafting on the beating heart:
newly devised methods of decreasing heart rate–a preliminary
report. Ann Thorac Cardiovasc Surg 7: 358–367, 2001.
260. Kuo HC, Cheng CF, Clark RB, Lin JJ, Lin JL, Hoshijima M,
Nguyen-Tran VT, Gu Y, Ikeda Y, Chu PH, Ross J, Giles WR,
Chien KR. A defect in the Kv channel-interacting protein 2
(KChIP2) gene leads to a complete loss of I
to
and confers suscep-
tibility to ventricular tachycardia. Cell 107: 801–813, 2001.
261. Kurachi Y, Noma A, Irisawa H. Electrogenic sodium pump in
rabbit atrio-ventricular node cell. Pflu¨ gers Arch 391: 261–266, 1981.
262. Kurata Y, Hisatome I, Imanishi S, Shibamoto T. Dynamical
description of sinoatrial node pacemaking: improved mathematical
model for primary pacemaker cell. Am J Physiol Heart Circ
Physiol 283: H2074–H2101, 2002.
263. Kurata Y, Hisatome I, Matsuda H, Shibamoto T. Dynamical
mechanisms of pacemaker generation in I
K1
-downregulated human
ventricular myocytes: insights from bifurcation analyses of a math-
ematical model. Biophys J 89: 2865–2887, 2005.
264. Kurata Y, Matsuda H, Hisatome I, Shibamoto T. Effects of
pacemaker currents on creation and modulation of human ventric-
ular pacemaker: theoretical study with application to biological
pacemaker engineering. Am J Physiol Heart Circ Physiol 292:
H701–H718, 2007.
265. Lacinova L, Klugbauer N, Hofmann F. Regulation of the calcium
channel alpha(1G) subunit by divalent cations and organic block-
ers. Neuropharmacology 39: 1254–1266, 2000.
266. Lamas GA, Lee K, Sweeney M, Leon A, Yee R, Ellenbogen K,
Greer S, Wilber D, Silverman R, Marinchak R, Bernstein R,
Mittleman RS, Lieberman EH, Sullivan C, Zorn L, Flaker
G, Schron E, Orav EJ, Goldman L. The mode selection trial
(MOST) in sinus node dysfunction: design, rationale, baseline char-
acteristics of the first 1000 patients. Am Heart J 140: 541–551, 2000.
267. Lancaster MK, Jones SA, Harrison SM, Boyett MR. Intracellu-
lar Ca
2
and pacemaking within the rabbit sinoatrial node: heter-
ogeneity of role and control. J Physiol 556: 481–494, 2004.
268. Lande G, Demolombe S, Bammert A, Moorman A, Charpen-
tier F, Escande D. Transgenic mice overexpressing human Kv-
LQT1 dominant-negative isoform. Part II: Pharmacological profile.
Cardiovasc Res 50: 328–334, 2001.
269. Laskowski MB, D’Agrosa LS. The ultrastructure of the sinu-atrial
node of the bat. Acta Anat 117: 85–101, 1983.
270. Le Bouter S, Demolombe S, Chambellan A, Bellocq C, Aimond
F, Toumaniantz G, Lande G, Siavoshian S, Baro I, Pond AL,
Nerbonne JM, Leger JJ, Escande D, Charpentier F. Microarray
analysis reveals complex remodeling of cardiac ion channel ex-
pression with altered thyroid status: relation to cellular and inte-
grated electrophysiology. Circ Res 92: 234–242, 2003.
271. Le Heuzey JY, Guize L, Valty J, Moutet JP, Kouz S, Lavergne
T, Boutjdir M, Peronneau P. Intracellular and extracellular re-
cordings of sinus node activity: comparison with estimated sino-
atrial conduction times during pacemaker shifts in rabbit heart.
Cardiovasc Res 20: 81–88, 1986.
272. Lees-Miller JP, Guo J, Somers JR, Roach DE, Sheldon RS,
Rancourt DE, Duff HJ. Selective knockout of mouse ERG1 B
potassium channel eliminates I(Kr) in adult ventricular myocytes
and elicits episodes of abrupt sinus bradycardia. Mol Cell Biol 23:
1856–1862, 2003.
273. Lees-Miller JP, Kondo C, Wang L, Duff HJ. Electrophysiologi-
cal characterization of an alternatively processed ERG K
channel
in mouse and human hearts. Circ Res 81: 719–726, 1997.
274. Lehmann H, Klein UE. Familial sinus node dysfunction with
autosomal dominant inheritance. Br Heart J 40: 1314–1316, 1978.
275. Lei M, Cooper PJ, Camelliti P, Kohl P. Role of the 293b-sensi-
tive, slowly activating delayed rectifier potassium current, i(Ks), in
pacemaker activity of rabbit isolated sino-atrial node cells. Cardio-
vasc Res 53: 68–79, 2002.
276. Lei M, Goddard C, Liu J, Leoni AL, Royer A, Fung SS, Xiao G,
Ma A, Zhang H, Charpentier F, Vandenberg JI, Colledge WH,
Grace AA, Huang CL. Sinus node dysfunction following targeted
disruption of the murine cardiac sodium channel gene Scn5a.
J Physiol 567: 387–400, 2005.
277. Lei M, Honjo H, Kodama I, Boyett MR. Characterisation of the
transient outward K
current in rabbit sinoatrial node cells. Car-
diovasc Res 46: 433–441, 2000.
278. Lei M, Honjo H, Kodama I, Boyett MR. Heterogeneous expres-
sion of the delayed-rectifier K
currents i(K,r) and i(K,s) in rabbit
sinoatrial node cells. J Physiol 535: 703–714, 2001.
279. Lei M, Jones SA, Liu J, Lancaster MK, Fung SS, Dobrzynski
H, Camelliti P, Maier SK, Noble D, Boyett MR. Requirement of
neuronal- and cardiac-type sodium channels for murine sinoatrial
node pacemaking. J Physiol 559: 835–848, 2004.
280. Leoni AL, Marionneau C, Demolombe S, Le Bouter S,
Mangoni ME, Escande D, Charpentier F. Chronic heart rate
reduction remodels ion channel transcripts in the mouse sinoatrial
node but not in the ventricle. Physiol Genomics 24: 4–12, 2005.
281. Lerman BB, Wesley RC Jr, DiMarco JP, Haines DE,
Belardinelli L. Antiadrenergic effects of adenosine on His-Pur-
kinje automaticity. Evidence for accentuated antagonism. J Clin
Invest 82: 2127–2135, 1988.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
975
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
282. Leuranguer V, Monteil A, Bourinet E, Dayanithi G, Nargeot J.
T-type calcium currents in rat cardiomyocytes during postnatal
development: contribution to hormone secretion. Am J Physiol
Heart Circ Physiol 279: H2540–H2548, 2000.
283. Levitzki A. Beta-adrenergic receptors and their mode of coupling
to adenylate cyclase. Physiol Rev 66: 819 854, 1986.
284. Levy MN. Sympathetic-parasympathetic interactions in the heart.
Circ Res 29: 437–445, 1971.
284a.Li J, Greener ID, Inada S, Nikolski VP, Yamamoto M, Hancox
JC, Zhang H, Billeter R, Efimov IR, Dobrzynski H, Boyett MR.
Computer three-dimensional reconstruction of the atrioventricular
node. Circ Res. In press.
285. Li J, Qu J, Nathan RD. Ionic basis of ryanodine’s negative chro-
notropic effect on pacemaker cells isolated from the sinoatrial
node. Am J Physiol Heart Circ Physiol 273: H2481–H2489, 1997.
286. Li M, West JW, Numann R, Murphy BJ, Scheuer T, Catterall
WA. Convergent regulation of sodium channels by protein kinase C
and cAMP-dependent protein kinase. Science 261: 1439–1442, 1993.
287. Lin W, Laitko U, Juranka PF, Morris CE. Dual stretch responses
of mHCN2 pacemaker channels: accelerated activation, acceler-
ated deactivation. Biophys J 92: 1559–1572, 2007.
288. Lipsius SL, Huser J, Blatter LA. Intracellular Ca
2
release
sparks atrial pacemaker activity. News Physiol Sci 16: 101–106,
2001.
289. Liu J, Dobrzynski H, Yanni J, Boyett MR, Lei M. Organisation
of the mouse sinoatrial node: structure and expression of HCN
channels. Cardiovasc Res 73: 729–738, 2007.
290. London B, Guo W, Pan X, Lee JS, Shusterman V, Rocco CJ,
Logothetis DA, Nerbonne JM, Hill JA. Targeted replacement of
KV1.5 in the mouse leads to loss of the 4-aminopyridine-sensitive
component of I(K,slow) and resistance to drug-induced qt prolon-
gation. Circ Res 88: 940–946, 2001.
291. London B, Wang DW, Hill JA, Bennett PB. The transient out-
ward current in mice lacking the potassium channel gene Kv1.4.
J Physiol 509: 171–182, 1998.
292. Lu HH. Shifts in pacemaker dominance within the sinoatrial region
of cat and rabbit hearts resulting from increase of extracellular
potassium. Circ Res 26: 339–346, 1970.
293. Lu ZJ, Pereverzev A, Liu HL, Weiergraber M, Henry M,
Krieger A, Smyth N, Hescheler J, Schneider T. Arrhythmia in
isolated prenatal hearts after ablation of the Cav2.3 (alpha1E)
subunit of voltage-gated Ca
2
channels. Cell Physiol Biochem 14:
11–22, 2004.
294. Ludwig A, Budde T, Stieber J, Moosmang S, Wahl C, Holthoff
K, Langebartels A, Wotjak C, Munsch T, Zong X, Feil S, Feil R,
Lancel M, Chien KR, Konnerth A, Pape HC, Biel M, Hofmann
F. Absence epilepsy and sinus dysrhythmia in mice lacking the
pacemaker channel HCN2. EMBO J 22: 216–224, 2003.
295. Ludwig A, Zong X, Hofmann F, Biel M. Structure and function of
cardiac pacemaker channels. Cell Physiol Biochem 9: 179–186,
1999.
296. Ludwig A, Zong X, Stieber J, Hullin R, Hofmann F, Biel M.
Two pacemaker channels from human heart with profoundly dif-
ferent activation kinetics. EMBO J 18: 2323–2329, 1999.
297. Lundberg JM, Hokfelt T. Multiple co-existence of peptides and
classical transmitters in peripheral autonomic and sensory neu-
rons–functional and pharmacological implications. Prog Brain Res
68: 241–262, 1986.
298. Luo CH, Rudy Y. A dynamic model of the cardiac ventricular
action potential. I. Simulations of ionic currents and concentration
changes. Circ Res 74: 1071–1096, 1994.
299. Lyashkov AE, Juhaszova M, Dobrzynski H, Vinogradova TM,
Maltsev VA, Juhasz O, Spurgeon HA, Sollott SJ, Lakatta EG.
Calcium cycling protein density and functional importance to au-
tomaticity of isolated sinoatrial nodal cells are independent of cell
size. Circ Res 100: 1723–1731, 2007.
300. Mackaay AJ, Op’t Hof T, Bleeker WK, Jongsma HJ, Bouman
LN. Interaction of adrenaline and acetylcholine on cardiac pace-
maker function. Functional inhomogeneity of the rabbit sinus
node. J Pharmacol Exp Ther 214: 417–422, 1980.
301. Mackintosh AF, Chamberlain DA. Sinus node disease affecting
both parents and both children. Eur J Cardiol 10: 117–122, 1979.
302. Macri V, Proenza C, Agranovich E, Angoli D, Accili EA. Sep-
arable gating mechanisms in a Mammalian pacemaker channel.
J Biol Chem 277: 35939–35946, 2002.
