ArticlePDF AvailableLiterature Review

Pathophysiological Mechanisms of Atrial Fibrillation: A Translational Appraisal

Authors:
  • Maastricht University, Maastricht, the Netherlands
  • University Heart and Vascular Center Hamburg
  • St Vincenz Hospital Paderborn

Abstract

Atrial fibrillation (AF) is an arrhythmia that can occur as the result of numerous different pathophysiological processes in the atria. Some aspects of the morphological and electrophysiological alterations promoting AF have been studied extensively in animal models. Atrial tachycardia or AF itself shortens atrial refractoriness and causes loss of atrial contractility. Aging, neurohumoral activation, and chronic atrial stretch due to structural heart disease activate a variety of signaling pathways leading to histological changes in the atria including myocyte hypertrophy, fibroblast proliferation, and complex alterations of the extracellular matrix including tissue fibrosis. These changes in electrical, contractile, and structural properties of the atria have been called "atrial remodeling." The resulting electrophysiological substrate is characterized by shortening of atrial refractoriness and reentrant wavelength or by local conduction heterogeneities caused by disruption of electrical interconnections between muscle bundles. Under these conditions, ectopic activity originating from the pulmonary veins or other sites is more likely to occur and to trigger longer episodes of AF. Many of these alterations also occur in patients with or at risk for AF, although the direct demonstration of these mechanisms is sometimes challenging. The diversity of etiological factors and electrophysiological mechanisms promoting AF in humans hampers the development of more effective therapy of AF. This review aims to give a translational overview on the biological basis of atrial remodeling and the proarrhythmic mechanisms involved in the fibrillation process. We pay attention to translation of pathophysiological insights gained from in vitro experiments and animal models to patients. Also, suggestions for future research objectives and therapeutical implications are discussed.
Pathophysiological Mechanisms of Atrial Fibrillation:
A Translational Appraisal
ULRICH SCHOTTEN, SANDER VERHEULE, PAULUS KIRCHHOF, AND ANDREAS GOETTE
Department of Physiology, University Maastricht, Maastricht, The Netherlands; Department of Cardiology and
Angiology, University Hospital of Münster, Münster; and Department of Cardiology, St. Vincenz Hospital,
Paderborn and Institute of Clinical Pharmacology, University Hospital Magdeburg, Magdeburg, Germany
I. Introduction 266
A. Clinical relevance: the “AF burden” 266
B. The natural history of AF 266
C. Clinical factors associated with AF 266
II. Atrial-Specific Aspects of Cardiac Physiology 267
A. Atrial electrophysiology 268
B. Electrophysiological basis of PV ectopy 269
C. Excitation-contraction coupling 272
III. Elementary Proarrhythmic Mechanisms During Atrial Fibrillation 273
A. Hierarchical and anarchical organization of AF 274
B. Cellular proarrhythmic mechanisms: automaticity and triggered activity 274
C. Mechanisms of reentry 276
IV. Experimental Paradigms of Atrial Fibrillation 278
A. AF and the autonomic nervous system 278
B. Chronic rapid atrial pacing 280
C. Heart failure 281
D. Animal models of chronic atrial dilatation 282
E. Models of AF combined with structural heart disease 282
F. Sterile pericarditis 283
G. Hypertension 283
H. Aging 283
I. Trangenic mouse models 284
V. Conditions and Mechanisms Contributing to the Initiation and Perpetuation of Atrial Fibrillation 284
A. Alterations in signaling pathways 284
B. Ion channel remodeling and shortening of refractoriness 289
C. Loss of atrial contractility and atrial dilatation 290
D. Alterations of atrial ca
2
handling and abnormal impulse formation 292
E. Atrial structural remodeling and conduction disturbances 294
F. Assessment of the AF substrate by fibrillation electrogram analysis 300
VI. Specific Forms of Atrial Fibrillation 304
A. Postoperative AF 304
B. Inherited cardiomyopathies and genetic defects associated with AF 305
VII. Summary and Future Perspectives 305
A. Summary 305
B. Current challenges and future perspectives 306
Schotten U, Verheule S, Kirchhof P, Goette A. Pathophysiological Mechanisms of Atrial Fibrillation: A Trans-
lational Appraisal. Physiol Rev 91: 265–325, 2011; doi:10.1152/physrev.00031.2009.—Atrial fibrillation (AF) is an
arrhythmia that can occur as the result of numerous different pathophysiological processes in the atria. Some
aspects of the morphological and electrophysiological alterations promoting AF have been studied extensively
in animal models. Atrial tachycardia or AF itself shortens atrial refractoriness and causes loss of atrial contractility.
Aging, neurohumoral activation, and chronic atrial stretch due to structural heart disease activate a variety of
signaling pathways leading to histological changes in the atria including myocyte hypertrophy, fibroblast prolifera-
tion, and complex alterations of the extracellular matrix including tissue fibrosis. These changes in electrical,
contractile, and structural properties of the atria have been called “atrial remodeling.” The resulting electrophysi-
ological substrate is characterized by shortening of atrial refractoriness and reentrant wavelength or by local
conduction heterogeneities caused by disruption of electrical interconnections between muscle bundles. Under
Physiol Rev 91: 265–325, 2011;
doi:10.1152/physrev.00031.2009.
www.prv.org 2650031-9333/11 Copyright © 2011 the American Physiological Society
on January 30, 2011physrev.physiology.orgDownloaded from
these conditions, ectopic activity originating from the pulmonary veins or other sites is more likely to occur and to
trigger longer episodes of AF. Many of these alterations also occur in patients with or at risk for AF, although the
direct demonstration of these mechanisms is sometimes challenging. The diversity of etiological factors and
electrophysiological mechanisms promoting AF in humans hampers the development of more effective therapy of
AF. This review aims to give a translational overview on the biological basis of atrial remodeling and the
proarrhythmic mechanisms involved in the fibrillation process. We pay attention to translation of pathophysiological
insights gained from in vitro experiments and animal models to patients. Also, suggestions for future research
objectives and therapeutical implications are discussed.
I. INTRODUCTION
A. Clinical Relevance: The “AF Burden”
Atrial fibrillation (AF) is the most common sustained
arrhythmia in humans, causing an increasing number of
complications and deaths (188, 289). Electrocardiogram
(ECG)-based surveys suggest that 1% of the total popu-
lation is affected (275). The number of patients with AF is
likely to double or triple within the next two to three
decades (240). The prevalence of AF is clearly age depen-
dent. The growing prevalence of AF can be explained in
part by the increasing average age in the human popula-
tion (85).
AF patients usually seek medical attention because
of AF-related symptoms. Treatment of these symptoms
has been the main motivation for AF therapy in the past.
In epidemiological and other observational studies, AF is
associated with excess death (43, 417). The available data
suggest that presence of AF approximately doubles death
rates in affected individuals, independent of other known
cardiovascular conditions. Similarly, AF worsens progno-
sis in patients with acute myocardial infarction and in
patients hospitalized for heart failure. Successful mainte-
nance of sinus rhythm was associated with longer survival
in the AFFIRM trial (123). Taken together, the available
data clearly suggest that AF increases mortality in af-
fected patients, although death rates in controlled trials
(AFFIRM, AF-CHF, RACE) are not affected when ion
channel-blocking drugs are needed to maintain sinus
rhythm.
Overall, 20 –25% of all strokes are caused by AF (384),
and AF-related strokes are more severe than strokes of
other origin. The importance of cardio-embolic stroke in
AF patients is highlighted by the fact that adequate anti-
coagulation in patients with AF can prevent strokes and
reduce mortality in patients at increased risk of stroke
(120, 227). Left ventricular function, the best-validated
clinical parameter for cardiac prognosis, can be markedly
impaired in AF patients and in some trials improved when
sinus rhythm is maintained for a longer period of time
(254, 284). However, it is worth noting that AF-CHF, a
recent large trial of sinus rhythm maintenance in patients
with already severely depressed left ventricular (LV) func-
tion, and AFFIRM, the first large “rate versus rhythm”
trial, did not detect an effect of sinus rhythm on LV
function (3, 471).
Apart from antithrombotic therapy, we have so far
failed to develop therapeutic interventions that improve
prognosis in AF patients (3, 244), underscoring the need
for better, possibly earlier and more comprehensive, man-
agement of AF, as highlighted by a recent consensus
statement (289).
B. The Natural History of AF
Systematic ECG monitoring studies indicate that a
substantial portion of AF episodes is asymptomatic (262).
Usually, these episodes are self-terminating. Over time,
typically over decades, the episodes become longer, and
eventually sustained forms of AF develop. This progres-
sive behavior of the arrhythmia has been demonstrated in
several large observational studies (283, 542). An example
of a patient with progressively longer episodes of AF is
depicted in Figure 1A. It should be noted, however, that
other courses of AF also occur, ranging from frequent
paroxysmal AF that never becomes persistent to perma-
nent AF developing as the first episode. As will be dis-
cussed in sections IV–VII, the progressive nature of AF is
partly caused by AF itself, but also reflects progression of
underlying structural heart diseases. There is evidence
that the time course of AF stabilization in patients with
structural heart disease is more rapid than in “lone AF”
patients (276).
C. Clinical Factors Associated With AF
Numerous clinical conditions are associated with an
increased incidence of AF. Most of them (Fig. 1B) con-
tribute to a gradual and progressive process of atrial
remodeling characterized by changes in ion channel func-
tion, Ca
2
homeostasis, and atrial structure such as cel-
lular hypertrophy, activation of fibroblasts, and tissue
fibrosis. These alterations may both favor the occurrence
of “triggers” for AF that initiate the arrhythmia and en-
hance the formation of a “substrate for AF” that promotes
its perpetuation. The association of clinical factors with
AF substrates and triggers will be discussed separately.
266 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
1. Clinical conditions associated with the initiation
of AF
Little is known about the mechanisms or clinical
conditions that initiate episodes of the arrhythmia. This
can be illustrated by the fact that many patients with an
accumulation of the above-mentioned AF-causing factors,
e.g., patients with advanced heart failure, never experi-
ence AF in their lifetime. A sizeable portion of patients
with lone AF suffer from “focal AF” that is initiated by
triggers that can be localized to preferential sites, mainly
the pulmonary veins (PV) (217). Electrical isolation of
PVs can prevent recurrence of AF in 70 80% of these lone
AF patients during a follow-up period of several years
(83). Stretch-activated or catecholamine-dependent auto-
maticity (52), as well as abnormal calcium handling (440),
have been suggested as mechanisms causing AF in focal
AF patients, but so far these mechanisms have not been
shown to be related to specific clinical conditions other
than AF itself.
2. Clinical conditions associated with the development
of a substrate for AF
Hypertension is found in 60 80% of AF patients
(396). Hypertension is an independent predictor of AF
(581), and it contributes to AF progression. Vascular dis-
ease, and most notably coronary artery disease, is found
in one-fourth to one-third of AF patients in surveys (396,
416), and may be associated with AF-related complica-
tions (255). Heart failure with dyspnea on exertion (NYHA
classes II-IV) is found in 30% of AF patients (416), and AF
is found in 30 40% of patients with heart failure (115).
Heart failure and AF appear to promote each other, with
AF compromising LV function, and LV dysfunction caus-
ing atrial dilation and pressure overload. Valvular heart
disease, especially mitral valve disease, was the most
common clinical condition associated with AF 50 years
ago. Early antibiotic therapy of streptococcal infections
has markedly reduced severe mitral valve disease in more
recent surveys (396, 416). These conditions are associated
with atrial dilatation, which plays an important causative
role in the development of a substrate of AF (sects. IV and
V). Diabetes mellitus is one of the established risk factors
for stroke in AF patients (190) and is found in 20% of all
patients with AF (396, 416). The high prevalence of dia-
betes mellitus in AF populations suggests that diabetes
may either cosegregate with AF due to similar conditions
that cause both AF and diabetes, or may imply that dia-
betes mellitus plays a causative role in the occurrence of
AF. Thyroid dysfunction, and especially hyperthyroidism,
is also associated with AF. Adequate therapy of thyroid
disease often terminates AF. Improved clinical manage-
ment of thyroid disease has rendered thyroid dysfuction
relatively rare in current AF populations (416).
It is well worth noting that all these clinical condi-
tions appear to enhance AF susceptibility in an additive
manner, as the prevalence of persistent AF increases
steadily depending on the number of such conditions
present (Fig. 1B). As we will discuss in sections IV–VI,
these conditions will increase AF propensity by many
diverse mechanisms. Understanding this diversity of the
mechanisms finally leading to AF is one of the unmet
challenges in unraveling AF pathophysiology.