303. Magyar CE, Wang J, Azuma KK, McDonough AA. Reciprocal
regulation of cardiac Na-K-ATPase and Na/Ca exchanger: hyperten-
sion, thyroid hormone, and development. Am J Physiol Cell
Physiol 269: C675–C682, 1995.
304. Maier SK, Westenbroek RE, Yamanushi TT, Dobrzynski H,
Boyett MR, Catterall WA, Scheuer T. An unexpected require-
ment for brain-type sodium channels for control of heart rate in the
mouse sinoatrial node. Proc Natl Acad Sci USA 100: 3507–3512,
2003.
305. Malo ME, Fliegel L. Physiological role and regulation of the
Na
/H
exchanger. Can J Physiol Pharmacol 84: 1081–1095, 2006.
306. Maltsev VA, Vinogradova TM, Bogdanov KY, Lakatta EG,
Stern MD. Diastolic calcium release controls the beating rate of
rabbit sinoatrial node cells: numerical modeling of the coupling
process. Biophys J 86: 2596–2605, 2004.
307. Maltsev VA, Vinogradova TM, Lakatta EG. The emergence of a
general theory of the initiation and strength of the heartbeat.
J Pharmacol Sci 100: 338–369, 2006.
308. 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 75: 233–244, 1994.
309. Mangoni ME, Couette B, Bourinet E, Platzer J, Reimer D,
Striessnig J, Nargeot J. Functional role of L-type Cav1.3 Ca
2
channels in cardiac pacemaker activity. Proc Natl Acad Sci USA
100: 5543–5548, 2003.
310. Mangoni ME, Couette B, Marger L, Bourinet E, Striessnig J,
Nargeot J. Voltage-dependent calcium channels and cardiac pace-
maker activity: from ionic currents to genes. Prog Biophys Mol Biol
90: 38 63, 2006.
311. Mangoni ME, Fontanaud P, Noble PJ, Noble D, Benkemoun H,
Nargeot J, Richard S. Facilitation of the L-type calcium current in
rabbit sino-atrial cells: effect on cardiac automaticity. Cardiovasc
Res 48: 375–392, 2000.
312. Mangoni ME, Nargeot J. Properties of the hyperpolarization-
activated current [I
f
] in isolated mouse sino-atrial cells. Cardiovasc
Res 52: 51–64, 2001.
313. Mangoni ME, Striessnig J, Platzer J, Nargeot J. Pacemaker
currents in mouse pacemaker cells. Circulation 104: R1047, 2001.
314. Mangoni ME, Traboulsie A, Leoni AL, Couette B, Marger L,
Le Quang K, Kupfer E, Cohen-Solal A, Vilar J, Shin HS,
Escande D, Charpentier F, Nargeot J, Lory P. Bradycardia and
slowing of the atrioventricular conduction in mice lacking CaV3.1/
alpha1G T-type calcium channels. Circ Res 98: 1422–1430, 2006.
315. Mangrum JM, DiMarco JP. The evaluation and management of
bradycardia. N Engl J Med 342: 703–709, 2000.
316. Marger L, Bouly M, Malhberg-Gaudin F, Cohen-Solal A, Leoni
A, Kupfer E, Striessnig J, Nargeot J, Mangoni ME. Physiolog-
ical effects of I
f
inhibition by ivabradine in mice lacking L-type
Cav1.3 calcium channels: a differential role for for HCN and Cav1.3
channels in the genesis and regulation of heart rate. European
Society of Cardiology Conference Vienna 2007, p. P1412.
317. Marionneau C, Couette B, Liu J, Li H, Mangoni ME, Nargeot
J, Lei M, Escande D, Demolombe S. Specific pattern of ionic
channel gene expression associated with pacemaker activity in the
mouse heart. J Physiol 562: 223–234, 2005.
318. Marshall PW, Rouse W, Briggs I, Hargreaves RB, Mills SD,
McLoughlin BJ. ICI D7288, a novel sinoatrial node modulator.
J Cardiovasc Pharmacol 21: 902–906, 1993.
319. Masumiya H, Tanaka H, Shigenobu K. Effects of Ca
2
channel
antagonists on sinus node: prolongation of late phase 4 depolariza-
tion by efonidipine. Eur J Pharmacol 335: 15–21, 1997.
320. Matsuura H, Ehara T, Ding WG, Omatsu-Kanbe M, Isono T.
Rapidly and slowly activating components of delayed rectifier K
current in guinea-pig sino-atrial node pacemaker cells. J Physiol
540: 815–830, 2002.
321. Matthes J, Huber I, Haaf O, Antepohl W, Striessnig J, Herzig
S. Pharmacodynamic interaction between mibefradil and other
calcium channel blockers. Naunyn-Schmiedebergs Arch Pharma-
col 361: 578–583, 2000.
976 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
322. Matthes J, Yildirim L, Wietzorrek G, Reimer D, Striessnig J,
Herzig S. Disturbed atrio-ventricular conduction and normal con-
tractile function in isolated hearts from Cav1.3-knockout mice.
Naunyn-Schmiedebergs Arch Pharmacol 369: 554–562, 2004.
323. McDonald TF, Pelzer S, Trautwein W, Pelzer DJ. Regulation
and modulation of calcium channels in cardiac, skeletal, smooth
muscle cells. Physiol Rev 74: 365–507, 1994.
324. McRory JE, Hamid J, Doering CJ, Garcia E, Parker R, Ham-
ming K, Chen L, Hildebrand M, Beedle AM, Feldcamp L,
Zamponi GW, Snutch TP. The CACNA1F gene encodes an L-type
calcium channel with unique biophysical properties and tissue
distribution. J Neurosci 24: 1707–1718, 2004.
325. Meijler FL, Janse MJ. Morphology and electrophysiology of the
mammalian atrioventricular node. Physiol Rev 68: 608 647, 1988.
326. Mery A, Aimond F, Menard C, Mikoshiba K, Michalak M,
Puceat M. Initiation of embryonic cardiac pacemaker activity by
inositol 1,4,5-trisphosphate-dependent calcium signaling. Mol Biol
Cell 16: 2414–2423, 2005.
327. Miake J, Marban E, Nuss HB. Biological pacemaker created by
gene transfer. Nature 419: 132–133, 2002.
328. Miake J, Marban E, Nuss HB. Functional role of inward rectifier
current in heart probed by Kir2.1 overexpression and dominant-
negative suppression. J Clin Invest 111: 1529–1536, 2003.
329. Milan DJ, Peterson TA, Ruskin JN, Peterson RT, MacRae CA.
Drugs that induce repolarization abnormalities cause bradycardia
in zebrafish. Circulation 107: 1355–1358, 2003.
330. Milanesi R, Baruscotti M, Gnecchi-Ruscone T, DiFrancesco D.
Familial sinus bradycardia associated with a mutation in the car-
diac pacemaker channel. N Engl J Med 354: 151–157, 2006.
331. Miquerol L, Meysen S, Mangoni M, Bois P, van Rijen HV,
Abran P, Jongsma H, Nargeot J, Gros D. Architectural and
functional asymmetry of the His-Purkinje system of the murine
heart. Cardiovasc Res 63: 77–86, 2004.
332. Mitchell JW, Larsen JK, Best PM. Identification of the calcium
channel alpha 1E [Ca(v)2.3] isoform expressed in atrial myocytes.
Biochim Biophys Acta 1577: 17–26, 2002.
333. Mitsuiye T, Guo J, Noma A. Nicardipine-sensitive Na
-mediated
single channel currents in guinea-pig sinoatrial node pacemaker
cells. J Physiol 521: 69–79, 1999.
334. Mitsuiye T, Shinagawa Y, Noma A. Sustained inward current
during pacemaker depolarization in mammalian sinoatrial node
cells. Circ Res 87: 88–91, 2000.
335. Mobley BA, Page E. The surface area of sheep cardiac Purkinje
fibers. J Physiol 220: 547–563, 1972.
336. Moe GK, Preston JB, Burlington H. Physiologic evidence for a
dual A-V transmission system. Circ Res 4: 357–375, 1956.
337. Moffat MP. Concentration-dependent effects of prostacyclin on
the response of the isolated guinea pig heart to ischemia and
reperfusion: possible involvement of the slow inward current.
J Pharmacol Exp Ther 242: 292–299, 1987.
338. Mommersteeg MT, Hoogaars WM, Prall OW, de Gier-de Vries
C, Wiese C, Clout DE, Papaioannou VE, Brown NA, Harvey
RP, Moorman AF, Christoffels VM. Molecular pathway for the
localized formation of the sinoatrial node. Circ Res 100: 354 –362,
2007.
339. Monteil A, Chemin J, Bourinet E, Mennessier G, Lory P,
Nargeot J. Molecular and functional properties of the human
alpha(1G) subunit that forms T-type calcium channels. J Biol Chem
275: 6090 6100, 2000.
340. Monteil A, Chemin J, Leuranguer V, Altier C, Mennessier G,
Bourinet E, Lory P, Nargeot J. Specific properties of T-type
calcium channels generated by the human alpha 1I subunit. J Biol
Chem 275: 16530–16535, 2000.
341. Moorman AF, Christoffels VM. Cardiac chamber formation: de-
velopment, genes, evolution. Physiol Rev 83: 1223–1267, 2003.
342. Moosmang S, Biel M, Hofmann F, Ludwig A. Differential distri-
bution of four hyperpolarization-activated cation channels in
mouse brain. Biol Chem 380: 975–980, 1999.
343. Mori T, Hashimoto A, Takase H, Kambe T. Nitric oxide (NO) is
not involved in accentuated antagonism for chronotropy in the
isolated mouse atrium. Naunyn-Schmiedebergs Arch Pharmacol
369: 363–366, 2004.
344. Moroni A, Gorza L, Beltrame M, Gravante B, Vaccari T, Bian-
chi ME, Altomare C, Longhi R, Heurteaux C, Vitadello M,
Malgaroli A, DiFrancesco D. Hyperpolarization-activated cyclic
nucleotide-gated channel 1 is a molecular determinant of the car-
diac pacemaker current I
f
. J Biol Chem 276: 29233–29241, 2001.
345. Mubagwa K, Flameng W. Adenosine, adenosine receptors and
myocardial protection: an updated overview. Cardiovasc Res 52:
25–39, 2001.
346. Mubagwa K, Mullane K, Flameng W. Role of adenosine in the
heart and circulation. Cardiovasc Res 32: 797–813, 1996.
347. Munk AA, Adjemian RA, Zhao J, Ogbaghebriel A, Shrier A.
Electrophysiological properties of morphologically distinct cells
isolated from the rabbit atrioventricular node. J Physiol 493: 801–
818, 1996.
348. Muramatsu H, Zou AR, Berkowitz GA, Nathan RD. Character-
ization of a TTX-sensitive Na
current in pacemaker cells isolated
from rabbit sinoatrial node. Am J Physiol Heart Circ Physiol 270:
H2108–H2119, 1996.
349. Musa H, Lei M, Honjo H, Jones SA, Dobrzynski H, Lancaster
MK, Takagishi Y, Henderson Z, Kodama I, Boyett MR. Heter-
ogeneous expression of Ca(2) handling proteins in rabbit sino-
atrial node. J Histochem Cytochem 50: 311–324, 2002.
350. Nakayama T, Kurachi Y, Noma A, Irisawa H. Action potential
and membrane currents of single pacemaker cells of the rabbit
heart. Pflu¨ gers Arch 402: 248–257, 1984.
351. Nathan RD. Two electrophysiologically distinct types of cultured
pacemaker cells from rabbit sinoatrial node. Am J Physiol Heart
Circ Physiol 250: H325–H329, 1986.
352. Nerbonne JM. Molecular basis of functional voltage-gated K
channel diversity in the mammalian myocardium. J Physiol 525:
285–298, 2000.
353. Nerbonne JM, Kass RS. Molecular physiology of cardiac repolar-
ization. Physiol Rev 85: 1205–1253, 2005.
354. Nerbonne JM, Nichols CG, Schwarz TL, Escande D. Genetic
manipulation of cardiac K
channel function in mice: what have we
learned, and where do we go from here? Circ Res 89: 944–956,
2001.
355. Nikmaram MR, Boyett MR, Kodama I, Suzuki R, Honjo H.
Variation in effects of Cs
, UL-FS-49, ZD-7288 within sinoatrial
node. Am J Physiol Heart Circ Physiol 272: H2782–H2792, 1997.
356. Nikolski V, Efimov I. Fluorescent imaging of a dual-pathway
atrioventricular-nodal conduction system. Circ Res 88: E23–30,
2001.
357. Nishi K, Yoshikawa Y, Takenaka F, Akaike N. Electrical activity
of sinoatrial node cells of the rabbit surviving a long exposure to
cold Tyrode’s solution. Circ Res 41: 242–247, 1977.
358. Niwa N, Yasui K, Opthof T, Takemura H, Shimizu A, Horiba M,
Lee JK, Honjo H, Kamiya K, Kodama I. Cav 3.2 subunit under-
lies the functional T-type Ca
2
channel in murine hearts during the
embryonic period. Am J Physiol Heart Circ Physiol 286: H2257–
H2263, 2004.
359. Noble D, Denyer JC, Brown HF, DiFrancesco D. Reciprocal
role of the inward currents ib, Na and i
f
in controlling and stabi-
lizing pacemaker frequency of rabbit sino-atrial node cells. Proc R
Soc Lond B Biol Sci 250: 199–207, 1992.
360. Noble D, Noble SJ. A model of sino-atrial node electrical activity
based on a modification of the DiFrancesco-Noble (1984) equa-
tions. Proc R Soc Lond B Biol Sci 222: 295–304, 1984.
361. Nof E, Luria D, Brass D, Marek D, Lahat H, Reznik-Wolf H,
Pras E, Dascal N, Eldar M, Glikson M. Point mutation in the
HCN4 cardiac ion channel pore affecting synthesis, trafficking,
functional expression is associated with familial asymptomatic
sinus bradycardia. Circulation 116: 463–470, 2007.
362. Noma A, Irisawa H. Contribution of an electrogenic sodium pump
to the membrane potential in rabbit sinoatrial node cells. Pflu¨ gers
Arch 358: 289–301, 1975.
363. Noma A, Irisawa H, Kokobun S, Kotake H, Nishimura M,
Watanabe Y. Slow current systems in the A-V node of the rabbit
heart. Nature 285: 228–229, 1980.
364. Noma A, Kotake H, Irisawa H. Slow inward current and its role
mediating the chronotropic effect of epinephrine in the rabbit
sinoatrial node. Pflu¨ gers Arch 388: 1–9, 1980.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
977
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
365. Noma A, Morad M, Irisawa H. Does the “pacemaker current”
generate the diastolic depolarization in the rabbit SA node cells?
Pflu¨ gers Arch 397: 190–194, 1983.
366. Noma A, Nakayama T, Kurachi Y, Irisawa H. Resting K conduc-
tances in pacemaker and non-pacemaker heart cells of the rabbit.
Jpn J Physiol 34: 245–254, 1984.
367. Noma A, Trautwein W. Relaxation of the ACh-induced potassium
current in the rabbit sinoatrial node cell. Pflu¨ gers Arch 377: 193–
200, 1978.
368. Nygren A, Fiset C, Firek L, Clark JW, Lindblad DS, Clark RB,
Giles WR. Mathematical model of an adult human atrial cell: the
role of K
currents in repolarization. Circ Res 82: 63–81, 1998.
369. Oehmen CS, Giles WR, Demir SS. Mathematical model of the
rapidly activating delayed rectifier potassium current I(Kr) in rab-
bit sinoatrial node. J Cardiovasc Electrophysiol 13: 1131–1140,
2002.
370. Oei HI, Van Ginneken AC, Jongsma HJ, Bouman LN. Mecha-
nisms of impulse generation in isolated cells from the rabbit sino-
atrial node. J Mol Cell Cardiol 21: 1137–1149, 1989.
371. Ono K, I
to
H. Role of rapidly activating delayed rectifier K
current in sinoatrial node pacemaker activity. Am J Physiol Heart
Circ Physiol 269: H453–H462, 1995.
372. Ono K, Shibata S, Iijima T. Properties of the delayed rectifier
potassium current in porcine sino-atrial node cells. J Physiol 524:
51–62, 2000.
373. Op’t Hof T, Mackaay AJ, Bleeker WK, Jongsma HJ, Bouman
LN. Differences between rabbit sinoatrial pacemakers in their
response to Mg, Ca and temperature. Cardiovasc Res 17: 526–532,
1983.
374. Opthof T. The mammalian sinoatrial node. Cardiovasc Drugs Ther
1: 573–597, 1988.
375. Opthof T. The normal range and determinants of the intrinsic
heart rate in man. Cardiovasc Res 45: 177–184, 2000.
376. Opthof T, Coronel R, Rademaker HM, Vermeulen JT, Wilms-
Schopman FJ, Janse MJ. Changes in sinus node function in a
rabbit model of heart failure with ventricular arrhythmias and
sudden death. Circulation 101: 2975–2980, 2000.
377. Opthof T, de Jonge B, Jongsma HJ, Bouman LN. Functional
morphology of the mammalian sinuatrial node. Eur Heart J 8:
1249–1259, 1987.
378. Opthof T, de Jonge B, Schade B, Jongsma HJ, Bouman LN.
Cycle length dependence of the chronotropic effects of adrenaline,
acetylcholine, Ca
2
and Mg
2
in the guinea-pig sinoatrial node. J
Auton Nerv Syst 11: 349–366, 1984.
379. Ornato JP, Peberdy MA. The mystery of bradyasystole during
cardiac arrest. Ann Emerg Med 27: 576–587, 1996.
380. Pachucki J, Burmeister LA, Larsen PR. Thyroid hormone regu-
lates hyperpolarization-activated cyclic nucleotide-gated channel
(HCN2) mRNA in the rat heart. Circ Res 85: 498–503, 1999.
381. Papadatos GA, Wallerstein PM, Head CE, Ratcliff R, Brady
PA, Benndorf K, Saumarez RC, Trezise AE, Huang CL, Van-
denberg JI, Colledge WH, Grace AA. Slowed conduction and
ventricular tachycardia after targeted disruption of the cardiac
sodium channel gene Scn5a. Proc Natl Acad Sci USA 99: 6210
6215, 2002.
382. Parsonnet V, Driller J, Hudson P, Villanueva A, Rough W,
Dick L. An experimental method for thermal control of heart rate:
work in progress. Pacing Clin Electrophysiol 3: 562–567, 1980.
383. Pascarel C, Hongo K, Cazorla O, White E, Le Guennec JY.
Different effects of gadolinium on I(KR), I(KS) and I(K1) in guinea-
pig isolated ventricular myocytes. Br J Pharmacol 124: 356 –360,
1998.
384. Pauza DH, Skripka V, Pauziene N, Stropus R. Anatomical study
of the neural ganglionated plexus in the canine right atrium: impli-
cations for selective denervation and electrophysiology of the si-
noatrial node in dog. Anat Rec 255: 271–294, 1999.
385. Pauza DH, Skripka V, Pauziene N, Stropus R. Morphology,
distribution, variability of the epicardiac neural ganglionated sub-
plexuses in the human heart. Anat Rec 259: 353–382, 2000.
386. Pennisi DJ, Rentschler S, Gourdie RG, Fishman GI, Mikawa
T. Induction and patterning of the cardiac conduction system. Int
J Dev Biol 46: 765–775, 2002.
387. Perez-Reyes E. Molecular characterization of a novel family of
low voltage-activated, T-type, calcium channels. J Bioenerg Bi-
omembr 30: 313–318, 1998.
388. Perez-Reyes E. Molecular physiology of low-voltage-activated T-
type calcium channels. Physiol Rev 83: 117–161, 2003.
389. Perski A, Olsson G, Landou C, de Faire U, Theorell T,
Hamsten A. Minimum heart rate and coronary atherosclerosis:
independent relations to global severity and rate of progression of
angiographic lesions in men with myocardial infarction at a young
age. Am Heart J 123: 609 616, 1992.
390. Petit-Jacques J, Bois P, Bescond J, Lenfant J. Mechanism of
muscarinic control of the high-threshold calcium current in rabbit
sino-atrial node myocytes. Pflu¨ gers Arch 423: 21–27, 1993.
391. Petrecca K, Amellal F, Laird DW, Cohen SA, Shrier A. Sodium
channel distribution within the rabbit atrioventricular node as ana-
lysed by confocal microscopy. J Physiol 501: 263–274, 1997.
392. Pian P, Bucchi A, Robinson RB, Siegelbaum SA. Regulation of
gating and rundown of HCN hyperpolarization-activated channels
by exogenous and endogenous PIP
2
. J Gen Physiol 128: 593– 604,
2006.
393. Platzer J, Engel J, Schrott-Fischer A, Stephan K, Bova S,
Chen H, Zheng H, Striessnig J. Congenital deafness and sino-
atrial node dysfunction in mice lacking class D L-type Ca
2
chan-
nels. Cell 102: 89–97, 2000.
394. Plotnikov AN, Sosunov EA, Qu J, Shlapakova IN, Anyuk-
hovsky EP, Liu L, Janse MJ, Brink PR, Cohen IS, Robinson
RB, Danilo P Jr, Rosen MR. Biological pacemaker implanted in
canine left bundle branch provides ventricular escape rhythms that
have physiologically acceptable rates. Circulation 109: 506 –512,
2004.
395. Pollack GH. Cardiac pacemaking. Science 199: 1234, 1978.
396. Pollack GH. Cardiac pacemaking: an obligatory role of cat-
echolamines? Science 196: 731–738, 1977.
397. Pond AL, Scheve BK, Benedict AT, Petrecca K, Van Wagoner
DR, Shrier A, Nerbonne JM. Expression of distinct ERG proteins
in rat, mouse, human heart. Relation to functional I
Kr
channels.
J Biol Chem 275: 5997–6006, 2000.
398. Pond AL, Scheve BK, Benedict AT, Petrecca K, Van Wagoner
DR, Shrier A, Nerbonne JM. Expression of Distinct ERG Pro-
teins in Rat, Mouse, Human Heart. Relation to functional I
Kr
chan-
nels. J Biol Chem 275: 5997–6006, 2000.
399. Posner P, Baker SP, Epstein ML, MacIntosh BR, Buss DD.
Effects of chronic hypoxia during maturation on the negative chro-
notropic effect of [H
] on the rabbit sino-atrial node. Biol Neonate
59: 109–113, 1991.
400. Potapova I, Plotnikov A, Lu Z, Danilo P Jr, Valiunas V, Qu J,
Doronin S, Zuckerman J, Shlapakova IN, Gao J, Pan Z, Her-
ron AJ, Robinson RB, Brink PR, Rosen MR, Cohen IS. Human
mesenchymal stem cells as a gene delivery system to create cardiac
pacemakers. Circ Res 94: 952–959, 2004.
401. Protas L, DiFrancesco D, Robinson RB. L-type but not T-type
calcium current changes during postnatal development in rabbit
sinoatrial node. Am J Physiol Heart Circ Physiol 281: H1252–
H1259, 2001.