II. ATRIAL-SPECIFIC ASPECTS OF
CARDIAC PHYSIOLOGY
Before proarrhythmic mechanisms in the atria will be
discussed, some atrial specific aspects of cardiac physi-
ology relevant to AF will be reviewed with a focus on
differences between atria and ventricles and regional dif-
ferences in function.
FIG.1.A: example of the “natural” course of the arrhythmia in an
atrial fibrillation (AF) patient. Typical pattern of time in AF (black) and
sinus rhythm (gray) over time (x-axis). AF progresses from undiagnosed
to first diagnosed, paroxysmal, persistent, and permanent. Flashes indi-
cate cardioversions as examples of therapeutic interventions that influ-
ence the time course of the arrhythmia. B: graph shows the type of AF
as a function of the number of “concomitant conditions” at the enroll-
ment visit into the AFNET registry that included 9,582 patients through-
out Germany from 2004 to 2006 (396). AF was classified as first episode,
paroxysmal, persistent, or permanent. Concomitant conditions were
defined as age 75 years, hypertension, diabetes (treated), cardiomyop-
athy, heart failure, valvular disease, or replacement. The proportion of
patients in permanent AF increases while the percentage of patients in
paroxysmal AF decreases almost linearly with increasing number of
concomitant conditions. The proportion of patients in persistent AF
remains relatively constant, suggesting that this is a transitory state
between paroxysmal and permanent AF. First episode of AF becomes
less likely in the patient population with many conditions.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 267
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
A. Atrial Electrophysiology
The predominant shape of the atrial action potential
is triangular with a gradual repolarization phase as shown
in the top right inset in Figure 2. If a plateau is present, it
is less pronounced than in ventricular myocytes. The left
panel of Figure 2 shows a human atrial action potential
and its main underlying ionic currents. As in ventricular
myocytes, the main depolarizing currents are the rapidly
activating and inactivating Na
-current (I
Na
) and the L-
type Ca
2
current (I
CaL
), which has somewhat slower
kinetics. The differences in action potential morphology
between atria and ventricles are mainly caused by differ-
ences in ion channel current density and kinetics of re-
polarizing currents. Some types of ion channels are selec-
tively expressed by atrial myocytes. Atrial-specific ion
channels represent interesting targets for cardioversion
of AF (160), given the risk for ventricular proarrhythmia
of traditional class I drugs (reduction of excitability and
conduction velocity, mainly by sodium channel blockade)
and class III drugs (prolongation of refractoriness, mainly
by potassium channel blockade) (465).
The ultrarapid delayed rectifier current (I
Kur
) acti-
vates 100 times more quickly than the rapid delayed
rectifier current (I
Kr
). I
Kur
significantly contributes to
atrial repolarization in most species. Kv1.5, the
-subunit
of I
Kur
, is expressed by atrial myocytes but to a much
lower extent in ventricular myocytes (373, 606). Most
compounds blocking I
Kur
also inhibit the transient out-
ward current (I
to
) and the acetylcholine-activated inward
rectifier current (I
KACh
) (488, 555). These drugs prolong
the atrial effective refractory period (AERP) and decrease
AF stability, without affecting the QT time (48, 136).
I
KACh
shortens the atrial action potential duration
(APD) during vagal activity. The pore-forming Kir3.x
-subunits are expressed in the atria but not in the ven-
tricles (149). Tertiapin-Q selectively inhibits Kir3.x chan-
nels like I
KACh
without affecting Kir2.x channels (inward
rectifier current I
K1
) (265). In dogs with sustained AF,
tertiapin-Q can terminate AF episodes without changing
ventricular repolarization (94). Another agent under de-
velopment, NIP-142, inhibits both Kir3.x and Kv1.5 chan-
nels, leading to a atrial-selective prolongation of APD and
AERP (369). In dogs, NIP-142 prolonged AERP by 10% and
cardioverted vagally induced AF without an effect on
ventricular refractoriness (401).
Recently, the presence of apamin-sensitive Ca
2
-ac-
tivated potassium channels (I
KCa
) in the heart was dem-
onstrated (636). In mouse hearts, the current density of
I
KCa
and its effect on APD are larger in the atria than in
the ventricles (636). Of the three genes encoding I
KCa
, SK1
(KCNN1) and SK2 (KCNN2) are expressed more in the
atria than in the ventricles, while the expression of SK3
(KCNN3) is similar (567). The degree of APD shortening
mediated by SK channels depends on the cytosolic Ca
2
concentration, thus possibly contributing to an interac-
tion of Ca
2
handling and repolarization. SK2 channels
colocalize with L-type calcium channels through a mutual
interaction with
-actinin 2 (349). In some animal models,
a role of SK2 channels in the development of atrial re-
modeling or AF has been suggested. In rabbit PVs, inter-
mittent burst pacing leads to an increased trafficking of
SK2 channels to the membrane, resulting in increased
apamine-sensitive current and APD shortening (432). In
SK2 knockout mice, APD was prolonged and AF induc-
ibility was increased probably due to enhanced triggered
activity (333). Furthermore, pharmacological block of
I
KCa
resulted in prolongation of atrial refractoriness and
termination of AF in rats, guinea pigs, and rabbits (144).
However, another study demonstrated little contribution
of SK channels to repolarization in normal rat and dog
atrial myocytes (402), and the role of I
KCa
blockade in
large-animal models of AF remains to be determined.
Recently, genome-wide association studies demonstrated
FIG. 2. The atrial action potential. Top right inset: comparison of
action potentials recorded at 37°C in different species at a pacing cycle
length close to the normal sinus cycle length. Top left: human atrial
action potential simulated using the Courtemanche model (126). The
traces below display the time course of the main ionic currents respon-
sible for the action potential. Inward (depolarizing) currents are shaded
gray, and outward (repolarizing) currents are black. Bottom right: table
lists the currents with the channel alpha subunits and the responsible
genes.
268 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
an association between lone AF and common gene vari-
ants of SK3 (166). It is currently unclear whether this
association is related to changes in myocyte repolariza-
tion or to effects in other cells types.
The inward rectifier current I
K1
stabilizes the resting
membrane potential and is an important determinant for
the initial depolarization and final repolarization of the
action potential (346). In atrial myocytes, inward rectifi-
cation of I
K1
at depolarized potentials is incomplete, al-
lowing I
K1
to play a role throughout the repolarization
phase (195). In addition, the density of I
K1
is 5- to 10-fold
smaller in atrial myocytes than in ventricular myocytes
(195). As a result, the membrane resistance of atrial myo-
cytes at rest is relatively high (205), and thus less depo-
larizing current is required to reach the action potential
threshold, making atrial myocytes inherently more excit-
able. This is reflected in model simulations that show that
the critical size of an ectopic focus required to drive the
surrounding myocardium is smaller in the atrium than in
the ventricle (269).
In the atria, gap junction channels are formed by
connexin (Cx) 40 and Cx43, and these two connexin
subtypes may combine to form heteromeric channels
(165). In contrast, ventricular myocytes predominantly
express Cx43 (575). Manipulation of Cx40 may offer a
way to affect atrial but not ventricular conduction, al-
though within the heart, Cx40 is also expressed in the
sinoatrial and atrioventricular nodes, the Purkinje system,
and endothelial cells (575).
1. Regional variation in atrial electrophysiology
Within the atria, considerable regional variation in
action potential morphology exists. As will be discussed
in detail in section IIB, PV myocytes differ from left atrial
myocytes, with a more depolarized resting membrane
potential, lower upstroke velocity, and shorter action po-
tential (161, 375). In isolated canine right atrial myocytes,
Feng et al. (178) reported relatively long APD with a
spike-and-dome morphology for myocytes from the crista
terminalis, while myocytes from close to the atrioventric-
ular (AV) groove had shorter action potentials (270 vs. 160
ms at 1-Hz stimulation, respectively). In this study, myo-
cytes from the pectinate muscles and atrial appendage
had intermediate durations. Differences in APD between
regions correlated to systematic differences in ion chan-
nel density. Burashnikov et al. (78) found a similar distri-
bution of action potential morphologies in perfused ca-
nine atria, in addition to a long APD in the bundle of
Bachmann, the main connection between right and left
atria. In some areas, like the crista terminalis, the epicar-
dium has shorter APDs than the endocardium (78, 178),
probably due to a relatively high I
to
current density in
epicardial myocytes (178). In older literature, myocytes
have been described with a fast action potential upstroke,
high conduction velocity, pronounced action potential
plateau, and long refractory period. These cells along the
caval border of the crista terminalis and in Bachmann’s
bundle remained excitable at higher potassium concen-
trations than normal atrial myocytes and may form spe-
cialized rapidly conducting tracts within the atrium (243,
579, 596).
B. Electrophysiological Basis of PV Ectopy
In humans, paroxysms of AF often originate in the
myocardial sleeves of the PVs. Haissaguerre et al. (217)
were the first to demonstrate that repetitive activity with
a very short cycle length may occur in the PVs, and that
radiofrequency ablation of such drivers can be used to
treat AF.
1. Morphology of PV myocytes
In theory, rapid activity in the PV area might result
from new impulse formation due to automaticity or trig-
gered activity or from (micro-)reentry due to abnormali-
ties in tissue structure. An argument favoring automatic-
ity is that during embryonic development, some markers,
in particular HNK-1 (51, 268), are expressed in areas
which will later form the conduction system and nodal
tissue as well as in the PV region. However, by all ac-
counts, the PVs in the adult heart consist mainly of myo-
cytes morphologically very similar to normal atrial myo-
cytes. Thus the PVs probably do not harbor large areas of
hidden nodal tissue. There are diverging reports about the
occurrence of abnormal myocytes scattered throughout
the myocardial sleeves. Using electron microscopy,
Masani (368) observed small, pale myocytes resembling
nodal P cells in normal rat PVs. In normal dog PVs, no
evidence for morphologically abnormal myocytes was
found in two studies (241, 585), whereas large PAS-posi-
tive cells (i.e., cells with a high glycogen content) were
found in another (110). In an electron microscopy study
on human PVs, Perez-Lugones et al. (447) have reported
the presence of nodal-like P cells, transitional cells, and
large Purkinje-like myocytes in patients with a history of
AF. These cell types were not present in PVs from pa-
tients without a history of AF (447). However, this study
has engendered criticism of the histological technique
itself and of the use of purely histological criteria to
classify electrical phenotypes of myocytes (12). Recently,
Morel et al. (391) have demonstrated the presence of
small Cajal-like cells in the interstitium of human PVs.
Similar cells are involved in pacemaking in the intestine,
but their role in PV automaticity is unclear, and their
presence has also been demonstrated in normal atrial
myocardium (235). The contribution of scattered abnor-
mal cell types with possible pacemaking properties re-
mains to be established. In general, a certain critical size
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 269
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
is required for an ectopic focus to act as a driver for the
surrounding myocardium (269). Propagation from an ec-
topic focus is most likely when electrical coupling grad-
ually increases from the focus to the surrounding muscle,
because a high degree of electrical coupling would effec-
tively silence the focus (reviewed in Ref. 270). Interest-
ingly, a recent study has indicated that at the PV-left atrial
junction, large PAS-positive myocytes expressing the
pacemaker channel protein HCN4, are more separated by
fibrosis and inflammatory infiltrates in chronic AF pa-
tients than in sinus rhythm patients (415).
2. Electrophysiology of isolated PV myocytes
Conflicting evidence about possible differences in
cellular electrophysiology between normal atrial myocar-
dium and myocardial sleeves has come from both record-
ings of isolated myocytes (current and voltage clamp) and
multicellular preparations (optical mapping and micro-
electrode recordings).
Voltage-clamp recordings on isolated PV myocytes
from healthy dogs indicate that PV myocytes have a larger
density of slow delayed rectifier current (I
Ks
) and I
Kr
,a
lower density of I
to
and I
CaL
than normal atrial myocytes
while I
Na
, the Na
/Ca
2
exchange current (I
NCX
) and
T-type Ca
2
current (I
CaT
) densities were similar to that in
normal atrial myocytes (161, 375). These data agree with
action potential recordings in myocardial sleeve prepara-
tions showing that PV myocytes have a more depolarized
resting membrane potential and action potentials with a
slower upstroke velocity and shorter duration, without
showing spontaneous diastolic depolarizations, afterde-
polarizations, or automaticity (161). In combination, the
lower upstroke velocity and relatively short APD may
contribute to a larger propensity for reentry within the
myocardial sleeve. As in the rest of the atria (see section
IVB), rapid atrial pacing reduced I
to
and I
CaL
in PV myo-
cytes. As a result, differences in APD between the PV and
left atrium became smaller (95), arguing against a major
role for the PVs in AF maintenance in this model. In
addition, Coutu et al. (127) reported that calcium handling
and its
-adrenergic response in PV myocytes was very
similar to that of left atrial myocytes, both for myocytes
isolated from normal dogs and after a week of rapid atrial
pacing.