402. Qu J, Altomare C, Bucchi A, DiFrancesco D, Robinson RB.
Functional comparison of HCN isoforms expressed in ventricular
and HEK 293 cells. Pflu¨ gers Arch 444: 597–601, 2002.
403. 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 89: E8 –14, 2001.
404. Qu J, Kryukova Y, Potapova IA, Doronin SV, Larsen M, Krish-
namurthy G, Cohen IS, Robinson RB. MiRP1 modulates HCN2
channel expression and gating in cardiac myocytes. J Biol Chem
279: 43497–43502, 2004.
405. Qu J, Plotnikov AN, Danilo P Jr, Shlapakova I, Cohen IS,
Robinson RB, Rosen MR. Expression and function of a biological
pacemaker in canine heart. Circulation 107: 1106–1109, 2003.
406. Qu Y, Baroudi G, Yue Y, Boutjdir M. Novel molecular mecha-
nism involving alpha1D (Cav1.3) L-type calcium channel in auto-
immune-associated sinus bradycardia. Circulation 111: 3034 –3041,
2005.
407. Qu Y, Rogers J, Tanada T, Scheuer T, Catterall WA. Modula-
tion of cardiac Na
channels expressed in a mammalian cell line
978 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
and in ventricular myocytes by protein kinase C. Proc Natl Acad
Sci USA 91: 3289–3293, 1994.
408. Randall WC, Jones SB, Lipsius SL, Rozanski GJ. Subsidiary
atrial pace-makers and their neuronal control. In: Nervous Control
of Cardiovascular Function, edited by Randall WC. New York:
Oxford Univ. Press, 1984, p. 199–224.
409. Rastan AJ, Eckenstein JI, Hentschel B, Funkat AK, Gummert
JF, Doll N, Walther T, Falk V, Mohr FW. Emergency coronary
artery bypass graft surgery for acute coronary syndrome: beating
heart versus conventional cardioplegic cardiac arrest strategies.
Circulation 114: I477–485, 2006.
410. Renaudon B, Lenfant J, Decressac S, Bois P. Thyroid hormone
increases the conductance density of f-channels in rabbit sino-
atrial node cells. Receptors Channels 7: 1–8, 2000.
411. Rentschler S, Vaidya DM, Tamaddon H, Degenhardt K, Sas-
soon D, Morley GE, Jalife J, Fishman GI. Visualization and
functional characterization of the developing murine cardiac con-
duction system. Development 128: 1785–1792, 2001.
412. Rigel DF. Effects of neuropeptides on heart rate in dogs: compar-
ison of VIP, PHI, NPY, CGRP, NT. Am J Physiol Heart Circ Physiol
255: H311–H317, 1988.
413. Rigel DF, Lathrop DA. Vasoactive intestinal polypeptide en-
hances automaticity of supraventricular pacemakers in anesthe-
tized dogs. Am J Physiol Heart Circ Physiol 261: H463–H468, 1991.
414. Rigel DF, Lathrop DA. Vasoactive intestinal polypeptide facili-
tates atrioventricular nodal conduction and shortens atrial and
ventricular refractory periods in conscious and anesthetized dogs.
Circ Res 67: 1323–1333, 1990.
415. Rigg L, Heath BM, Cui Y, Terrar DA. Localisation and functional
significance of ryanodine receptors during beta-adrenoceptor stim-
ulation in the guinea-pig sino-atrial node. Cardiovasc Res 48: 254
264, 2000.
416. Rigg L, Mattick PA, Heath BM, Terrar DA. Modulation of the
hyperpolarization-activated current [I
f
] by calcium and calmodulin
in the guinea-pig sino-atrial node. Cardiovasc Res 57: 497–504,
2003.
417. Rigg L, Terrar DA. Possible role of calcium release from the
sarcoplasmic reticulum in pacemaking in guinea-pig sino-atrial
node. Exp Physiol 81: 877–880, 1996.
418. Robberecht P, Gillet L, Chatelain P, De Neef P, Camus JC,
Vincent M, Sassard J, Christophe J. Specific decrease of secre-
tin/VIP-stimulated adenylate cyclase in the heart from the Lyon
strain of hypertensive rats. Peptides 5: 355–358, 1984.
419. Roberts LA, Slocum GR, Riley DA. Morphological study of the
innervation pattern of the rabbit sinoatrial node. Am J Anat 185:
74 88, 1989.
420. Robinson RB, Baruscotti M, DiFrancesco D. Autonomic modu-
lation of heart rate: pitfalls of nonselective channel blockade. Am J
Physiol Heart Circ Physiol 285: H2865, 2003.
421. Robinson RB, Brink PR, Cohen IS, Rosen MR. I
f
and the bio-
logical pacemaker. Pharmacol Res 53: 407–415, 2006.
422. Roden DM, Balser JR, George AL Jr, Anderson ME. Cardiac
ion channels. Annu Rev Physiol 64: 431–475, 2002.
423. Rosen MR, Brink PR, Cohen IS, Robinson RB. Genes, stem cells
and biological pacemakers. Cardiovasc Res 64: 12–23, 2004.
424. Rosen MR, Robinson RB, Brink P, Cohen IS. Recreating the
biological pacemaker. Anat Rec A Discov Mol Cell Evol Biol 280:
1046–1052, 2004.
425. Rottbauer W, Baker K, Wo ZG, Mohideen MA, Cantiello HF,
Fishman MC. Growth and function of the embryonic heart depend
upon the cardiac-specific L-type calcium channel alpha1 subunit.
Dev Cell 1: 265–275, 2001.
426. Rozanski GJ, Lipsius SL. Electrophysiology of functional subsid-
iary pacemakers in canine right atrium. Am J Physiol Heart Circ
Physiol 249: H594–H603, 1985.
427. Rozanski GJ, Lipsius SL, Randall WC, Jones SB. Alterations in
subsidiary pacemaker function after prolonged subsidiary pace-
maker dominance in the canine right atrium. J Am Coll Cardiol 4:
535–542, 1984.
428. Rubenstein DS, Lipsius SL. Mechanisms of automaticity in sub-
sidiary pacemakers from cat right atrium. Circ Res 64: 648 657,
1989.
429. Saikawa T, Carmeliet E. Slow recovery of the maximal rate of
rise (V
max
) of the action potential in sheep cardiac Purkinje fibers.
Pflu¨ gers Arch 394: 90–93, 1982.
430. Saito K, Gutkind JS, Saavedra JM. Angiotensin II binding sites
in the conduction system of rat hearts. Am J Physiol Heart Circ
Physiol 253: H1618–H1622, 1987.
431. Sakai R, Hagiwara N, Matsuda N, Kassanuki H, Hosoda S.
Sodium-potassium pump current in rabbit sino-atrial node cells.
J Physiol 490: 51–62, 1996.
432. Sakmann B, Noma A, Trautwein W. Acetylcholine activation of
single muscarinic K
channels in isolated pacemaker cells of the
mammalian heart. Nature 303: 250–253, 1983.
433. Sanders L, Rakovic S, Lowe M, Mattick PA, Terrar DA. Fun-
damental importance of Na
-Ca
2
exchange for the pacemaking
mechanism in guinea-pig sino-atrial node. J Physiol 571: 639 649,
2006.
434. Sanders P, Kistler PM, Morton JB, Spence SJ, Kalman JM.
Remodeling of sinus node function in patients with congestive
heart failure: reduction in sinus node reserve. Circulation 110:
897–903, 2004.
435. Santoro B, Liu DT, Yao H, Bartsch D, Kandel ER, Siegelbaum
SA, Tibbs GR. Identification of a gene encoding a hyperpolariza-
tion-activated pacemaker channel of brain. Cell 93: 717–729, 1998.
436. Sarachek NS, Leonard JL. Familial heart block and sinus brady-
cardia. Classification and natural history. Am J Cardiol 29: 451–
458, 1972.
437. Sasaki S, Daitoku K, Iwasa A, Motomura S. NO is involved in
MCh-induced accentuated antagonism via type II PDE in the canine
blood-perfused SA node. Am J Physiol Heart Circ Physiol 279:
H2509–H2518, 2000.
438. Satin J, Cribbs LL. Identification of a T-type Ca
2
channel iso-
form in murine atrial myocytes (AT-1 cells). Circ Res 86: 636 642,
2000.
439. Satoh H. Role of T-type Ca
2
channel inhibitors in the pacemaker
depolarization in rabbit sino-atrial nodal cells. Gen Pharmacol 26:
581–587, 1995.
440. Schimerlik MI. Structure and regulation of muscarinic receptors.
Annu Rev Physiol 51: 217–227, 1989.
441. Schram G, Pourrier M, Melnyk P, Nattel S. Differential distri-
bution of cardiac ion channel expression as a basis for regional
specialization in electrical function. Circ Res 90: 939–950, 2002.
442. Schuessler RB, Boineau JP, Bromberg BI. Origin of the sinus
impulse. J Cardiovasc Electrophysiol 7: 263–274, 1996.
443. Schulze-Bahr E, Neu A, Friederich P, Kaupp UB, Breithardt
G, Pongs O, Isbrandt D. Pacemaker channel dysfunction in a
patient with sinus node disease. J Clin Invest 111: 1537–1545, 2003.
444. Seifert R, Scholten A, Gauss R, Mincheva A, Lichter P, Kaupp
UB. Molecular characterization of a slowly gating human hyperpo-
larization-activated channel predominantly expressed in thalamus,
heart, testis. Proc Natl Acad Sci USA 96: 9391–9396, 1999.
445. Seisenberger C, Specht V, Welling A, Platzer J, Pfeifer A,
Kuhbandner S, Striessnig J, Klugbauer N, Feil R, Hofmann F.
Functional embryonic cardiomyocytes after disruption of the L-
type alpha1C (Cav1.2) calcium channel gene in the mouse. J Biol
Chem 275: 39193–39199, 2000.
446. Senges J, Mizutani T, Pelzer D, Brachmann J, Sonnhof U,
Kubler W. Effect of hypoxia on the sinoatrial node, atrium, atrio-
ventricular node in the rabbit heart. Circ Res 44: 856 863, 1979.
447. Severi S, Cavalcanti S. Electrolyte and pH dependence of heart
rate during hemodialysis: a computer model analysis. Artif Organs
24: 245–260, 2000.
448. Severi S, Cavalcanti S, Mancini E, Santoro A. Effect of elec-
trolyte and pH changes on the sinus node pacemaking in humans.
J Electrocardiol 35: 115–124, 2002.
449. Seyama I. Characteristics of the anion channel in the sino-atrial
node cell of the rabbit. J Physiol 294: 447–460, 1979.
450. Shi W, Wymore R, Yu H, Wu J, Wymore RT, Pan Z, Robinson
RB, Dixon JE, McKinnon D, Cohen IS. Distribution and preva-
lence of hyperpolarization-activated cation channel (HCN) mRNA
expression in cardiac tissues. Circ Res 85: e1–6, 1999.
451. Shibasaki T. Conductance and kinetics of delayed rectifier potas-
sium channels in nodal cells of the rabbit heart. J Physiol 387:
227–250, 1987.
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
979
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
452. Shibasaki T. Conductance and kinetics of delayed rectifier potas-
sium channels in nodal cells of the rabbit heart. J Physiol 387:
227–250, 1987.
453. Shibata EF, Giles W, Pollack GH. Threshold effects of acetyl-
choline on primary pacemaker cells of the rabbit sino-atrial node.
Proc R Soc Lond B Biol Sci 223: 355–378, 1985.
454. Shigekawa M, Iwamoto T. Cardiac Na
-Ca
2
exchange: molecu-
lar and pharmacological aspects. Circ Res 88: 864 876, 2001.
455. Shimoni Y, Fiset C, Clark RB, Dixon JE, McKinnon D, Giles
WR. Thyroid hormone regulates postnatal expression of transient
K
channel isoforms in rat ventricle. J Physiol 500: 65–73, 1997.
456. Shimoni Y, Han X, Severson D, Giles WR. Mediation by nitric
oxide of the indirect effects of adenosine on calcium current in
rabbit heart pacemaker cells. Br J Pharmacol 119: 1463–1469, 1996.
457. Shinagawa Y, Satoh H, Noma A. The sustained inward current
and inward rectifier K
current in pacemaker cells dissociated
from rat sinoatrial node. J Physiol 523: 593–605, 2000.
458. Shvilkin A, Danilo P Jr, Chevalier P, Chang F, Cohen IS,
Rosen MR. Vagal release of vasoactive intestinal peptide can
promote vagotonic tachycardia in the isolated innervated rat heart.
Cardiovasc Res 28: 1769–1773, 1994.
459. Silva J, Rudy Y. Mechanism of pacemaking in I
K1
-downregulated
myocytes. Circ Res 92: 261–263, 2003.
460. Silverman ME, Grove D, Upshaw CB Jr. Why does the heart
beat? The discovery of the electrical system of the heart. Circula-
tion 113: 2775–2781, 2006.
461. Simon L, Ghaleh B, Puybasset L, Giudicelli JF, Berdeaux A.
Coronary and hemodynamic effects of S 16257, a new bradycardic
agent, in resting and exercising conscious dogs. J Pharmacol Exp
Ther 275: 659 666, 1995.
462. Sinnegger-Brauns MJ, Hetzenauer A, Huber IG, Renstrom E,
Wietzorrek G, Berjukov S, Cavalli M, Walter D, Koschak A,
Waldschutz R, Hering S, Bova S, Rorsman P, Pongs O,
Singewald N, Striessnig JJ. Isoform-specific regulation of mood
behavior and pancreatic beta cell and cardiovascular function by
L-type Ca
2
channels. J Clin Invest 113: 1430–1439, 2004.
463. Slovut DP, Wenstrom JC, Moeckel RB, Wilson RF, Osborn
JW, Abrams JH. Respiratory sinus dysrhythmia persists in trans-
planted human hearts following autonomic blockade. Clin Exp
Pharmacol Physiol 25: 322–330, 1998.
464. Sorota S, Du XY. Delayed activation of cardiac swelling-induced
chloride current after step changes in cell size. J Cardiovasc Elec-
trophysiol 9: 825–831, 1998.
465. Soufan AT, Ruijter JM, van den Hoff MJ, de Boer PA, Hagoort
J, Moorman AF. Three-dimensional reconstruction of gene ex-
pression patterns during cardiac development. Physiol Genomics
13: 187–195, 2003.
466. Stieber J, Herrmann S, Feil S, Loster J, Feil R, Biel M,
Hofmann F, Ludwig A. The hyperpolarization-activated channel
HCN4 is required for the generation of pacemaker action potentials
in the embryonic heart. Proc Natl Acad Sci USA 100: 15235–15240,
2003.
467. Stieber J, Hofmann F, Ludwig A. Pacemaker channels and sinus
node arrhythmia. Trends Cardiovasc Med 14: 23–28, 2004.
468. Stjarne L, Lundberg JM. On the possible roles of noradrenaline,
adenosine 5-triphosphate and neuropeptide Y as sympathetic co-
transmitters in the mouse vas deferens. Prog Brain Res 68: 263–
278, 1986.
469. Stowe DF, Bosnjak ZJ, Kampine JP. Effects of hypoxia on adult
and neonatal pacemaker rates. Obstet Gynecol 66: 649 656, 1985.
470. Striessnig J. Pharmacology, structure and function of cardiac
L-type Ca
2
channels. Cell Physiol Biochem 9: 242–269, 1999.
471. Takano M, Noma A. Distribution of the isoprenaline-induced
chloride current in rabbit heart. Pflu¨ gers Arch 420: 223–226, 1992.
472. Takimoto K, Li D, Nerbonne JM, Levitan ES. Distribution,
splicing and glucocorticoid-induced expression of cardiac alpha 1C
and alpha 1D voltage-gated Ca
2
channel mRNAs. J Mol Cell Car-
diol 29: 3035–3042, 1997.
473. Tamargo J, Caballero R, Gomez R, Valenzuela C, Delpon E.
Pharmacology of cardiac potassium channels. Cardiovasc Res 62:
9–33, 2004.
474. Tanabe T, Beam KG, Powell JA, Numa S. Restoration of exci-
tation-contraction coupling and slow calcium current in dysgenic
muscle by dihydropyridine receptor complementary DNA. Nature
336: 134–139, 1988.
475. Tanabe T, Mikami A, Numa S, Beam KG. Cardiac-type excita-
tion-contraction coupling in dysgenic skeletal muscle injected with
cardiac dihydropyridine receptor cDNA. Nature 344: 451– 453,
1990.
476. Tawara S. Das Reizleitungssystem des Saiigethierherzens. Jena:
Gustav Fischer, 1906.
477. Tellez JO, Dobrzynski H, Greener ID, Graham GM, Laing E,
Honjo H, Hubbard SJ, Boyett MR, Billeter R. Differential ex-
pression of ion channel transcripts in atrial muscle and sinoatrial
node in rabbit. Circ Res 99: 1384–1393, 2006.
478. Temple J, Frias P, Rottman J, Yang T, Wu Y, Verheijck EE,
Zhang W, Siprachanh C, Kanki H, Atkinson JB, King P, Ander-
son ME, Kupershmidt S, Roden DM. Atrial fibrillation in KCNE1-
null mice. Circ Res 97: 62–69, 2005.
479. Thollon C, Bidouard JP, Cambarrat C, Lesage L, Reure H,
Delescluse I, Vian J, Peglion JL, Vilaine JP. Stereospecific in
vitro and in vivo effects of the new sinus node inhibitor ()-S
16257. Eur J Pharmacol 339: 43–51, 1997.
480. 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: compar-
ison with UL-FS 49. Br J Pharmacol 112: 37–42, 1994.
481. Toda N, West TC. The influence of ouabain on cholinergic re-
sponses in the sinoatrial node. J Pharmacol Exp Ther 153: 104–113,
1966.
482. Toyama J, Boyett MR, Watanabe E, Honjo H, Anno T, Kodama
I. Computer simulation of the electrotonic modulation of pace-
maker activity in the sinoatrial node by atrial muscle. J Electrocar-
diol 28 Suppl: 212–215, 1995.
483. Tse HF, Xue T, Lau CP, Siu CW, Wang K, Zhang QY, Tomaselli
GF, Akar FG, Li RA. 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 114: 1000–1011, 2006.
484. Tseng GN, Boyden PA. Multiple types of Ca
2
currents in single
canine Purkinje cells. Circ Res 65: 1735–1750, 1989.
485. Tsien RW, Carpenter DO. Ionic mechanisms of pacemaker ac-
tivity in cardiac Purkinje fibers. Federation Proc 37: 2127–2131,
1978.
486. Ueda K, Nakamura K, Hayashi T, Inagaki N, Takahashi M,
Arimura T, Morita H, Higashiuesato Y, Hirano Y, Yasunami M,
Takishita S, Yamashina A, Ohe T, Sunamori M, Hiraoka
M, Kimura A. Functional characterization of a trafficking-defec-
tive HCN4 mutation, D553N, associated with cardiac arrhythmia.
J Biol Chem 279: 27194–27198, 2004.
487. Valenzuela D, Han X, Mende U, Fankhauser C, Mashimo H,
Huang P, Pfeffer J, Neer EJ, Fishman MC. G alpha(o) is nec-
essary for muscarinic regulation of Ca
2
channels in mouse heart.
Proc Natl Acad Sci USA 94: 1727–1732, 1997.
488. Van Bogaert PP, Goethals M. Pharmacological influence of spe-
cific bradycardic agents on the pacemaker current of sheep cardiac
Purkinje fibers. A comparison between three different molecules.
Eur Heart J 8 Suppl L: 35–42, 1987.
489. Van Bogaert PP, Goethals M, Simoens C. Use- and frequency-
dependent blockade by UL-FS 49 of the if pacemaker current in
sheep cardiac Purkinje fibers. Eur J Pharmacol 187: 241–256, 1990.
490. Van Bogaert PP, Pittoors F. Use-dependent blockade of cardiac
pacemaker current (I
f
) by cilobradine and zatebradine. Eur J Phar-
macol 478: 161–171, 2003.
491. Van der Heyden MA, Wijnhoven TJ, Opthof T. Molecular as-
pects of adrenergic modulation of cardiac L-type Ca
2
channels.
Cardiovasc Res 65: 28–39, 2005.
492. Van Ginneken AC, Giles W. Voltage clamp measurements of the
hyperpolarization-activated inward current I
f
in single cells from
rabbit sino-atrial node. J Physiol 434: 57–83, 1991.
493. Van Wagoner DR. Mechanosensitive gating of atrial ATP-sensitive
potassium channels. Circ Res 72: 973–983, 1993.
494. Vandecasteele G, Eschenhagen T, Scholz H, Stein B, Verde I,
Fischmeister R. Muscarinic and beta-adrenergic regulation of
heart rate, force of contraction and calcium current is preserved in
980 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
mice lacking endothelial nitric oxide synthase. Nat Med 5: 331–334,
1999.
495. Vassalle M. Analysis of cardiac pacemaker potential using a “volt-
age clamp” technique. Am J Physiol 210: 1335–1341, 1966.
496. Verheijck EE, van Ginneken AC, Bourier J, Bouman LN. Ef-
fects of delayed rectifier current blockade by E-4031 on impulse
generation in single sinoatrial nodal myocytes of the rabbit. Circ
Res 76: 607–615, 1995.
497. Verheijck EE, van Ginneken AC, Wilders R, Bouman LN.
Contribution of L-type Ca
2
current to electrical activity in sino-
atrial nodal myocytes of rabbits. Am J Physiol Heart Circ Physiol
276: H1064–1077, 1999.
498. Verheijck EE, van Kempen MJ, Veereschild M, Lurvink J,
Jongsma HJ, Bouman LN. Electrophysiological features of the
mouse sinoatrial node in relation to connexin distribution. Cardio-
vasc Res 52: 40–50, 2001.
499. Verheijck EE, Wessels A, van Ginneken AC, Bourier J, Mark-
man MW, Vermeulen JL, de Bakker JM, Lamers WH, Opthof
T, Bouman LN. Distribution of atrial and nodal cells within the
rabbit sinoatrial node: models of sinoatrial transition. Circulation
97: 1623–1631, 1998.
500. Verheijck EE, Wilders R, Bouman LN. Atrio-sinus interaction
demonstrated by blockade of the rapid delayed rectifier current.
Circulation 105: 880 885, 2002.
501. Verkerk AO, Veldkamp MW, Abbate F, Antoons G, Bouman
LN, Ravesloot JH, van Ginneken AC. Two types of action po-
tential configuration in single cardiac Purkinje cells of sheep. Am J
Physiol Heart Circ Physiol 277: H1299–H1310, 1999.
502. Verkerk AO, Veldkamp MW, Bouman LN, van Ginneken AC.
Calcium-activated Cl
current contributes to delayed afterdepolar-
izations in single Purkinje and ventricular myocytes. Circulation
101: 2639–2644, 2000.
503. Verkerk AO, Wilders R, Coronel R, Ravesloot JH, Verheijck
EE. Ionic remodeling of sinoatrial node cells by heart failure.
Circulation 108: 760–766, 2003.