A competing body of evidence has been presented by
Chen and co-workers (101, 104). In isolated myocytes
from both canine and rabbit PVs, a large proportion of
cells showed spontaneous activity (40% resp. 76%), in
some cases with unusual action potential morphologies
(101, 104). In voltage-clamp recordings, spontaneously
active myocytes had a low density of I
K1
, a larger delayed
rectifier amplitude, and sometimes showed a current re-
sembling the hyperpolarization-activated pacemaker
(funny) current (I
f
) (104). After 6 8 wk of rapid atrial
pacing in dogs, the beating rates of spontaneously active
PV myocytes had doubled, and a higher incidence of early
and delayed afterdepolarizations was observed (101). In
this and further studies from the same group, the isolation
procedure, in which left atrial cavity (104) or PV lumen
(101) were perfused, may have affected the electrical
behavior of the resulting myocytes. Also, if normal PV
myocytes would have such a pronounced tendency to-
wards automaticity, it is surprising that AF paroxysms or
premature atrial beats originating in the PVs are not ob-
served more frequently in healthy animals.
3. Electrical activity of myocardial sleeves
The earliest reports of PV automaticity in atrial prep-
arations far predate the discovery of PV involvement in
paroxysmal AF (70, 557). Cheung (106) later described
activity of guinea pig PVs as a subsidiary pacemaker with
a low, regular intrinsic frequency.
In a study on superfused PVs from normal dogs, Chen
et al. (103) reported a large heterogeneity of action po-
tential morphologies in the myocardial sleeves. This was
accompanied by a high incidence of high-frequency irreg-
ular activity, which was further increased in PVs from
dogs after 6 8 wk of rapid atrial pacing (103). However,
this high intrinsic arrhythmic tendency of canine PVs was
not confirmed in studies by other laboratories, which
showed a homogeneous distribution of action potentials
throughout the myocardial sleeves without spontaneous
activity under baseline conditions (Fig. 3A) (95). Rat PVs
did not display spontaneous activity, but it was observed
during infusion of norepinephrine or a combination of
-
and
-agonists (371). Rabbit PVs did not show spontane-
ous activity under baseline conditions, but PV pacemaker
activity with a diastolic depolarization was readily in-
duced by an acceleration in pacing rate in the presence of
low concentration of ryanodine (249). Thus most studies
on PV preparations indicate that the PVs are not sponta-
neously active under normal conditions, but that sponta-
neous and triggered activity can sometimes be induced,
particularly during sympathicomimetic treatment (see
also sect. IVA). In addition to the PVs, arrhythmic activity
may also originate from “myocardial sleeves” in other
atrial regions, such as myocardial extensions in the atrio-
ventricular valves and coronary sinus (37, 267, 621– 623).
Wit and co-workers have amassed an extensive body of
data on catecholamine-induced delayed afterdepolariza-
tions (DADs) and triggered activity elicited by fast pacing
or extrastimulation in the canine coronary sinus (232, 566,
621). DADs in coronary sinus myocytes involve Ca
2
release from the sarcoplasmic reticulum (21) and activa-
tion of a transient inward current (565). The contribution
of catecholamine-induced triggered activity from the cor-
onary sinus in human AF is unclear at this point (280).
270 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
4. Tissue structure of PVs
Apart from differences in cellular electrophysiology, the
PV area also shows salient features in gross anatomy and fiber
geometry. The area between the PVs shows a strong preferen-
tial superior-inferior fiber orientation on the epicardial side,
corresponding with the main propagation direction during si-
nus rhythm (364). In dog PVs, fiber orientation on the endocar-
dial slide of the sleeves was predominantly circumferential,
often with sharp transitions in fiber orientation to the epicardial
side at the proximal region of the veins (Fig. 3B) (585). After
3–5 mo of rapid atrial pacing, canine PVs showed a heteroge-
neous increase in extracellular matrix volume, coinciding with
rapid repetitive activity originating in the veins (107). In pa-
tients, there may be a relation between PV structure and the
presence of AF, with a marked variation in PV anatomy be-
tween individuals (228, 285). The superior PVs had longer and
thicker sleeves in AF patients (285). Guerra et al. (215) specif-
ically linked areas of PV wall thickening to high-frequency
potentials and the origin of ectopic beats. PVs in AF patients
also showed more fibrosis and discontinuities (228).
5. Conduction patterns in myocardial sleeves
Fractionated electrograms have been reported in the
PV-left atrial junction in humans and dogs (261, 528). This
FIG. 3. Electrophysiology of pulmonary veins (PVs). A: homogeneous distribution of action potential morphologies along the myocardial sleeve.
Photomicrograph shows the muscular layer tapering off from the left atrial ostium to the vein, endocardial side up. [Modified from Wang et al. (602),
with permission from Elsevier.] B: partial cross section of a canine PV sleeve, Masson’s trichrome staining. Within the sleeve, sharp transitions in
fiber orientation can be observed. Fibers on the luminal side (indicated by “L”) predominantly show a circumferential orientation. [Modified from
Verheule et al. (585), with permission from Oxford University Press.] C: lines of conduction block within a sleeve correlate with transition in fiber
orientation. Right: endocardial fiber orientation, with asterisk and circle indicating abrupt changes in fiber orientation. Lines of conduction block,
perpendicular to the length axis of the vein, were observed during slow pacing (left) and especially when paced with extrastimuli (middle). [Modified
from Hocini et al. (241), with permission from Wolters Kluwer Health.] D: repetitive firing in a human PV. Initial three beats of PV “focal” firing
pattern following a normal sinus beat. All three beats originated at the same site, with reentrant conduction into the vein during beat 2 and
conduction block at part of the ostium during beat 3. [Modified from Patterson et al. (440), with permission from John Wiley and Sons.]
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 271
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
fractionation was explained by slow conduction and/or
conduction block between the myocardial sleeves of the
PVs and the adjacent atrial myocardium. Zones of con-
duction delay and fractionated electrograms in normal
canine PVs were associated with abrupt changes in fiber
orientation (Fig. 3C) (241). Using optical mapping on the
canine PV area, Arora et al. (23) demonstrated 2:1 con-
duction block from the proximal to distal part of PVs
during fast pacing. Reentry around a functional line of
block close to the PV-left atrial junction could be induced
by premature stimuli. The core of reentrant circuits were
also clustered in a zone of increased anisotropy at the
PV-left atrial junction in another study on canine PVs
(110). Using monophasic action potentials in AF patients,
Narayan et al. (405) described that the slope of the APD
restitution curve was larger than 1 in paroxysmal AF
patients, which allowed single premature beats to initiate
AF. In contrast, in persistent AF patients, the slope was
smaller than 1, due to the occurrence of local activation
delays (405). While application of acetylcholine flattened
the APD restitution curve and reduced APD alternans in
canine PVs, the inducibility of rapid reentrant activity was
increased (453). In addition, activation of inward rectifier
K
channels by adenosine increased dominant frequen-
cies at the PV-left atrial transition in paroxysmal AF pa-
tients, supporting a reentrant mechanism (25). Indeed,
unstable reentry circuits with entry and exit points at the
PV-LA junction have been mapped in human PVs, in con-
junction with relatively short effective refractory periods
and anisotropic conduction (310). Rapid reentrant PV
activity and fibrillatory conduction in the rest of the atria
may be interdependent, because PV isolation by RF abla-
tion reduced the probability both of bursts of PV tachy-
cardia and of persistent AF (429).
On the other hand, examples of focal spread of acti-
vation have also been reported. In normal canine atrial
preparation, optical mapping revealed focal activity with
a relatively long cycle length near the PV-left atrial junc-
tion during isoproterenol infusion (23). Using high-density
epicardial mapping, Zhou et al. (653) observed repetitive
focal activation patterns with a short cycle length in PVs
of dogs after 1 mo of rapid atrial pacing. Focal spread of
activation in the PVs, often with block within the myocar-
dial sleeve, was also reported in a canine heart failure
model (423) as well as in patients with AF (Fig. 3D). The
mechanism underlying these focal PV activation patterns
was compatible with triggered activity related to in-
creased Ca
2
load (440, 441) (see also sections III and V).
In summary, PVs show differences from the normal
atrial myocardium in cellular electrophysiology and fiber
geometry. Also, abnormal cell types have been described
in some studies. These peculiarities may act in concert to
cause arrhythmogenic activity of PVs.
C. Excitation-Contraction Coupling
Research during the past 50 years has revealed that
Ca
2
ubiquitously mediates excitation-contraction, exci-
tation-secretion, and excitation-transcription coupling.
Furthermore, numerous cellular responses are directly or
indirectly regulated by this important second messenger.
Mechanisms of excitation coupling in atria and ven-
tricles have been expertly reviewed in great detail else-
where (46, 55). Basically, excitation-contraction coupling
is initiated by depolarization of the cell membrane by an
action potential that triggers opening of voltage-depen-
dent L-type Ca
2
channels. Influx of Ca
2
triggers the
release of Ca
2
from the sarcoplasmic reticulum through
Ca
2
release channels or “ryanodine channels” (RYR) in a
process called Ca
2
-induced Ca
2
release. Ca
2
binds to
troponin C activating the actin-myosin filaments, and the
myocyte starts to contract. Relaxation is initiated by a
decline of the cytosolic Ca
2
concentration mainly by
resequestration of Ca
2
into the sarcoplasmic reticulum
by the sarcoplasmic reticulum Ca
2
-ATPase (SERCA),
which is tightly controlled by its inhibitory protein phos-
pholamban, and extrusion to the extracellular space by
the Na
/Ca
2
exchanger (NCX). In hearts of larger mam-
mals, like dog, rabbit, and human, reuptake into the sar-
coplasmic reticulum and elimination by the NCX account
for 70 and 30% of Ca
2
removal, respectively (36), while
in mice and rats up to 90% of Ca
2
is reuptaken into the
sarcoplasmic reticulum (see sect. IVI) (36).
Due to the specific ultrastructure of atrial myocytes,
the spatiotemporal pattern of Ca
2
release and the way
Ca
2
release is controlled differs from that in ventricular
myocytes (55). Most atrial myocytes lack appreciable t
tubules or possess only a rudimentary and more irregular
transverse t-tubular system (Fig. 4, Aand B) (32, 560). An
electron microscopic study in rat atrial myocytes showed
a transversely oriented tubular system along the Z-lines
which was formed from the membrane of the sarcoplas-
mic reticulum but not the sarcolemma (639). As in ven-
tricular myocytes, immunostaining of RYRs revealed that
most of these receptors are aligned along the Z-lines of
the sarcomeres (nonjunctional sarcoplasmic reticulum).
In atrial myocytes, an additional population of RYRs em-
bedded in sarcoplasmic reticular membranes is located
directly underneath the sarcolemma in close vicinity to
the L-type Ca
2
channels, thereby forming the junctional
sarcoplasmic reticulum (90, 229).
These ultrastructural properties of atrial myocytes
have important functional consequences. Upon opening
of voltage-dependent L-type Ca
2
channels, release of
Ca
2
starts underneath the sarcolemma at the junctional
sarcoplasmic reticulum. From there, Ca
2
release spreads
towards the center of the cell through Ca
2
-induced Ca
2
release mechanisms. This centripetal Ca
2
“wave propa-
gation” has been documented in atrial myocytes from
272 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
guinea pigs (342), rats (353), cats (258, 297), and humans
(229) (Fig. 4, Cand D). It has been suggested that the
regulation of centripetal wave propagation contributes to
positive inotropic modulation of excitation-contraction
coupling in atrial myocytes (354).
Another consequence of the atrial-specific cellular
ultrastructure is that in the junctional sarcoplasmic retic-
ulum, where RYR clusters are located in close vicinity to
L-type Ca
2
channels, the frequency of elementary Ca
2
release events (sparks), which can provoke proarrhyth-
mic Ca
2
waves, is higher than in central cellular regions.
In addition, L-type Ca
2
channel antagonists reduce the
spark frequency only in the junctional but only slightly in
the nonjunctional regions (515).