504. Verkerk AO, Wilders R, Zegers JG, van Borren MM, Ravesloot
JH, Verheijck EE. Ca
2
-activated Cl
current in rabbit sinoatrial
node cells. J Physiol 540: 105–117, 2002.
505. Viatchenko-Karpinski S, Fleischmann BK, Liu Q, Sauer H,
Gryshchenko O, Ji GJ, Hescheler J. Intracellular Ca
2
oscilla-
tions drive spontaneous contractions in cardiomyocytes during
early development. Proc Natl Acad Sci USA 96: 8259 8264, 1999.
506. Vinogradova TM, Bogdanov KY, Lakatta EG. beta-Adrenergic
stimulation modulates ryanodine receptor Ca
2
release during di-
astolic depolarization to accelerate pacemaker activity in rabbit
sinoatrial nodal cells. Circ Res 90: 73–79, 2002.
507. Vinogradova TM, Fedorov VV, Yuzyuk TN, Zaitsev AV, Rosen-
shtraukh LV. Local cholinergic suppression of pacemaker activity
in the rabbit sinoatrial node. J Cardiovasc Pharmacol 32: 413– 424,
1998.
508. Vinogradova TM, Lyashkov AE, Zhu W, Ruknudin AM, Si-
renko S, Yang D, Deo S, Barlow M, Johnson S, Caffrey JL,
Zhou YY, Xiao RP, Cheng H, Stern MD, Maltsev VA, Lakatta
EG. High basal protein kinase A-dependent phosphorylation drives
rhythmic internal Ca
2
store oscillations and spontaneous beating
of cardiac pacemaker cells. Circ Res 98: 505–514, 2006.
509. Vinogradova TM, Zhou YY, Bogdanov KY, Yang D, Kuschel M,
Cheng H, Xiao RP. Sinoatrial node pacemaker activity requires
Ca
2
/calmodulin-dependent protein kinase II activation. Circ Res
87: 760–767, 2000.
510. Vinogradova TM, Zhou YY, Maltsev V, Lyashkov A, Stern M,
Lakatta EG. Rhythmic ryanodine receptor Ca
2
releases during
diastolic depolarization of sinoatrial pacemaker cells do not re-
quire membrane depolarization. Circ Res 94: 802–809, 2004.
511. Viswanathan PC, Coles JA Jr, Sharma V, Sigg DC. Recreating
an artificial biological pacemaker: insights from a theoretical
model. Heart Rhythm 3: 824 831, 2006.
512. Waltuck J, Buyon JP. Autoantibody-associated congenital heart
block: outcome in mothers and children. Ann Intern Med 120:
544–551, 1994.
513. Warren KS, Baker K, Fishman MC. The slow mo mutation
reduces pacemaker current and heart rate in adult zebrafish. Am J
Physiol Heart Circ Physiol 281: H1711–H1719, 2001.
514. Watanabe EI, Honjo H, Anno T, Boyett MR, Kodama I,
Toyama J. Modulation of pacemaker activity of sinoatrial node
cells by electrical load imposed by an atrial cell model. Am J
Physiol Heart Circ Physiol 269: H1735–H1742, 1995.
515. Watanabe Y, Dreifus LS. Sites of impulse formation within the
atrioventricular junction of the rabbit. Circ Res 22: 717–727, 1968.
516. Weiergraber M, Henry M, Sudkamp M, de Vivie ER, Hescheler
J, Schneider T. Ablation of Ca
v
2.3 / E-type voltage-gated calcium
channel results in cardiac arrhythmia and altered autonomic con-
trol within the murine cardiovascular system. Basic Res Cardiol
100: 1–13, 2005.
517. West GA, Belardinelli L. Correlation of sinus slowing and hyper-
polarization caused by adenosine in sinus node. Pflu¨ gers Arch 403:
75–81, 1985.
518. West GA, Belardinelli L. Sinus slowing and pacemaker shift
caused by adenosine in rabbit SA node. Pflu¨ gers Arch 403: 66–74,
1985.
519. West TC, Toda N. Response of the A-V node of the rabbit to
stimulation of intracardiac cholinergic nerves. Circ Res 20: 18 –31,
1967.
520. Wickman K, Clapham DE. Ion channel regulation by G proteins.
Physiol Rev 75: 865–885, 1995.
521. Wickman K, Krapivinsky G, Corey S, Kennedy M, Nemec J,
Medina I, Clapham DE. Structure, G protein activation, func-
tional relevance of the cardiac G protein-gated K
channel, I
KACh
.
Ann NY Acad Sci 868: 386–398, 1999.
522. Wickman K, Nemec J, Gendler SJ, Clapham DE. Abnormal
heart rate regulation in GIRK4 knockout mice. Neuron 20: 103–114,
1998.
523. Wickman KD, Iniguez-Lluhl JA, Davenport PA, Taussig R,
Krapivinsky GB, Linder ME, Gilman AG, Clapham DE. Recom-
binant G-protein beta gamma-subunits activate the muscarinic-
gated atrial potassium channel. Nature 368: 255–257, 1994.
524. Wikstrom L, Johansson C, Salto C, Barlow C, Campos Barros
A, Baas F, Forrest D, Thoren P, Vennstrom B. Abnormal heart
rate and body temperature in mice lacking thyroid hormone recep-
tor alpha 1. EMBO J 17: 455–461, 1998.
525. Wilders R. Computer modelling of the sinoatrial node. Med Biol
Eng Comput 45: 189–207, 2007.
526. Wilders R, Jongsma HJ, van Ginneken AC. Pacemaker activity
of the rabbit sinoatrial node. A comparison of mathematical mod-
els. Biophys J 60: 1202–1216, 1991.
527. Wilson SJ and Bolter CP. Do cardiac neurons play a role in the
intrinsic control of heart rate in the rat? Exp Physiol 87: 675–682,
2002.
528. Xiao RP, Cheng H, Lederer WJ, Suzuki T, Lakatta EG. Dual
regulation of Ca
2
/calmodulin-dependent kinase II activity by
membrane voltage and by calcium influx. Proc Natl Acad Sci USA
91: 9659–9663, 1994.
529. Xu M, Welling A, Paparisto S, Hofmann F, Klugbauer N. En-
hanced expression of L-type Cav1.3 calcium channels in murine
embryonic hearts from Cav1 2-deficient mice. J Biol Chem 278:
40837–40841, 2003.
530. Yamada M. The role of muscarinic K
channels in the negative
chronotropic effect of a muscarinic agonist. J Pharmacol Exp Ther
300: 681–687, 2002.
531. Yamamoto M, Dobrzynski H, Tellez J, Niwa R, Billeter R,
Honjo H, Kodama I, Boyett MR. Extended atrial conduction
system characterised by the expression of the HCN4 channel and
connexin45. Cardiovasc Res 72: 271–281, 2006.
532. Yamane T, Shah DC, Jais P, Haissaguerre M. Pseudo sinus
rhythm originating from the left superior pulmonary vein in a
patient with paroxysmal atrial fibrillation. J Cardiovasc Electro-
physiol 12: 1190–1191, 2001.
533. Yanagihara K, Noma A, Irisawa H. Reconstruction of sino-atrial
node pacemaker potential based on the voltage clamp experiments.
Jpn J Physiol 30: 841–857, 1980.
534. Yu FH, Catterall WA. Overview of the voltage-gated sodium
channel family. Genome Biol 4: 207, 2003.
535. Yu H, Chang F, Cohen IS. Pacemaker current exists in ventricular
myocytes. Circ Res 72: 232–236, 1993.
536. Yu H, Wu J, Potapova I, Wymore RT, Holmes B, Zuckerman J,
Pan Z, Wang H, Shi W, Robinson RB, El-Maghrabi MR, Ben-
GENESIS AND REGULATION OF THE HEART AUTOMATICITY
981
Physiol Rev VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
jamin W, Dixon J, McKinnon D, Cohen IS, Wymore R. MinK-
related peptide 1: a beta subunit for the HCN ion channel subunit
family enhances expression and speeds activation. Circ Res 88:
E84 87, 2001.
537. Yu X, Chen XW, Zhou P, Yao L, Liu T, Zhang B, Li Y, Zheng H,
Zheng LH, Zhang CX, Bruce I, Ge JB, Wang SQ, Hu ZA, Yu HG,
Zhou Z. Calcium influx through I
f
channels in rat ventricular
myocytes. Am J Physiol Cell Physiol 292: C1147–C1155, 2007.
538. Yu X, Duan KL, Shang CF, Yu HG, Zhou Z. Calcium influx
through hyperpolarization-activated cation channels [I
h
channels]
contributes to activity-evoked neuronal secretion. Proc Natl Acad
Sci USA 101: 1051–1056, 2004.
539. Yuill KH, Hancox JC. Characteristics of single cells isolated from
the atrioventricular node of the adult guinea-pig heart. Pflu¨ gers
Arch 445: 311–320, 2002.
540. Zaza A, Maccaferri G, Mangoni M, DiFrancesco D. Intracellular
calcium does not directly modulate cardiac pacemaker (i
f
) chan-
nels. Pflu¨ gers Arch 419: 662–664, 1991.
541. Zaza A, Micheletti M, Brioschi A, Rocchetti M. Ionic currents
during sustained pacemaker activity in rabbit sino-atrial myocytes.
J Physiol 505: 677–688, 1997.
542. Zaza A, Robinson RB, DiFrancesco D. Basal responses of the
L-type Ca
2
and hyperpolarization-activated currents to auto-
nomic agonists in the rabbit sino-atrial node. J Physiol 491:
347–355, 1996.
543. Zaza A, Rocchetti M, DiFrancesco D. Modulation of the hyper-
polarization-activated current [I
f
] by adenosine in rabbit sinoatrial
myocytes. Circulation 94: 734–741, 1996.
544. Zhang H, Holden AV, Boyett MR. Sustained inward current and
pacemaker activity of mammalian sinoatrial node. J Cardiovasc
Electrophysiol 13: 809 812, 2002.
545. Zhang H, Holden AV, Kodama I, Honjo H, Lei M, Varghese T,
Boyett MR. Mathematical models of action potentials in the pe-
riphery and center of the rabbit sinoatrial node. Am J Physiol
Heart Circ Physiol 279: H397–H421, 2000.
546. Zhang Z, Xu Y, Song H, Rodriguez J, Tuteja D, Namkung Y,
Shin HS, Chiamvimonvat N. Functional roles of Ca(v)1.3
[alpha(1D)] calcium channel in sinoatrial nodes: insight gained
using gene-targeted null mutant mice. Circ Res 90: 981–987., 2002.
547. Zhou Z, Lipsius SL. Na
-Ca
2
exchange current in latent pace-
maker cells isolated from cat right atrium. J Physiol 466: 263–285,
1993.
548. Zicha S, Fernandez-Velasco M, Lonardo G, L’Heureux N, Nat-
tel S. Sinus node dysfunction and hyperpolarization-activated
(HCN) channel subunit remodeling in a canine heart failure model.
Cardiovasc Res 66: 472–481, 2005.
982 MATTEO E. MANGONI AND JOE
¨
L NARGEOT
Physiol Rev
VOL 88 JULY 2008 www.prv.org
on July 15, 2008 physrev.physiology.orgDownloaded from
... Heart automaticity is a fundamental physiological function that is reliant on the presence of a highly specialized population of cardiomyocytes in the sinoatrial node (SAN). These cells are referred to as pacemaker cells [1]. The spontaneous activity of pacemaker cells is due to diastolic depolarization (DD), a slow depolarization phase that drives the membrane potential from the end of an action potential (AP) to the threshold of a new AP. ...