III. ELEMENTARY PROARRHYTHMIC
MECHANISMS DURING ATRIAL
FIBRILLATION
For a long time, reentry of excitation wavefronts was
considered the main mechanism of AF. The seminal discov-
ery of localized sources of paroxysmal AF originating from
the PVs by Haissaguerre et al. in 1998 (217) has renewed the
interest in “focal” sources of AF. Both cellular proarrhyth-
mic mechanisms, like automaticity or triggered activity, and
reentrant mechanisms might underlie these phenomena. In
the authors’ view, the relative contribution of these distinct
mechanisms, however, is likely to vary between individual
patients and cannot be fully determined at present.
FIG. 4. Excitation-contraction coupling in atrial myocytes. A: confocal image of a membrane staining of a rabbit atrial myocyte with
di-8-ANNEPS. The myocyte lacks a t-tubular system. B: schematic drawing of the ultrastructure of atrial myocytes. Ryanodine receptors (RYR) are
located underneath the sarcolemma in close vicinity to the L-type Ca
2
channels (junctional SR) or along transverse proportions of the SR oriented
along the Z-lines (nonjuntional SR). Inositol 1,4,5-trisphosphate receptors (IP
3
R) are located in the junctional SR. Corbular portions of the SR serve
as reuptake and storage sites. [Modified from Bootman et al. (55), with permission from J Cell Sci.] C: confocal line scan of a Ca
2
transient in a
cat atrial myocyte. The pseudo-color indicates Ca
2
concentration (blue represents low, red a high concentration). Ca
2
release is initiated
underneath the sarcolemma and spreads towards the center of the cell. [Modified from Blatter et al. (50), with permission from John Wiley and Sons.]
D: the central cellular Ca
2
transient (ct) follows the subsarcolemmal (ss) transient with a delay of 10 –30 ms but reaches a comparable amplitude.
[Modified from Blatter et al. (50), with permission from John Wiley and Sons.]
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 273
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
A. Hierarchical and Anarchical Organization of AF
In general, AF can be perpetuated by “hierarchical”
or “anarchical” mechanisms. In case of a hierarchical
organization of AF, the arrhythmia is driven by a rapid
localized source. As the atrial myocardium remote from
this site cannot follow the driver in a 1:1 fashion, irregular
conduction at a lower frequency ensues. Both focal dis-
charges and reentrant circuits can act as localized
sources. Irrespective of its nature, ablation of the source
will terminate AF. In the case of anarchical AF, multiple
nonlocalized sources act anarchically to sustain AF. As
long as a sufficient number of sources is present simulta-
neously, AF will be sustained. In such a scenario, ablation
of a localized driver is not possible, but ablation strategies
that restrict the propagation of wavelets may still be
successful.
Table 1 gives a few classic examples of hierarchical
and anarchical types of AF. Reentrant waves can drive AF
in both a hierarchical as well as in an anarchical form of
AF. Hierarchical AF can be maintained by multiple forms
of reentry. Stable reentry circuits (“mother waves”) driv-
ing AF can vary in size (430, 509) and are in some cases
determined by anatomical structures (73, 370). Even if
reentry circuits are unstable, they still might be able to
maintain sustained AF as long as they continuously re-
form and at least one reentry circuit is always present
(307). Other examples of hierarchical AF maintained by
reentry are rotors that can be either fixed or wandering
through the atria (356, 474, 525). Perpetuation of AF by
“multiple wavelets” represents an anarchical form of AF
with reentrant mechanism.
Cellular proarrhythmic mechanisms (automaticity
and triggered activity) can clearly drive AF in a hierarchi-
cal manner under certain circumstances (492). However,
whether cellular proarrhythmic mechanisms can produce
AF in an anarchical type of the arrhythmia is so far
hypothetical and technically difficult to determine. Until
today, no convincing experiments have been reported
supporting “polytopic ectopy” as a cause of AF. This does
not mean, however, that disseminated atrial focal dis-
charges cannot occur during AF. In fact, as will be dis-
cussed in section V, many experimental studies indirectly
imply the existence of disseminated atrial ectopy during
the fibrillation process as a contributor to the perpetua-
tion of AF.
B. Cellular Proarrhythmic Mechanisms:
Automaticity and Triggered Activity
1. Enhanced automaticity
Enhanced automaticity occurs when myocytes pos-
sessing pacemaker activity increase their rate of sponta-
neous discharge. Enhanced automaticity can be due to a
lowered threshold of the action potential upstroke (phase
0), a less negative maximal diastolic potential, or an in-
crease in the slope of spontaneous diastolic depolariza-
tion (phase 4). A characteristic feature of both normal as
well as enhanced automaticity is the phenomenon of over-
drive suppression, which is related to intracellular Na
accumulation (118, 374). As mentioned in section II B, the
presence of pacemaker cells in the pulmonary veins or in
other regions of the atria outside the sinus node is debat-
able (391). Over all, there is little evidence for the exis-
tence of enhanced automaticity as a proarrhythmic mech-
anism during AF.
2. Abnormal automaticity
Abnormal automaticity occurs when cells are depo-
larized by any kind of depolarizing current and the thresh-
old of inward currents is reached (179). In many cases,
the degree of depolarization does not allow for complete
recovery of Na
channels. Thus the upstroke of the action
potentials is often mediated by Ca
2
inward currents.
Abnormal automaticity is less sensitive to overdrive sup-
pression (118) but might be abolished by compounds that
shift the membrane potential to more negative values
such as activators of inward rectifier currents (acetylcho-
line and adenosine) (468).
3. Triggered activity
Triggered activity arises from membrane oscillations
following normal action potentials (i.e., the trigger). If
TABLE 1. Examples for hierarchical and anarchical organization of AF
Hierarchical AF Anarchical AF
Reentrant mechanisms Stable mother wave macroreentrant (73, 370, 430) Multiple wavelets (7, 386, 388, 509)
Stable mother wave microreentrant (356, 509)
Unstable reentry circuits (307)
Leading circle (5)
Rotor (fixed, wandering) (356, 474, 525)
Cellular proarrhythmic mechanisms
(automaticity/triggered activity) Automatic foci (491, 492) Disseminated atrial focal discharges (hypothetical)
While reentrant mechanisms can underlie both hierarchical and anarchical types of atrial fibrillation (AF), automaticity and triggered activity have
so far been demonstrated only in hierarchical forms of AF. Reference numbers are given in parentheses.
274 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
such membrane oscillations reach threshold of depolariz-
ing currents, they can provoke new action potentials.
Under certain circumstances, such triggered responses
can in turn elicit new action potentials, resulting in self-
sustaining runs of triggered activity (Fig. 5A). Depending
on when the membrane oscillation occurs, “early” and
“delayed” afterdepolarizations are distinguished.
A) DELAYED AFTERDEPOLARIZATIONS. Delayed afterdepolar-
izations (DADs) are membrane potential oscillations oc-
curring after full repolarization of the triggering action
potential. DADs are favored by conditions producing
Ca
2
overload, like ischemia,
-adrenergic stimulation,
low extracellular K
concentration, and tachycardia (117,
277). Excess Ca
2
is extruded to the extracellular space
by the Na
/Ca
2
exchanger. Due to the 3:1 stochiometry
(3 Na
are exchanged for 1 Ca
2
), the I
NCX
is electrogenic
and produces an inward current that depolarizes the cell.
This mechanism forms the basis for the characteristic rate
dependence of DADs. The faster the triggering rhythm,
the shorter the interval of the triggered response is and
the faster self-sustaining episodes of DADs are (277).
B) EARLY AFTERDEPOLARIZATIONS. Early afterdepolariza-
tions (EADs) are membrane oscillations occurring during
phase 2 or 3 of the action potential. They occur in the
presence of action potential prolongation. During the pro-
longed action potentials, “Ca
2
window currents” may get
activated producing a new action potential upstroke
(382). Ca
2
window currents occur when the voltage
threshold for activation and inactivation overlap, allowing
rapid transformational changes from inactivated to closed
and open states of the channel (264). Another mechanism
likely involved in EADs occurring during
-adrenergic
stimulation is spontaneous release of Ca
2
from the sar-
coplasmic reticulum due to elevated cytosolic Ca
2
con-
centrations (593). In most cases, EADs occur under bra-
dycardic conditions while DADs are more likely to occur
during tachycardia or rapid pacing.
More recently, another type of EADs was described
by Burashnikov and Antzelevitch (76). In dog atria ex-
posed to combined sympathetic and parasympathetic
stimulation, the first action potential occurring after a
pause can trigger an EAD during late phase 3 of the action
potential (“late phase 3 EADs,” see Fig. 5B) (76). These
EADs exclusively occurred when both a shortening of the
action potential (parasympathetic stimulation) and an in-
crease in Ca
2
load of the cells (sympathetic stimulation)
were present. During the pause, Ca
2
is accumulated in
the sarcoplasmic reticulum, and the triggering action po-
tential produces a strong release of Ca
2
that exceeds the
action potential in duration. Because the action potential
is short, the high cytosolic Ca
2
concentration and the
negative membrane potential generate a strong inwardly
directed I
NCX
that will produce the EAD. The important
conceptual difference between this type and other sorts
of triggered activity lies in the fact that late phase 3 EADs
are triggered by a strong but essentially normal Ca
2
release from the sarcoplasmic reticulum, while DADs and
many forms of EADs are due to abnormal spontaneous
Ca
2
release from intracellular Ca
2
stores. Late phase 3
EADs have been suggested to play a role in the immediate
recurrences of AF (77). Interestingly, shortly after cardio-
version of AF, ultra-short refractory periods have been
reported that might further increase the likelihood for
these EADs to occur and to result in reinitiation of AF
(155). On the other hand, the occurrence of late phase 3
EADs is strictly limited to the simultaneous presence of
agents abbreviating the action potential and enhancing
Ca
2
load (76, 77, 236, 440). The pharmacological inter-
ventions used in these studies (e.g., 1
M acetylcholine
1
M isoproterenol) produce massive shortening of the
action potential and provide strong enhancement of Ca
2
load. Whether such extreme conditions can occur in pa-
tients with or prone to AF is currently unclear.
FIG. 5. Cellular proarrhythmic mechanisms in the atria. A: sche-
matic illustration of delayed (DAD) and early afterdepolarizations
(EAD). Both forms of triggered activity can elicit single or runs of action
potentials (AP). B: comparison of the mechanism of DAD and late phase
3 EAD. While a DAD is elicited by an abnormal Ca
2
release from the
sarcoplasmic reticulum, late phase 3 EADs are due to a strong but
normal Ca
2
release which outlasts the action potential (dotted line). In
both cases, the I
NCX
involved in Ca
2
removal depolarizes the cell.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 275
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
C. Mechanisms of Reentry
1. Circus movement reentry
Circus movement reentry was first demonstrated in
rings prepared from jellyfish (372) and cardiac tissue
(381). It is characterized by an activation that can travel
along a preformed anatomical structure and reactivate
(reexcite) previously excited tissue. A prerequisite for
circus movement reentry is recovery of excitability after
the previous activation before the next activation reaches
the tissue again. As a consequence of this, a short refrac-
tory period and a low conduction velocity make circus
movement reentry more likely. The minimal pathlength
for circus movement reentry can be calculated as the
product of conduction velocity and refractory period
(wavelength) (613). If the path of the circuit is longer than
the wavelength, there is a delay between recovery of the
tissue and the moment of reexcitation which is called the
“temporal excitable gap.” The section of the path which
regained excitability before the next excitation is called
the “spatial excitable gap,” which can be calculated as the
product of conduction velocity and temporal excitable
gap (Fig. 6A).
Initiation of circus movement reentry requires unidi-
rectional conduction block often occurring in regions
with long refractory periods. Because of the presence of
an excitable gap, circus movement reentry can be en-
trained (600). External waves can invade the reentrant
circuit, causing block and termination of reentry.
2. The leading circle concept
In 1924, Garrey (192) proposed a concept of AF that
involved aspects of reentry without a clearly defined an-
atomical structure. In turtle cardiac muscle, he demon-
strated a sustained excitation wave rotating around a
stimulation electrode. While in these experiments the
stimulus site still might have represented an obstacle for
propagation, in 1973, Allessie et al. (6) provided the first
evidence that reentry does not necessarily require an
anatomical obstacle. In left atrial rabbit atria, tachycardia
induced by premature stimulation was due to excitation
by rotating waves. Transmembrane electrode recordings
demonstrated that the core was not fully activated but
instead showed electrotonic depolarizations preventing
the tissue from regaining full excitability. According to
the “leading circle concept,” the size of the reentry circuit
adapts to the smallest possible loop in which the wave
can continue to propagate (Fig. 6B) (5). Excitation
wavefronts are propagating through tissue with limited
excitability, and the excitable gap is small. Therefore,
the arrhythmia is relatively unstable so that small
changes in the properties of the tissue can significantly
affect the dynamics of the reentry process, the local-
ization of the circuit, and the activation cycle length.
Because of the small excitable gap, premature stimula-
tion is less likely (compared with circus movement
reentry) but still able to invade the reentry circuit
(600). In line with this, entrainment of AF in a limited
area around the stimulation electrode (1– 4 cm) has
been demonstrated in experimental (288) and clinical
studies (437).
3. Spiral wave reentry
The theory of spiral wave reentry originally stems
from observations of chemical reactions in excitable me-
dia (619) and has strongly been influenced by insights
obtained from imaging of intracellular Ca
2
waves in
FIG. 6. Mechanisms of reentry. A: circus movement reentry. The size of the anatomical obstacle, the conduction velocity, and the refractory
period are the main determinants of this kind of reentry. The spatial excitable gap is the section of the path in which full excitability has been
regained. B: leading circle concept. As no anatomic obstacle exists, the reentry path adopts the minimal possible path length, which depends on
conduction velocity and refractory period. The spatial excitable gap is small. The central region is rendered unexcitable by electrotonic
depolarization by the circulating fibrillation wave. C: rotor theory reentry. The rotor rotates around an excitable yet unexcited core. Lengths of
arrows show conduction velocity. D: chaotic activation pattern caused by multiple wavelets. Waves are separated by multiple lines of conduction
block. Block lines may also occur within waves and form pivot points. Asterisks denote waves appearing within the mapped area presumably due
to transmural conduction breakthroughs reflecting a 3-dimensional substrate for AF. See text for further explanation.
276 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
Xenopus oocytes (114, 316) and the results of computer
simulations (438, 571).
To understand spiral wave reentry, it is important to
realize that propagation in the heart depends on a critical
balance between the “source” and the “sink” of a depo-
larizing current. The source of a wavefront is the diffusion
current generated by excited tissue tending to depolarize
downstream cells, which act as a sink. If the sink is too
large, the source current is not sufficient to excite the
downstream cells and propagation fails. In convex wave-
fronts, cells at the leading tip have to activate more cells
in front of it, resulting in a relatively small source
current and a low conduction velocity. In a concave
wavefront, many cells contribute to the activation of a
lower number of downstream cells which accelerates
conduction (Fig. 7).
The classical protocol to induce spiral wave reentry
is to provoke a perpendicular collision of a wavefront
with the wavetail of another wave (119). Where the tissue
is still refractory, the colliding wave will block while the
wave encounters excitable tissue behind the wave tail.
The colliding wave will turn towards newly recovered
cells and in this way the reentry wave adopts the shape of
a rotor (Fig. 6C). Importantly, the curvature of the wave-
front increases towards the core. Because of increasing
source-sink mismatch, the conduction velocity declines
towards the core until block occurs (dotted line in Fig.
6C). Thus the core, though being excitable, remains un-
excited during spiral wave reentry. Also, the core will
tend to shorten the action potential duration in its vicinity
which together with the low conduction velocity explains
the short wavelength (conduction velocity times refrac-
tory period) in its proximity.
In general, spiral wave reentry provides a compre-
hensive concept for reentry during AF. The only, though
significant, shortcoming of this concept is that spiral wave
reentry has never been documented in AF in humans.
More specifically, to the best of the authors’ knowledge,
there is not a single graphical depiction of sustained spiral
waves occurring during AF in patients. A possible expla-
nation might be that in human atria electrophysiological
heterogeneity due to structural changes such as fibrosis is
much more pronounced than in many animal models, and
therefore, more complex propagation patterns occur. Inter-
estingly, structural remodeling resulting in fibrosis of the
atrial wall in dogs with heart failure has been shown to
reduce the stability of rotors and promote the existence of
“multiple unstable rotors” (556), a conduction pattern essen-
tially resembling multiple wavelets.
4. The multiple wavelet hypothesis
In the late 1950s, computer models of AF demon-
strated that, based on simple assumptions regarding re-
fractoriness and conduction velocity, reentrant wavelets
might wander through an excitable medium in a seem-
ingly chaotic pattern (387). According to Moe’s “multiple
wavelet hypothesis,” fibrillation wavefronts continuously
undergo wavefront-wavetail interactions resulting in
wavebreak and generation of new wavefronts. On the
other hand, block, collision, and fusion of wavefronts will
tend to reduce their number. As long as the number of
wavefronts does not decline below a critical level, multi-
ple wavelets will be capable to sustain the arrhythmia
(386, 388). Factors increasing the stability of the fibrilla-
tion process include shortening of the refractory period,
increased heterogeneity of refractoriness, slowing of con-
duction, and an increase of the tissue mass. In contrast,
prolongation of refractoriness, enhancement of conduc-
tion velocity, and reduction of the available substrate will
reduce the number of wavefronts until the arrhythmia
ceases.
In 1985, Allessie et al. (7) demonstrated for the first
time the existence of multiple wavefronts in canine atria
exposed to acetylcholine. Numerous experimental and
clinical observations could be reconciled with the multi-
ple wavelet hypothesis. For example, during the Maze
procedure, the atria are subdivided in multiple electrically
independent compartments that are too small to sustain
the arrhythmia (128, 129). A comparable mechanism can
be postulated for some ablation procedures (see sect. VI).
Furthermore, prolongation of refractoriness indeed has
been shown to reduce AF stability (87, 452).
It has recently been suggested that the multiple wave-
let hypothesis would actually not exclude the coexistence
of local sources of AF (578). These authors argue that in
certain substrates, stable rotors might act as a source of
multiple wavelets. As such, this might be true in specific
cases (655). Moe’s hypothesis, however, goes beyond the
simple existence of multiple wavelets. It implies that the
fibrillation process is actually driven by them and no
localized sources of AF exist (“anarchical” organization of
AF). The actual experimental demonstration of multiple
wavelets as the mechanism sustaining AF, however, is
technically challenging, since fibrillatory conduction
FIG. 7. Effect of wavefront curvature on conduction velocity. In a
convex wavefront, cells serve as current source for more than one
downstream cell. Because of the relatively strong sink, conduction
velocity is low. In contrast, current source of more than one cell adds up
in a convex wavefront. As a result, the source is relatively strong and
conduction velocity high.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 277
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
remote from a localized source cannot be distinguished
from multiple wavelets with current mapping tech-
niques. Thus the direct demonstration of Moe’s multiple
wavelet hypothesis would essentially require that all
other mechanisms potentially sustaining AF are ruled
out. This could only be achieved by recording all elec-
trical activity in the entire atrium, which is not possible
with the techniques currently available. As long as it is
not possible to identify all local sources of the arrhyth-
mia in AF patients, Moe’s multiple wavelet hypothesis
will remain an important but hypothetical model for the
perpetuation of AF.
IV. EXPERIMENTAL PARADIGMS OF
ATRIAL FIBRILLATION
In many patients, AF is associated with some form of
underlying heart disease. Congestive heart failure (CHF),
hypertension, valvular disease, and aging are all strong
clinical predictors of AF (275). To unravel the mecha-
nisms leading to AF, several of these factors have been
mimicked in animal models. However, in some patients,
AF is observed in the absence of underlying structural
heart disease or hypertension (“lone AF”). The analogous
animal model is that of rapid atrial pacing (RAP). In some
patients, AF episodes may be triggered by increased au-
tonomic activity, which also has been studied in animal
models. Postoperative AF has been mimicked in a canine
model of sterile pericarditis.
The controlled conditions in animal studies have al-
lowed analysis of separate factors contributing to AF
(e.g., rapid activation rates or dilatation). However, in
humans, these factors may be present in a mild form over
prolonged periods of time, and the final substrate of AF
may have evolved very slowly over a period of decades.
For practical purposes, stimuli in animal studies are ap-
plied in a more intense form, for example, creating acute
severe mitral regurgitation to study chronic dilatation
(587) and pronounced ventricular tachycardia causing a
progression towards decompensating heart failure within
a few weeks (326).
In addition, there are further important differences
between other animal models and human patient popula-
tions. Postoperative AF after open-heart surgery may not
be fully represented by sterile paricarditis models, in
which the most common arrhythmia is atrial flutter. Sev-
eral weeks of RAP are required for AF to become self-
sustaining, whereas lone AF patients often present with
sustained AF episodes. Finally, long-term animal studies
in models of aging and long-standing hypertension, two of
the most important clinical predictors of AF, are relatively
rare. With these caveats, we will give an overview of
various animal models and the mechanistic insights
gained from them.
A. AF and the Autonomic Nervous System
In paroxysmal AF patients, the onset of AF is fre-
quently preceded by altered autonomic activity (138). One
study has indicated that patients with lone AF tend to
show a vagal pattern of AF onset, while patients with
structural heart disease tend to show a sympathetic pat-
tern (125). However, other studies in lone AF patients
have indicated that AF onset is associated with a change
in autonomic balance rather than with an increase in
vagal or sympathetic drive alone (181, 563). Paroxysms of
AF and atrial tachycardia were observed in dog models of
both intermittent RAP and CHF induced by rapid ventric-
ular pacing (420, 554). In both models, monitoring of the
activities of the left stellate ganglion and vagal nerve in
ambulatory dogs revealed that the onset of AF or atrial
tachycardia was predominantly associated with simulta-
neous sympathovagal discharges.
It has long been recognized that cholinergic activity
can promote AF (192). Electrophysiological effects are
instantaneous upon application of acetylcholine or vagal
stimulation. Mainly through an increase in I
KACh
(and in
the presence of
-adrenergic stimulation also by a reduc-
tion in I
Ca,L
) APD and AERP are shortened, thereby de-
creasing the wavelength of reentry circuits (509). The
degree of AERP shortening depends on the acetylcholine
concentration applied or the intensity of vagal stimula-
tion. At very high concentrations of acetylcholine, AERP
may become as short as 30 ms in dogs (509). Although
cholinergic AF has been investigated for decades, the
exact mechanism is still controversial. In isolated atrial
preparations from normal animals, AF can often not be
induced. However, in the presence of acetylcholine, pro-
longed AF episodes are frequently observed after a single
premature beat. In isolated canine atrial preparations in
the presence of 0.5–2
M acetylcholine, Allessie et al. (7)
described multiple wavelets wandering through the atria
in a chaotic pattern, without any single tissue area dom-
inating the activation pattern. In contrast, Jalife and co-
workers (360, 525) reported a gradient in activation fre-
quency during AF, with the highest activation frequencies
occurring in the left atrium in the presence of acetylcho-
line. These high frequencies (up to 30 Hz) were proposed
to originate from stable microreentrant circuits or “ro-
tors” (356).
The mechanism of cholinergic AF may actually depend
on the acetylcholine concentration applied. Schuessler et
al. (509) have described conduction in a canine right
atrium preparation. Here, a premature stimulus caused
multiple wavelet reentry at acetylcholine concentrations
up to 1
M. In this small preparation, multiple wavelet
reentry did not sustain for longer than 2 s. Fibrillation was
only sustained (2 min) at very high acetylcholine con-
centrations (10
M), due to the formation of small and
localized reentrant circuits with a fast cycle length of
278 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
30 –50 ms that dominated activation in the entire prepa-
ration (Fig. 8A). Using a detailed mathematical model,
Kneller et al. (296) have demonstrated that ACh can sta-
bilize spiral wave reentry. A heterogeneous distribution of
ACh, resulting in dispersion of refractoriness, led to fibril-
latory conduction away from the core of the spiral wave.
Adrenergic regulation of atrial electrophysiology is
more complex.
-Adrenergic receptor activation in-
creases I
CaL
(331), I
Kur
(330, 648), I
Ks
(479), and I
KACh
/I
KH
(641). On the other hand,
-adrenergic stimulation inhib-
its I
to
(175), I
K1
(174), and I
KACh
/I
KH
(63, 641). In canine
atrial myocytes, the
-adrenergic agonist phenylephrine
slightly increases APD, whereas the
-agonist isoprotero-
nol decreases APD (334). As a net result of sympathetic
stimulation, the plateau potential of the action potential is
increased (527), while the total APD is unaffected or even
decreased (654). The resulting increase in sarcoplasmic
Ca
2
load may enhance triggered activity (620).