... The spontaneous activity of pacemaker cells is due to diastolic depolarization (DD), a slow depolarization phase that drives the membrane potential from the end of an action potential (AP) to the threshold of a new AP. SAN cells express a wide array of ion channels, which underlie the generation and regulation of DD by the autonomic nervous system (ANS) [1]. Knowledge of the functional role of ion channels in pacemaker activity mostly comes from small-animal models (i.e., rodents) [2][3][4]. ...
... Along with markers of cardiac development, adult SAN is characterized by a specific pattern of ion channels that ensures the generation of spontaneous activity ( Figure 2). The generation of automaticity in cardiac pacemaker cells is due to DD, a spontaneous, slowly depolarizing phase of the AP cycle [1]. During this phase, the membrane potential progressively becomes less negative until it reaches the threshold for triggering a new AP. ...
Article
Full-text available
Human-induced pluripotent stem cell (hiPSC)-derived cardiomyocytes raise the possibility of generating pluripotent stem cells from a wide range of human diseases. In the cardiology field, hiPSCs have been used to address the mechanistic bases of primary arrhythmias and in investigations of drug safety. These studies have been focused primarily on atrial and ventricular pathologies. Consequently, many hiPSC-based cardiac differentiation protocols have been developed to differentiate between atrial- or ventricular-like cardiomyocytes. Few protocols have successfully proposed ways to obtain hiPSC-derived cardiac pacemaker cells, despite the very limited availability of human tissues from the sinoatrial node. Providing an in vitro source of pacemaker-like cells would be of paramount importance in terms of furthering our understanding of the mechanisms underlying sinoatrial node pathophysiology and testing innovative clinical strategies against sinoatrial node dysfunction (i.e., biological pacemakers and genetic- and pharmacological- based therapy). Here, we summarize and detail the currently available protocols used to obtain patient-derived pacemaker-like cells.
... This result is in line with HCN2 and HCN4 channel differences in function, distribution, and regulation. HCN4 channels are mostly responsible for the I f current of the sinoatrial node, the pacemaker region of the heart, generating intrinsic heartbeat (see [82] for a review) and regulating oscillation activity in thalamocortical neurons. Substantial differences in cellular localization have been observed for HCN2 and HCN4. ...
Article
Full-text available
Pacemaking activity in substantia nigra dopaminergic neurons is generated by the coordinated activity of a variety of distinct somatodendritic voltage- and calcium-gated ion channels. We investigated whether these functional interactions could arise from a common localization in macromolecular complexes where physical proximity would allow for efficient interaction and co-regulations. For that purpose, we immunopurified six ion channel proteins involved in substantia nigra neuron autonomous firing to identify their molecular interactions. The ion channels chosen as bait were Cav1.2, Cav1.3, HCN2, HCN4, Kv4.3, and SK3 channel proteins, and the methods chosen to determine interactions were co-immunoprecipitation analyzed through immunoblot and mass spectrometry as well as proximity ligation assay. A macromolecular complex composed of Cav1.3, HCN, and SK3 channels was unraveled. In addition, novel potential interactions between SK3 channels and sclerosis tuberous complex (Tsc) proteins, inhibitors of mTOR, and between HCN4 channels and the pro-degenerative protein Sarm1 were uncovered. In order to demonstrate the presence of these molecular interactions in situ, we used proximity ligation assay (PLA) imaging on midbrain slices containing the substantia nigra, and we could ascertain the presence of these protein complexes specifically in substantia nigra dopaminergic neurons. Based on the complementary functional role of the ion channels in the macromolecular complex identified, these results suggest that such tight interactions could partly underly the robustness of pacemaking in dopaminergic neurons.
... I f is proposed to initiate diastolic depolarization. [9][10][11] T-type calcium channels are low-voltage activated and are also open at voltages negative to − 40 mV. [6] They are suggested to be responsible for late phase diastolic depolarization. ...
Article
Full-text available
Background The major membrane currents responsible for sinoatrial (SA) rhythm generation are generally studied in isolated cardiac cells using electrophysiological tools. Such studies are resource and labor-intensive. Materials and Methods Here, we have studied four major currents in isolated rat heart preparations, perfused in Langendorff mode, and demonstrate that this is a good preparation for such studies. Heart rates of isolated perfused rat hearts were recorded using surface electrocardiogram before and after perfusion with drugs and solutions that affect the four major currents responsible for SA rhythm generation. Results The rates of whole isolated hearts beating with SA rhythm decreased with cesium and decreased by about half with ivabradine, both blockers of the funny current (I f ). Importantly, the rhythm was not abolished even with a high dose of ivabradine at which total blockade of I f is expected. The rate was not affected by nickel, a blocker of T-type calcium current. The SA rhythm was abolished by the reduction or removal of sodium from the perfusate (interventions that inhibit the calcium-extrusive mode of the sodium-calcium exchanger) or by nifedipine, the L-type calcium channel blocker. Discussion The inferences made based on these observations are (a) I f contributes significantly to pacemaking, (b) I CaT does not play a role and (c) I NCX and I CaL are obligatory rhythm-generating currents in the SA node. Cyclical calcium release from SR during diastole (the calcium clock), responsible for driving I NCX in its forward mode is probably a phenomenon independent of membrane events, as total I f blockade did not abolish rhythm generation. These results corroborate with published literature where most studies were done on single cells.
Article
Full-text available
The sinus node (SN) serves as the primary pacemaker of the heart and is the first component of the cardiac conduction system. Due to its anatomical properties and sample scarcity, the cellular composition of the human SN has been historically challenging to study. Here, we employed a novel deep learning deconvolution method, namely Bulk2space, to characterise the cellular heterogeneity of the human SN using existing single-cell datasets of non-human species. As a proof of principle, we used Bulk2Space to profile the cells of the bulk human right atrium using publicly available mouse scRNA-Seq data as a reference. 18 human cell populations were identified, with cardiac myocytes being the most abundant. Each identified cell population correlated to its published experimental counterpart. Subsequently, we applied the deconvolution to the bulk transcriptome of the human SN and identified 11 cell populations, including a population of pacemaker cardiomyocytes expressing pacemaking ion channels (HCN1, HCN4, CACNA1D) and transcription factors (SHOX2 and TBX3). The connective tissue of the SN was characterised by adipocyte and fibroblast populations, as well as key immune cells. Our work unravelled the unique single cell composition of the human SN by leveraging the power of a novel machine learning method.
Article
Full-text available
This Committee Report provides methodological, interpretive, and reporting guidance for researchers who use measures of heart rate (HR) and heart rate variability (HRV) in psychophysiological research. We provide brief summaries of best practices in measuring HR and HRV via electrocardiographic and photoplethysmographic signals in laboratory, field (ambulatory), and brain‐imaging contexts to address research questions incorporating measures of HR and HRV. The Report emphasizes evidence for the strengths and weaknesses of different recording and derivation methods for measures of HR and HRV. Along with this guidance, the Report reviews what is known about the origin of the heartbeat and its neural control, including factors that produce and influence HRV metrics. The Report concludes with checklists to guide authors in study design and analysis considerations, as well as guidance on the reporting of key methodological details and characteristics of the samples under study. It is expected that rigorous and transparent recording and reporting of HR and HRV measures will strengthen inferences across the many applications of these metrics in psychophysiology. The prior Committee Reports on HR and HRV are several decades old. Since their appearance, technologies for human cardiac and vascular monitoring in laboratory and daily life (i.e., ambulatory) contexts have greatly expanded. This Committee Report was prepared for the Society for Psychophysiological Research to provide updated methodological and interpretive guidance, as well as to summarize best practices for reporting HR and HRV studies in humans.
Article
Full-text available
Loss or dysregulation of the normally precise control of heart rate via the autonomic nervous system plays a critical role during the development and progression of cardiovascular disease—including ischemic heart disease, heart failure, and arrhythmias. While the clinical significance of regulating changes in heart rate, known as the chronotropic effect, is undeniable, the mechanisms controlling these changes remain not fully understood. Heart rate acceleration and deceleration are mediated by increasing or decreasing the spontaneous firing rate of pacemaker cells in the sinoatrial node. During the transition from rest to activity, sympathetic neurons stimulate these cells by activating β-adrenergic receptors and increasing intracellular cyclic adenosine monophosphate. The same signal transduction pathway is targeted by positive chronotropic drugs such as norepinephrine and dobutamine, which are used in the treatment of cardiogenic shock and severe heart failure. The cyclic adenosine monophosphate–sensitive hyperpolarization–activated current (I f ) in pacemaker cells is passed by hyperpolarization-activated cyclic nucleotide-gated cation channels and is critical for generating the autonomous heartbeat. In addition, this current has been suggested to play a central role in the chronotropic effect. Recent studies demonstrate that cyclic adenosine monophosphate–dependent regulation of HCN4 (hyperpolarization-activated cyclic nucleotide-gated cation channel isoform 4) acts to stabilize the heart rate, particularly during rapid rate transitions induced by the autonomic nervous system. The mechanism is based on creating a balance between firing and recently discovered nonfiring pacemaker cells in the sinoatrial node. In this way, hyperpolarization-activated cyclic nucleotide-gated cation channels may protect the heart from sinoatrial node dysfunction, secondary arrhythmia of the atria, and potentially fatal tachyarrhythmia of the ventricles. Here, we review the latest findings on sinoatrial node automaticity and discuss the physiological and pathophysiological role of HCN pacemaker channels in the chronotropic response and beyond.
Article
Full-text available
Cardiac rhythm regulated by micro-macroscopic structures of heart. Pacemaker abnormalities or disruptions in electrical conduction, lead to arrhythmic disorders may be benign, typical, threatening, ultimately fatal, occurs in clinical practice, patients on digitalis, anaesthesia or acute myocardial infarction. Both traditional and genetic animal models are: In-vitro: Isolated ventricular Myocytes, Guinea pig papillary muscles, Patch-Clamp Experiments, Porcine Atrial Myocytes, Guinea pig ventricular myocytes, Guinea pig papillary muscle: action potential and refractory period, Langendorff technique, Arrhythmia by acetylcholine or potassium. Acquired arrhythmia disorders: Transverse Aortic Constriction, Myocardial Ischemia, Complete Heart Block and AV Node Ablation, Chronic Tachypacing, Inflammation, Metabolic and Drug-Induced Arrhythmia. In-Vivo: Chemically induced arrhythmia: Aconitine antagonism, Digoxin-induced arrhythmia, Strophanthin/ouabain-induced arrhythmia, Adrenaline-induced arrhythmia, and Calcium-induced arrhythmia. Electrically induced arrhythmia: Ventricular fibrillation electrical threshold, Arrhythmia through programmed electrical stimulation, sudden coronary death in dogs, Exercise ventricular fibrillation. Genetic Arrhythmia: Channelopathies, Calcium Release Deficiency Syndrome, Long QT Syndrome, Short QT Syndrome, Brugada Syndrome. Genetic with Structural Heart Disease: Arrhythmogenic Right Ventricular Cardiomyopathy/Dysplasia, Dilated Cardiomyopathy, Hypertrophic Cardiomyopathy, Atrial Fibrillation, Sick Sinus Syndrome, Atrioventricular Block, Preexcitation Syndrome. Arrhythmia in Pluripotent Stem Cell Cardiomyocytes. Conclusion: Both traditional and genetic, experimental models of cardiac arrhythmias’ characteristics and significance help in development of new antiarrhythmic drugs.