The macroscopic innervation of the heart has been
described in detail by Kawashima et al. (282). Within the
atria, the autonomic nervous system forms an intricate
system of ganglionated plexi, especially concentrated in a
number of interconnected fat pads (434). Targetted abla-
tion of these fat pads may (490, 493) or may not (131, 421)
be a useful adjunct in treating AF in patients. The effect of
vagal stimulation on atrial refractoriness is heteroge-
neous because of heterogeneity in the distribution of
parasympathetic nerve endings and/or M
2
-cholinoceptors
(4, 343). In contrast, the effect of sympathetic stimulation
on refractoriness is more homogeneous (343). Enhanced
heterogeneity in vagal innervation (by ablation of the right
PV fat pad) (237) and in sympathetic innervation (by
application of phenol) (424) can both increase AF stabil-
ity.
The PV region, and the PV-left atrial junction in par-
ticular, is rich in both parasympathetic and sympathetic
FIG. 8. Mechanism of AF in animal models. A: during infusion of acetylcholine (ACh), the mechanism of AF can depend on the applied dose.
At 10
6
M ACh, multiple wavelet reentry with functional lines of conduction block was observed in a canine right atrial preparation. At a
concentration of 10
4.5
M, a single micro-reentrant circuit became dominant in the same preparation. [Adapted from Schuessler et al. (509), with
permission from Wolters Kluwer Health.] B: AF begets AF: in control goats, a high-frequency burst evokes only5sofAF.After 24 h of artificially
maintained AF with a “fibrillation pacemaker,” episode length has increased to 20 s. After 2 wk of AF, episodes last for more than 24 h. [Modified
from Wijffels et al. (614).] C: AF with and without underlying structural heart disease. Fibrillation maps recorded after 48 h of rapid atrial pacing
(RAP) from the left atrial free wall of a control goat (left) and a goat after 4 wk of atrioventricular block (right). In dilated atria, fibrillation waves
were less uniform, and local crowding of isochrones occurred more frequently. Local conduction delays of 8 ms between neighboring electrodes
are displayed in the lower panels. A higher degree of dissociation of fibrillation waves was observed in dilated atria. [Adapted from Neuberger et
al. (410), with permission from Elsevier.]
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 279
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
nerve endings (22, 553). In addition, APD shortening dur-
ing acetylcholine infusion is relatively large in the poste-
rior left atrium (335), probably due to a relatively high
expression of I
KACh
(487). The normal physiological rele-
vance of the high degree of autonomic innervation and
responsiveness of the PV myocardium is unclear. As de-
tailed in section IIB, PV myocardial sleeves in healthy
animals or preparations probably do not show rapid spon-
taneous activity, but it can be elicited by both parasym-
pathetic and sympathetic stimulation. As explained in
section III, the combination of vagal and sympathetic
activity may provoke triggered activity (mainly late phase
3 EADs) or promote AF triggered by PV firing (441, 494).
This combination can create a state with short APD and
increased Ca
2
load in atrial myocytes and may explain
observations that simultaneous sympathovagal discharge
is associated with the onset of AF paroxysms in both
patients and animal models (described above).
The autonomic ganglia in the PV area can be acti-
vated directly by high-frequency electrical stimulation. In
dogs, this led to episodes of AF and atrial tachycardia that
could be inhibited by both sympathetic and vagal phar-
macological blockade (489). In isolated canine PV prepa-
rations, autonomic nerve stimulation decreased APD and
increased rapid firing associated with EADs (442). Vagal
blockade by atropine prevented APD shortening in these
preparations and abolished rapid firing. Sympathetic
blockade by atenolol or a high dose of ryanodine (prob-
ably reducing sarcoplasmic Ca
2
load) also prevented the
spontaneous PV activity.
In summary, the local electrophysiological properties
of the PV myocardium may play an important role in AF
paroxysms triggered by autonomic imbalance. However,
as shown recently by Lemola et al. (321), AF due to strong
vagal activation is still observed after selective PV isola-
tion, but it disappeared after selective ablation of the
autonomic ganglia overlying the PV ostia. Therefore, the
current clinical ablation strategies may be effective be-
cause they affect both the PV myocardium and the adja-
cent autonomic ganglia.
B. Chronic Rapid Atrial Pacing
A major advance in the understanding of AF was the
observation that AF itself induces atrial changes that
promote AF (393, 614). In a goat model, only seconds of
AF could be induced by burst pacing in control animals,
whereas AF episodes lasting hours were induced after 2
days of pacing. Sustained AF (24 h) was observed after
a week of artificially maintained AF (Fig. 8B) (614). In a
canine model, continuous RAP (400 beats/min) for 6 wk
increased the AF duration from seconds to minutes (194,
393, 647). The increased AF stability was associated with
a decrease in AERP to 55% of baseline values in both
goats and dogs (614, 647). In the dog, the spatial hetero-
geneity of the AERP was increased by RAP, and this
seemed to be an independent determinant of enhanced
AF vulnerability (172). RAP causes a progressive decrease
in the densities of I
CaL
and I
to
and an increase in I
K1
(96,
647). This process of “electrical remodeling” results in a
shift towards the repolarizing currents, leading to APD
and AERP shortening (see also section VBand Fig. 10). As
in atria infused with acetylcholine, AERP shortening can
contribute to AF stability.
Interestingly, the time courses of AERP shortening
and AF stabilization diverge. In the goat, the AERP was
already maximally reduced after 24 h of AF, whereas AF
stability continued to increase over the ensuing weeks
(614). Moreover, in goats subjected to sequential 4-wk
periods of AF separated by 1 wk of sinus rhythm to allow
full recovery of the AERP, an acceleration of AF stabili-
zation was observed, independent of alterations in AERP
(562). These findings indicate that besides electrical re-
modeling, at least one other factor contributes to the
progression of AF.
Structural remodeling has been proposed as a candi-
date for this “second factor,” since it takes place in a
slower time domain of weeks to months. RAP induces
structural alterations that develop progressively, includ-
ing atrial dilatation (517), myocyte hypertrophy, loss of
sarcomeres, accumulation of glycogen, and mitochondrial
abnormalities (30, 393). In contrast to the rapid recovery
of the AERP after restoration of sinus rhythm, recovery of
morphological abnormalities is slow and incomplete (29).
The contribution of these structural changes to AF stabil-
ity is unknown at this moment. Overall, fibrosis is not
increased in the atria of AF goats and dogs (30, 326).
However, the volume of extracellular matrix (ECM) per
myocyte in goats increases during months of AF (29). This
pattern is similar to interstitial fibrosis or “microfibrosis”
observed during aging, which can lead to slow, discontin-
uous conduction during transverse propagation (529). In-
deed, a recent study compared goats with 6 mo of AF to
goats with 3 mo of AF, demonstrating a higher degree of
dissociation at the later time point, leading to a larger
number of smaller fibrillation waves (587a). This increase
in complexity of fibrillatory conduction was accompanied
by myocyte hypertrophy and increased interstitial fibro-
sis. In cell cultures, rapid pacing can cause myocytes to
release factors that alter the behavior of cultured fibro-
blasts (81). In vivo, such paracrine factors released by
myocytes at high activation rates may modulate fibro-
blasts function, possibly contributing to ECM remodeling.
Another factor that can directly affect myocardial
conduction is electrical coupling between myocytes me-
diated by gap junctions. In the goat model, no quantitative
changes in the total expression of the gap junction pro-
teins Cx40 and Cx43 levels were found, but the distribu-
tion of Cx40 became more heterogeneous during persis-
280 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
tent AF (572). The significance of Cx40 heterogeneity for
AF is unclear; conduction during slow pacing was not
affected (572), but the effects during fibrillatory conduc-
tion may be more pronounced.
For the interpretation of RAP models, the possible
contribution of high and irregular ventricular rates to
atrial structural remodeling should be considered. In
the dog, most early studies were performed with a
single atrial burst pacemaker, leading to a high ventric-
ular rate. Most researchers now combine RAP with
complete AV block and a separate low-rate ventricular
pacemaker (80 –100 beats/min). This distinction among
RAP models may explain some of the discrepancies in
the literature. For example, dogs with a single atrial
pacemaker show a progressive reduction in atrial con-
duction velocity (194) and an underlying decrease in
I
Na
(193), whereas the double pacemaker/AV block
model does not (326, 586). One study in goats has
investigated the role of high ventricular rates. Atrial
structural abnormalities were observed in a model of
RAP alone, but were virtually absent in a model of RAP
with a controlled ventricular rate (500). In dogs, 3 mo
of RAP increased atrial fibrosis even after AV node
ablation, but to a smaller degree than in a group with-
out controlled ventricular rate (31). In sheep, 15 wk of
RAP with and without controlled ventricular rate were
compared (13). With a controlled ventricular rate, many
animals did not develop persistent AF, and no signifi-
cant increase in fibrosis was observed. In contrast, in
the group without AV block, persistent AF developed
more rapidly, along with conduction heterogeneity and
extensive atrial fibrosis. Thus RAP without a controlled
ventricular rate appears to provoke structural changes
that increase AF stability. These changes might be
caused by a mild form of a ventricular tachycardiomy-
opathy. RAP models with and without controlled ven-
tricular rate are both relevant to human AF, corre-
sponding to AF without and with effective rate control,
respectively.
Taken together, the increased stability of AF in mod-
els of RAP involves electrical remodeling by shortening of
AERP and contribution of a “second factor,” which is
likely to involve structural alterations.
C. Heart Failure
To investigate how CHF promotes AF, an experimen-
tal model of CHF has been used by Nattel and co-workers
(326). Overt heart failure was induced by pacing dogs
with a ventricular pacemaker at a high rate (200 beats/
min) for 5 wk. The duration of induced AF episodes was
increased from a few seconds in control dogs to several
minutes in CHF dogs. Although in this model some ionic
currents were altered (see sect. Vand Fig. 10), CHF did
not reduce APD and AERP or increase AERP heteroge-
neity. In fact, AERP was significantly increased in intact
dogs (97, 539). In addition, CHF did not alter the overall
conduction velocity during slow pacing (326). Therefore,
the mechanisms leading to increased AF stability due to
CHF and RAP show important differences.
In contrast to RAP models, atrial fibrosis is dramati-
cally increased in CHF dogs (326), with large areas of
connective tissue. These structural abnormalities were
accompanied by regional conduction heterogeneity (Fig.
12C). Further studies have supported the key role of atrial
fibrosis in the promotion of AF in this model (97, 522).
When rapid ventricular pacing was followed by a 1-mo
period of slow ventricular pacing, ventricular function
and atrial dimensions recovered completely (97). At this
point, atrial electrophysiology had also normalized, but
atrial fibrosis, conduction heterogeneities, and AF stabil-
ity remained present, indicating that AF stability was
determined by structural rather than electrical remodel-
ing (97, 522).
Atrial damage in the CHF model occurs rapidly,
with a peak in inflammation, apoptosis, and necrosis
within 24 h after activation of the ventricular pace-
maker (220). These indicators of tissue injury gradually
disappeared in the following 5 wk. In the ventricle,
fibrosis developed only slowly over the entire period of
ventricular pacing, leading to a modest increase in
fibrosis over a 5-wk period. In fact, changes in gene
expression were far more extensive in the left atrium
than in the left ventricle (89).
The renin-angiotensin system is an important media-
tor in the development of an AF substrate in CHF. In the
atria of CHF dogs, tissue angiotensin II levels were rapidly
increased by rapid ventricular pacing and remained ele-
vated thereafter (220). Inhibition of ANG II production by
the angiotensin-converting enzyme (ACE) blocker enala-
pril not only attenuated CHF-induced atrial fibrosis but
also reduced conduction heterogeneity and AF stability
(328, 518). Atrial fibrosis and increased AF stability in the
canine CHF model can also be inhibited by simvastatin
(523), pirfenidone (318), polyunsaturated
3 fatty acids
(478), and sprionolactone (640). The PPAR
activator
fenofibrate was not effective in this model (523), but in
rabbits with tachycardiomyopathy-induced CHF, the
PPAR
acitvator pioglitazone did reduce atrial structural
remodeling and AF stability (519).
The differences in atrial pathology between CHF and
RAP models are reflected in gene expression profiles as
assessed by microarrays (88). In the canine RAP model,
changes entailed mainly downregulation of gene expres-
sion. In the CHF model, changes are quantitatively larger
and include a strong upregulation of extracellular matrix-
related genes.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 281
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
D. Animal Models of Chronic Atrial Dilatation
Atrial dilatation can be both cause and consequence
of AF. In patients, atrial enlargement correlates with the
incidence of AF (183, 233), and left atrial size is a strong
independent predictor for the development of AF (42). On
the other hand, several studies imply that AF also causes
atrial enlargement (485), and cardioversion to SR leads to
a decrease in atrial dimensions (607).
Chronic atrial dilatation occurs in several animal
models with an increased AF vulnerability. In a dog model
of RAP with a controlled ventricular rate, the diastolic left
atrial surface area increased by 24% (estimated by con-
sidering the atrium as a sphere with an area of 4
r
2
)
(517). In a sheep model of chronic hypertension, left atrial
surface area increased by 22% (292), and by an estimated
80% in a canine CHF model (517). However, in these
models, the role of atrial dilatation itself in the develop-
ment of a substrate of AF cannot be established. For this
purpose, several models of chronic atrial dilatation with-
out heart failure have been developed. Although most of
these models display an increased AF vulnerability, they
differ considerably in the structural changes associated
with atrial dilatation. Apart from any structural changes,
the increase in substrate size inherent to atrial dilatation
will also contribute to AF stability, as shown by Zou et al.
in a mathematical model (655).
Partial occlusion of the pulmonary artery and partial
avulsion of the tricuspid valve in dogs resulted in enlarged
right atria with cellular hypertrophy and increased inter-
stitial fibrosis (59). The increased AF stability in this
model was not due to electrical remodeling since action
potentials were not significantly different from control
dogs. In dogs with mitral regurgitation (MR) due to partial
mitral valve avulsion, the left atrium dilated within min-
utes, followed by a gradual increase to 167% of the base-
line left atrial area after 3 wk (587). Histological analysis
demonstrated areas of inflammatory infiltrates and
slightly increased fibrosis, without myocyte hypertrophy
(214, 587). Cx40 and Cx43 expression in the atrial free
walls did not show marked changes (214). The increased
duration of induced AF episodes could not be explained
by a decrease in wavelength, because the AERP was
homogeneously increased and the overall conduction ve-
locity was unchanged. High-resolution optical mapping
revealed heterogeneous left atrial conduction heterogene-
ity during pacing with short cycle lengths and extrastimu-
lation (see Fig. 12A), indicating that the increased AF
vulnerability was caused by increased conduction heter-
ogeneity associated with structural changes (586), remi-
niscent of the canine CHF model. By analyzing activation
sequences during AF in a canine MR model, Cox et al.
(129) showed that AF was maintained by the presence of
multiple unstable reentrant circuits.
In another model of chronic atrial dilatation, the time
course of atrial dilatation and AF stabilization was studied
(411). Chronic complete AV block in goats resulted in a
slow idioventricular rhythm, and volume overload of the
ventricles caused a progressive atrial enlargement and
myocyte hypertrophy. In a time period of 4 wk of AV
block, the right atrial area gradually increased by 29%.
Atrial dilatation was paralleled by a gradual increase in
AF stability while AERP and dispersion of refractoriness
remained constant. During sinus rhythm and slow pacing,
atrial conduction velocity was slightly increased. How-
ever, atrial mapping showed a higher incidence of areas
with slow conduction during fast pacing in dilated atria
(see Fig. 12B). Interestingly, atrial fibrosis was not in-
creased in this model, suggesting that atrial fibrosis does
not form a necessary condition for increased AF vulner-
ability in dilated atria. Also, the expression of Cx40 and
Cx43 did not show marked alterations. The difference in
atrial fibrosis between models may be related to the time
course of atrial dilatation. In the goat AV block model, the
slower time course may allow atrial myocytes to adapt
and undergo cellular hypertrophy. In the canine models,
partial valvular avulsion causes a sudden onset of dilata-
tion, which may lead to acute damage and subsequent
replacement fibrosis.
In a rabbit model, an arteriovenous shunt led to
chronic overload with an estimated increase in left atrial
surface area of 112% (238). Atrial conduction velocity was
significantly decreased by 30%. The inducibility of AF
was increased, and in the majority of cases, the arrhyth-
mias arose from the posterior left atrium, with either a
focal pattern of origin or a single reentrant circuit. In this
model, the expression levels of both Cx40 and Cx43 pro-
tein were significantly reduced (230).
In summary, animal models have demonstrated that
chronic atrial dilatation increases AF stability without
shortening of refractoriness. The contribution of in-
creased tissue mass is also probably limited. Fibrosis and
cellular hypertrophy occur in most but not all models of
atrial dilatation. The relationship between these factors
and alterations in conduction will be discussed further in
section VE.
E. Models of AF Combined With Structural
Heart Disease
Some 70 80% of AF patients have some preexisting
form of structural heart disease (325). There is evidence
that the time course of AF stabilization in patients with
structural heart disease is more rapid than in lone AF
patients (276, 283). Thus the most common scenario of AF
stabilization is that electrical remodeling takes place in
atria that are already structurally remodeled. From animal
studies, little is known about the process of remodeling
282 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
due to AF, AF stabilization, and the efficacy of antiar-
rhythmic drugs in this setting.
Everett et al. (170) studied a dog model in which
atrial dilatation due to mitral avulsion was combined with
RAP for 6 wk, leading to sustained AF. After cardiover-
sion, spontaneous AF episodes and a high incidence of
atrial premature beats were observed, but these phenom-
ena disappeared within a week. However, atrial structural
changes and the vulnerability to induced AF remained.
Neuberger et al. (410) created AV block in a goat model,
leading to atrial dilatation. After 4 wk, RAP was applied
for 48 h. In dilated atria, RAP still led to AERP shortening,
but the AF cycle length did not decrease, indicating a
longer excitable gap during AF. The duration of AF epi-
sodes after 48 h of RAP was significantly prolonged in
dilated atria compared with nondilated atria, with an
increased incidence of conduction block (Fig. 8C).
In dogs, RAP with a preexisting substrate of CHF
caused less AERP shortening than RAP alone (521). How-
ever, the conduction velocity was decreased by RAP in
the presence of CHF but not by either CHF or RAP alone.
Atrial fibrosis and AF stability due to CHF was not exac-
erbated by 1 wk of RAP, but the incidence of sustained AF
was increased. Also, RAP in the presence of CHF failed to
alter the density of I
to
,I
NCX
, and I
Ks
and had a smaller
effect on I
CaL
and I
K1
than RAP alone (96). Thus the
effects of AF in atria with preexisting structural heart
disease, a common clinical scenario, may differ from
those of lone AF.
F. Sterile Pericarditis
A clinically highly relevant form of AF occurs in up to
50% of patients in the first days after cardiac surgery
(157). This postoperative form of AF has been mimicked
in a canine model of sterile pericarditis, in which the atria
are dusted with talcum powder and covered with a layer
of gauze at the end of open heart surgery (435). The
predominant arrhythmia in this model is atrial flutter.
However, as in patients (467, 598), the initiation of flutter
requires a previous period of AF (520). During this tran-
sitional phase of AF, a line of functional block can de-
velop, most often along the crista terminalis between the
superior and inferior vena cava (73, 370, 569). If this line
of block becomes long enough, a stable reentry circuit
can develop and AF converts to flutter. Conversely, if the
line of block at the core of the flutter circuit becomes too
short, flutter can convert to AF (430, 601). During AF in
the sterile pericarditis model, unstable reentrant circuits
with a very short cycle length have been observed (307).
These unstable circuits often disappear and reform
around anatomic obstacles or functional lines of block.
Even if only one circuit is present, fibrillatory conduction
can be maintained when the circuit is too fast for 1:1
conduction to the rest of the atrium.
Altered connexin expression may be an important
contributor to atrial conduction disturbances in sterile
pericarditis. Cx40, Cx43, and
-actinin disappeared in the
epicardial layer and were reduced in the midmyocardial
layer, while the endocardial distribution was normal
(472). These changes may lead to transmural heterogene-
ity in electrical coupling and a disruption of epicardial
conduction. The importance of the inflammatory compo-
nent in arrhythmia vulnerability is underscored by studies
showing that anti-inflammatory agents such as atorvasta-
tin (308) and prednisone (202) can prevent AF and atrial
flutter in this model.
G. Hypertension
Long-term animal studies in models of hypertension,
an important clinical predictor of AF, are relatively rare.
In 11-mo-old spontaneously hypertensive rats, the induc-
ibility of atrial tachycardia was increased, accompanied
by a decrease in I
CaL
and an increase in fibrosis (108). In
a sheep model of long-standing elevated blood pressure
induced by prenatal corticosteroid exposure (mean arte-
rial pressure 94 vs. 71 mmHg in control sheep), 4- to
5-yr-old animals had increased AF stability, reduced con-
duction velocities, no change in refractoriness, and in-
creased fibrosis with myocyte hypertrophy and myolysis
(292). Thus models of elevated blood pressure seem to
share some important features of the AF substrate with
the more extensively studied models of structural heart
disease of CHF and dilatation. Unfortunately, no data are
available on the signaling pathways involved in the struc-
tural remodeling process induced by hypertension.
H. Aging
Although the clinical prevalence of AF strongly in-
creases with age (176), the intrinsic contribution of aging
to AF promotion is difficult to study in humans due to the
long time span of senescence and the presence of numer-
ous confounding factors. AF is also quite common in
some species of domestic animals during aging, but its
pathogenesis has not been studied systematically (71).
Several animal models have been used to investigate ag-
ing-related AF. Spach et al. (536) studied conduction pat-
terns in canine atria and found an age-dependent slowing
of transverse propagation that correlated with the devel-
opment of extensive collagenous septa that separated
small groups of fibers. Similarly, in a comparative study of
young and old rats, AF could only be induced in old rats,
while the AERP was not different between age groups
(231). Instead, conduction slowing and enhanced AF vul-
nerability were associated with an age-dependent in-
crease in heterogeneous interstitial fibrosis.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 283
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
In dogs, Koura et al. (304) demonstrated that with
age, the amount of interstitial fibrosis and fatty infiltrates
increased and that Cx43 became increasingly concen-
trated at end-to-end connections between myocytes.
High-resolution optical mapping of a small tissue area
revealed enhanced anisotropy of conduction in old dogs,
while the APD was not altered (304). Most notably, ex-
tremely slow transverse conduction, sometimes causing a
“zig-zag” conduction pattern, was only observed in old
atria. Such dissociated conduction between adjacent fi-
bers may become widespread throughout the atria at
advanced age.
At the cellular level, a study performed in adult (1–5
yr) and old (8 yr) dogs showed an age-induced shift in
membrane currents that gave rise to a slight (15%)
increase in APD and ERP (Fig. 10) (17). In addition,
increased APD heterogeneity and slower conduction of
premature beats was observed in aged atrial tissue (16,
17). Despite these changes in atrial electrophysiology, the
inducibility of AF was not significantly increased in this
study (16). In rabbits, aging increased the tendency to-
ward formation of DADs in the PVs (624, 625).
Taken together, these studies show that even in the
absence of any underlying pathology, the senescent heart
possesses structural characteristics that predispose to
AF. However, in these animal models, the extent of elec-
trical and structural alterations is often not sufficient to
lead to the increased AF vulnerability that is associated
with aging in humans.
I. Trangenic Mouse Models
It was originally thought that fibrillation required a
substrate of a certain size (192, 390, 618). According to
this “critical mass” hypothesis, the mouse heart (at 0.0005
times the weight of the adult human heart) would be too
small to be able to fibrillate. However, more recent ob-
servations of fibrillation in very small hearts represent a
challenge to the critical mass hypothesis (570). In a study
on mice with selective atrial fibrosis due to overexpres-
sion of TGF-
1, the atrial wavelength was 15 mm (582).
With its length of 5 mm, these atria could probably not
accommodate more than one reentrant wavelet, assuming
a homogeneous substrate. However, increased atrial fi-
brosis in this model made the atria a structurally hetero-
geneous substrate with an increased AF inducibility (Fig.
12E).
APD shortening in response to administration of cho-
linergic agonists can also be sufficient to increase AF
vulnerability in normal mice (305, 597), and in one study,
AF could be induced in normal mice without pharmaco-
logical intervention (508). Thus, at least as a “proof of
principle,” the mouse heart can be a useful model for AF,
even when the underlying mechanism is likely to be re-
entrant. However, the question remains open to which
extent arrhythmias in very small mouse atria with the
complete absence or marked overexpression of a single
gene can be translated to human pathology.