Preprint
Full-text available
The cardiac conduction system (CCS) is a network of specialized cardiomyocytes coordinating electrical impulse propagation for synchronized heart contractions. Although the components of the CCS, including the sinoatrial node, atrioventricular node, His bundle, bundle branches, and Purkinje fibers, were anatomically discovered more than 100 years ago, their molecular constituents as well as regulators remain not fully understood. Here, we demonstrated the transcriptomic landscape of the postnatal mouse CCS at a single-cell resolution with spatial information. Integration of single-cell and spatial transcriptomics uncovered region-specific markers and zonation patterns of gene expression across the CCS, which were histologically validated. Network inference showed heterogeneous gene regulatory networks across the CCS. Notably, CCS region-specific gene regulation was recapitulated in a dish using neonatal mouse atrial and ventricular myocytes overexpressing CCS-specific transcription factors, Tbx3 and/or Irx3. This finding was supported by ATAC-seq of different CCS regions, Tbx3-ChIP-seq, and the identification of Irx motifs. Overall, this study demonstrates comprehensive molecular profiles within the postnatal CCS and provides evidence of the regulatory mechanisms contributing to its heterogeneity.
Article
The electrophysiological properties of sinoatrial (SA) node pacemaker cells vary in different regions of the node. In this study, we have investigated variation of the 4-aminopyridine (4-AP)-sensitive current as a function of the size (as measured by the cell capacitance) of SA node cells to elucidate the ionic mechanisms. The 10 mM 4-AP-sensitive current recorded from rabbit SA node cells was composed of transient and sustained components ( I trans and I sus , respectively). The activation and inactivation properties [activation: membrane potential at which conductance is half-maximally activated ( V h ) = 19.3 mV, slope factor ( k) = 15.0 mV; inactivation: V h = −31.5 mV, k = 7.2 mV] as well as the density of I trans (9.0 pA/pF on average at +50 mV) were independent of cell capacitance. In contrast, the density of I sus (0.97 pA/pF on average at +50 mV) was greater in larger cells, giving rise to a significant correlation with cell capacitance. The greater density of I sus in larger cells (presumably from the periphery) can explain the shorter action potential in the periphery of the SA node compared with that in the center. Thus variation of the 4-AP-sensitive current may be involved in regional differences in repolarization within the SA node.
Article
Evidence is presented that chloride ions are able to carry an appreciable amount of electric charge through the membrane of atrial muscle fibers, Purkinje fibers, and fibers of the S-A nodal pacemaker. An increase in the slope of diastolic depolarization of pacemaker fibers was recorded when Cl – was replaced by larger anions (sulfate) in the external medium. The rate of repolarization decreased when larger anions were substituted for chloride ions, and it was increased in nitrate or bromide solutions. The anionic permeability of the cell membrane of S-A node, atrial muscle, and Purkinje fibers seems to follow the series: NO 3 – > Br – > Cl – > CH 3 COO – > SO 4 = Evidence is presented that chloride ions contribute to diastolic depolarization of pacemaker fibers. The K ⁺ electrode properties of resting membrane of fibers of the S-A node were investigated at low and at constant extracellular Cl – concentration. It was found that above 16 mm K 2 SO 4 there is agreement between the resting potential and E k . Below this K ⁺ concentration a deviation of the resting potential from the line for a K ⁺ electrode was observed. Determinations of the Cl – content and of volume changes support the idea that the membrane of fibers of the S-A node and atrial muscle is permeable to Cl – and to K ⁺ ions.
Article
1. Single pacemaker cells were isolated from the sinus venosus of cane toad (Bufo marinus) in order to study the mechanisms involved in the spontaneous firing rate of action potentials. Intracellular calcium concentration ([Ca2+](i)) was measured with indo-1 to determine whether [Ca2+](i) influenced firing rate. A rapid transient rise of [Ca2+](i) was recorded together with each spontaneous action potential. [Ca2+](i) at the peak of systole was 655 +/- 64 nM and the minimum at the end of diastole was 195 +/- 15 nM. 2. Reduction of extracellular Ca2+ concentration from 2 to 0.5 mM caused a reduction in both systolic and diastolic [Ca2+](i) and the spontaneous firing rate also gradually declined. 3. Application of the acetoxymethyl (AM) ester of BAPTA (10 mu M), in order to increase intracellular calcium buffering, caused a decline in systolic and diastolic [Ca2+](i). The firing rate declined progressively until the cells stopped firing after 10-15 min. At the time that firing ceased, the diastolic [Ca2+](i) had declined by 141 +/- 38 nM. 4. In the presence of ryanodine (2 mu M), which interferes with Ca2+ release from the sarcoplasmic reticulum, the systolic and diastolic [Ca2+](i) both declined and the firing rate decreased until the cells stopped firing. At quiescence diastolic [Ca2+](i) had declined by 93 +/- 20 nM. 5. Exposure of the cells to Na+-free solution caused a rise in [Ca2+](i) which exceeded the systolic level after 4.8 +/- 0.3 s. This rise is consistent with Ca2+ entry on a Na+-Ca2+ exchanger. 6. Rapid application of caffeine (10-20 mM) to cells clamped at -60 mV caused a rapid increase in [Ca2+](i) which then spontaneously declined. An inward current with a similar time course to that of [Ca2+](i) was also generated. Application of Ni2+ (5 mM) or 2,4-dichlorobenzamil (25 mu M) reduced the amplitude of the inward current produced by caffeine by 96 +/- 1 % and 74 +/- 10%, respectively. In a Na+-free solution the caffeine-induced current was reduced by 93 +/- 7%. 7. Under a variety of circumstances the diastolic [Ca2+](i) showed a close association viith pacemaker firing rate. The existence of a Na+-Ca2+ exchanger and its estimated contribution to inward current during the pacemaker potential suggest that the Na+-Ca2+ exchange current makes a contribution to pacemaker activity.
Article
Sarcoplasmic reticulum (SR) Ca2+ cycling, that is, the Ca2+ clock, entrained by externally delivered action potentials has been a major focus in ventricular myocyte research for the past 5 decades. In contrast, the focus of pacemaker cell research has largely been limited to membrane-delimited pacemaker mechanisms (membrane clock) driven by ion channels, as the immediate cause for excitation. Recent robust experimental evidence, based on confocal cell imaging, and supported by numerical modeling suggests a novel concept: the normal rhythmic heart beat is governed by the tight integration of both intracellular Ca2+ and membrane clocks. In pacemaker cells the intracellular Ca2+ clock is. manifested by spontaneous, rhythmic submembrane local Ca2+ releases from SR, which are tightly controlled by a high degree of basal and reserve PKA-dependent protein phosphorylation. The Ca2+ releases rhythmically activate Na+/Ca2+ exchange inward currents that ignite action potentials, whose shape and ion fluxes are tuned by the membrane clock which, in turn, sustains operation of the intracellular Ca2+ clock. The idea that spontaneous SR Ca2+ releases initiate and regulate normal automaticity provides the key that reunites pacemaker and ventricular cell research, thus evolving a general theory of the initiation and strength of the heartbeat.
Article
Concomitant with the development of surgical treatment of cardiac arrythmias and management of myocardial ischemia, there is renewed interest in morphology of the intrinsic cardiac nervous system. In this study, we analyze the topography and structure of the human epicardiac neural plexus (ENP) as a system of seven ganglionated subplexuses. The morphology of the ENP was revealed by a histochemical method for acetylcholinesterase in whole hearts of 21 humans and examined by stereoscopic, contact, and bright-field microscopy. According to criteria established to distinguish ganglionated subplexuses, they are epicardiac extensions of mediastinal nerves entering the heart through discrete sites of the heart hilum and proceeding separately into regions of innervation by seven pathways, on the courses of which epicardiac ganglia, as wide ganglionated fields, are plentifully located. It was established that topography of epicardiac subplexuses was consistent from heart to heart. In general, the human right atrium was innervated by two subplexuses, the left atrium by three, the right ventricle by one, and the left ventricle by three subplexuses. The highest density of epicardiac ganglia was identified near the heart hilum, especially on the dorsal and dorsolateral surfaces of the left. atrium, where up to 50% of all cardiac ganglia were located. The number of epicardiac ganglia identified for the human hearts in this study ranged from 706 up to 1,560 and was not correlated with age in most heart regions. The human heart contained on average 836 +/- 76 epicardiac ganglia. The structural organization of ganglia and nerves within subplexuses was observed to vary considerably from heart to heart and in relation to age. The number of neurons identified for any epicardiac ganglion was significantly fewer in aged human compared with infants. By estimating the number of neurons within epicardiac ganglia and relating this to the number of ganglia in the human epicardium, it was calculated that approximately 43,000 intrinsic neurons might be present in the ENP in adult hearts and 94,000 neurons in young hearts (fetuses, neonates, and children). In conclusion, this study demonstrates the total ENP in humans using staining for acetylcholinesterase, and provides a morphological framework for an understanding of how intrinsic ganglia and nerves are structurally organized within the human heart. Anat Rec 259: 353-382, 2000. (C) 2000 Wiley-Liss, Inc.
Article
The effect of block of the L-type Ca2+ current by 2 ;/M nifedipine and of the Na+ current by 20 //M tetrodotoxin on the center (normally the leading pacemaker site) and periphery (latent pacemaker tissue) of the rabbit sinoatrial node was investigated. Spontaneous action potentials were recorded with microelectrodes from either an isolated right atrium containing the whole node or small balls of tissue (-0.3-0.4 mm in diameter) from different regions of the node. Nifedipine abolished the action potential in the center, but not usually in the periphery, in both the intact sinoatrial node and the small balls. Tetrodotoxin had no effect on electrical activity in small balls from the center, but it decreased the takeoff potential and upstroke velocity and slowed the spontaneous activity (by 49 ±10%; n = 11) in small balls from the periphery. It is concluded that whereas the L-type Ca2+ current plays an obligatory role in pacemaking in the center, the Na+ current plays a major role in pacemaking in the periphery. heart; cardiac
Book
1. Major cellular structures involved in E-C coupling. 2. Myofilaments: The end effector of E-C Coupling. 3. Sources and sinks of activator calcium. 4. Cardiac action potentials and ion channels. 5. Ca influx via sarcolemmal Ca channels. 6. Na/Ca exchange and the sarcolemmal Ca-pump. 7. Sarcoplasmic reticulum Ca uptake, content and release. 8. Excitation-contraction coupling. 9. Control of cardiac contraction by SR and sarcolemmal Ca fluxes. 10. Cardiac inotropy and Ca mismanagement. References. Index.
Conference Paper
Study objectives: We evaluate the hypothesis that pulse rate increases linearly with increased body temperature in infants and determine how much tachycardia in infants can be explained by a 1degreesC (1.8degreesF) increase in body temperature. Methods: Infants younger than 1 year and presenting to a pediatric emergency department were prospectively enrolled. Rectal temperature and pulse rate were measured. Research personnel rated behavioral state as sleeping, awake and quiet, fussy, or crying. Patients were excluded if they were fussy or crying or if they had any medical condition expected to cause tachycardia. The remaining patients were divided into 6 age-based groups. Linear regression analysis of pulse rate and temperature was performed for each group. Results: Four hundred ninety patients were enrolled. Pulse rate increased linearly with temperature in all age groups older than 2 months (adjusted r(2)=0.102 to 0.376) but not in infants younger than 2 months (adjusted r(2)=0.004). In infants aged 2 months or older, a multivariate linear regression model adjusted for age showed that pulse rate increased an average of 9.6 beats/min (95% confidence interval 7.7 to 11.5) per 1degreesC (1.8degreesF) increase in temperature (adjusted r(2)= 0.225). At any given temperature, the prediction interval for an individual's pulse rate had a span of approximately 64 beats/min. Conclusion: In infants 2 to 12 months of age, pulse rate increases linearly with body temperature, with a mean increase of 9.6 beats/min for each 1degreesC (1.8degreesF) increase in body temperature. Pulse rates of individual infants vary greatly, however, with a broad range of pulse rates observed at any given temperature.