In recent years, increased AF vulnerability in a num-
ber of transgenic mouse models with altered expression
of ion channels has been reported. Deletion of the gap
junction protein Cx40 leads to decreased atrial conduc-
tion velocity and increased AF stability (216, 583). In mice
overexpressing Kir2.1, the increase in I
K1
current was
associated with spontaneous AF episodes (332), perhaps
through the stabilization of rotors (141). Deletion of
KNCE1, an auxiliary subunit for I
Ks
that normally stabi-
lizes the open state, unexpectedly shortened APD and led
to spontaneous AF episodes (558). Deletion of Ca
2
-acti-
vated SK2 potassium channels also increased APD and AF
inducibility (333).
A growing number of transgenic mice display pro-
nounced atrial enlargement and an associated increase in
spontaneous or inducible AF: mice overexpressing junc-
tin (247), mice with cardiac-specific overexpression of
angiotensin converting enzyme (634), cAMP-response el-
ement modulator (395), TNF-
(473), junctate-1 (248), and
overexpression of the G
q protein (239). In these mice, it
is difficult to assess to what extent AF vulnerability re-
sults directly from the genetic defect or from atrial dila-
tation and the concomitant alterations in tissue structure.
While mouse models might be useful to study the
mechanisms of specific proarrhythmic phenomena as
such, translation of these findings to pathophysiology of
AF in humans is inherently problematic. The action po-
tential is much shorter so that the currents contributing to
atrial repolarization differ between mice and men (Fig. 2).
Also, Ca
2
handling in mice is characterized by a very
rapid Ca
2
reuptake and a very pronounced contribution
of Ca
2
reuptake to diastolic Ca
2
elimination (418).
Proarrhythmic mechanisms occurring in mice that are
related to altered Ca
2
handling or ion channel function
therefore not necessarily, and in some cases are not likely
to, reflect the electrophysiological phenomena that would
occur in humans with similar defects in cell function.
V. CONDITIONS AND MECHANISMS
CONTRIBUTING TO THE INITIATION AND
PERPETUATION OF ATRIAL FIBRILLATION
A. Alterations in Signaling Pathways
The structural and functional adaptations of the atria
to underlying heart disease or AF are the result of the
regulation by multiple signaling pathways that can occur
either as a consequence of AF or before the onset of AF,
usually triggered by underlying structural heart disease.
Figure 9 summarizes the AF-related changes in atrial sig-
naling.
284 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
1. Changes in atrial signaling preceding AF
A) ATRIAL PRESSURE AND VOLUME OVERLOAD. Chronic atrial
stretch appears to be one of the most prominent trigger
mechanisms for signaling changes involved in the patho-
genesis of AF. Interestingly, atria appear to react much
faster and more strongly to increased wall stress due to
dilatation than ventricular myocardium (221). The induc-
tion of heart failure by rapid ventricular pacing induces
the development of apoptosis and increased collagen syn-
thesis in the atria within a couple of days, whereas the
degree of such changes is substantially smaller and the time
course much slower in the ventricles (326). At the molecular
level, the development of atrial fibrosis due to pressure
and/or volume overload is mediated by both angiotensin
II-dependent and angiotensin II-independent mechanism
(328, 523). Left ventricular failure increases atrial synthesis
of angiotensin II, and thereby atrial fibrosis is induced via
activation of mitogen-activated protein kinases (201). Sig-
naling pathways mediated by angiotensin II type 1 recep-
tors (AT1 receptors) are linked to G proteins. Binding of
angiotensin II to AT1 receptors leads to tyrosine phos-
phorylation of receptor tyrosine kinases (656). The acti-
vated monomeric G protein Ras interacts with Raf-1 and
activated Raf-1 then phosphorylates the kinases MEK-1
and MEK-2. In the final step of this signaling cascade,
extracellular signal regulated kinases (ERK-1 and ERK-2)
are activated by phosphorylation. ERKs lead to activation
of transcription factors, such as Elk-1 and c-fos, which are
responsible for the effects on gene transcription (545).
Studies have shown a linear correlation between angio-
FIG. 9. Atrial signal transduction pathways regulating gene expression in atrial tissue. Signal transduction is induced by several autocrine and
paracrine factors. After binding of various ligands (angiotensin II, TGF-
1, PDGF) to cell-surface receptors, numerous intracellular signaling
cascades are activated, which regulate essential programs of gene expression responsible for hypertrophy, proliferation, cell survival, and cell death.
A central role for redox-signaling is played by NF
B, which is activated after reactive oxygen species (ROS) are generated. One source for ROS is
the activated NADPH oxidase. Under baseline conditions, NF
B is retained in the cytosol as an inactive complex with inhibitory I
B proteins.
Activation occurs via rapid degradation of the complex and phosphorylation by I
B kinases. Angiotensin signaling encompasses several phosphor-
ylation steps. As a final step, phosphorylated MAP-kinases (ERK, p38, and JNK) induce the cellular response by activation of transcription factors.
Cytokines bind to corresponding cytokine receptors, leading to the activation of Janus kinases (JAKs). STAT proteins are the common targets of
JAKs. Upon phosphorylation, STATs dimerize and translocate to the nucleus where they bind to interferon response elements or
-interferon-
activated sequences and thereby initiate target gene transcription. AF induces activation of several signal transduction pathways and intracellular
signaling molecules. Therefore, structural changes (hypertrophy, fibroblast activation and proliferation, apoptosis) are induced at the cellular and
tissue level. These alterations contribute to electrophysiological and morphological alterations of fibrillating atria. After binding of various ligands
(angiotensin II, TGF-
1, CTGF, PDGF) to cell-surface receptors or via increased calcium influx, intracellular signaling cascades are activated during
AF to induce the development of atrial hypertrophy, proliferation of fibroblast, increased collagen synthesis, increased expression of adhesion
molecules (VCAM-1, PAI-1), induction of apoptosis, and increased expression of autocrine and paracrine factors like angiotensin II, TGF, CTGF, and
matrix metalloproteases (MMP). AT-1, angiotensin II type 1 receptors; TGF-
, transforming growth factor
; PDGF, platelet-derived growth factor;
JNK, c-jun terminal kinase; ERK, extracellular signal regulated kinases; STAT, signal tranducer and activators of transcription; CaMK II, Ca
2
/
calmodulin-dependent protein kinase II; PAI, plasminogen activator inhibitor; VCAM, vascular cell adhesion molecule.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 285
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
tensin II, ERK1/2 activation, and the degree of atrial fi-
brosis (328). Another tyrosine kinase that is activated by
angiotensin II is janus kinase 2 (JAK2) (367). JAK2 ini-
tiates activation of transcription factors STAT-1 and
STAT-3. A recent study demonstrated that the angiotensin
II/Rac1/STAT3 pathway is an important signaling pathway
involved in atrial structural remodeling (564). In addition
to angiotensin II, pacing-induced ventricular failure also
increased atrial TGF-
and PDGF levels. TGF-
operates
predominantly by autocrine and paracrine mechanisms.
Binding of a TGF-
homodimer to two TGF-
type II
receptors causes phosphorylation of signaling molecules
belonging to the family known as SMADs. When phos-
phorylated, SMADs aggregate and enter the nucleus to
induce myocardial fibrosis (61). In addition, TGF-
can
redirect protein synthesis to favor expression of fetal
genes as described in fibrillating atria (439). A mouse
model that overexpresses TGF-
1 develops profound
atrial fibrosis and AF, whereas the ventricles are normal
(582). Interestingly, atrial fibroblasts are activated signif-
icantly faster than ventricular fibroblasts in CHF models,
explaining the rapid and more severe degree of interstitial
fibrosis in the atria (79). Burstein et al. (79) demonstrated
that atrial fibroblasts react more strongly to angiotensin II
than ventricular fibroblasts. Importantly, proliferation of
atrial fibroblasts was consistently more pronounced than
that of ventricular fibroblasts when stimulated with a
range of growth factors including angiotensin II and
TGF-
1 (79). Connective tissue growth factor (CTGF) is
implicated in various fibrotic disorders and is produced
by fibroblasts after activation by TGF-
1. Moreover, atrial
CTGF was identified as a candidate factor in CHF-induced
atrial fibrosis by analysis of gene networks (79).
Several studies clearly suggest that oxidative stress
contributes to atrial remodeling in CHF models (98). The
peroxisome proliferator-activated receptor-
(PPAR-
)
activator pioglitazone antagonizes angiotensin II actions
and possesses anti-inflammatory and antioxidant proper-
ties. Pioglitazone attenuated CHF-induced atrial struc-
tural remodeling and AF vulnerability (519). In the same
study, both pioglitazone and candesartan reduced TGF-
,
TNF-
, and mitogen-activated protein kinase, but neither
affected p38-kinase or c-Jun NH
2
-terminal kinase activa-
tion. In contrast,
-3 polyunsaturated fatty acids attenuate
CHF-related phosphorylation of the mitogen-activated
protein kinases ERK and p38 (478).
In failing human myocardium, NADPH oxidase-re-
lated ROS production increases, which in turn enhances
expression and activity of Rac1. Application of 3-hydroxy-
3-methylglutaryl coenzyme A reductase inhibitors (st-
atins) like simvastatin downregulate Rac1-GTPase activ-
ity by reducing isoprenylation and translocation of Rac1
to the cell membrane (2). Inhibition of Rac1 by statins
decreases NADPH oxidase-related reactive oxygen spe-
cies production in cardiac myocytes and reduces myocar-
dial hypertrophy. Furthermore, simvastatin reduces hu-
man atrial myofibroblast proliferation via a RhoA path-
way (455). In addition, simvastatin, but not fenofibrate
(PPAR-
agonist), inhibits canine atrial fibroblast prolif-
eration, which paralleled collagen-synthetic fibroblast
function (449). Thus statins and PPAR-
agonists have
very different efficacy in preventing CHF-related atrial
structural remodeling.
Chronic atrial stretch is also associated with an al-
tered expression of matrix metalloproteinases (MMP).
MMPs are a large group of enzymes that function to break
down the extracellular matrix. Patients with persistent AF
show decreased atrial expression/activity of MMP-1 and
increased tissue inhibitor of metalloproteinase (TIMP)-1
levels (200, 363). In AF patients with congestive heart
failure, an increased collagen I fraction appears to be
associated with upregulation of MMP-2 and downregula-
tion of TIMP-1 (121, 635). These apparent discrepancies
could be explained by temporal changes in MMPs func-
tion and the presence of concomitant cardiac diseases
(valvular regurgitation, heart failure, etc.), which have a
substantial effect on atrial MMP expression (14). In addi-
tion to MMPs, ADAMs (a disintegrin and metalloprotein-
ase) also influence interstitial matrix composition and
cell-cell and cell-matrix interactions. ADAMs form a large
family of membrane-bound glycoproteins that function in
proteolysis, signaling, cell adhesion, and cleavage-secre-
tion of membrane-bound proteins (498). Arndt et al. (20)
described the effect of AF in patients with concomitant
heart diseases on regulation of ADAMs. Atrial tissue of
patients with permanent AF shows increased levels of
ADAM10 and ADAM15. Membrane expression of ADAM15
is significantly upregulated during AF, whereas most
ADAM15 is largely confined to the cytoplasm during sinus
rhythm. The ADAM15/
1-integrin ratio is significantly in-
creased in fibrillating tissue and correlates with the left
atrial diameter and the duration of fibrillation. Thus reg-
ulation of MMPs and ADAMs may influence the composi-
tion of interstitial matrix and, furthermore, contributes to
geometrical changes and dilatation of the atria and ven-
tricles (257).
B) AGING. Another important factor that influences
atrial pathology is aging (199). Both replacement and
reactive fibrosis, possibly induced by increased levels of
TGF-
are postulated to contribute to structural changes
in the elderly. Other profibrotic molecular mechanisms
(e.g., angiotensin II, bradykinin, endothelin-1) influencing
this process still need to be clarified. In very old myocar-
dium, higher levels of active p38
MAPK
in atrial trabeculae
after ischemia point towards an increased cellular stress,
which is even more pronounced after postischemic reper-
fusion. A recent study showed that aging significantly
influences PV anatomy, which may contribute to generate
triggers for AF. CT analysis revealed that left atrial ap-
286 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
pendages and the four PV trunks become dilated in pa-
tients older than 50 years (436).
In addition to morphological changes, the aging pro-
cess is also accompanied by modified cellular Ca
2
han-
dling (624). Aging is accompanied by upregulation of
InsP
3
Rs and transient receptor potential C (TRPC) chan-
nels but decreased SERCA activity. Whether these factors
contribute to PV arrhythmogenesis is currently unknown.
Another structural change commonly observed with
increasing age is isolated atrial amyloidosis (IAA). The
incidence of IAA increases with age (466). The formation
of amyloid fibrils takes place by a nucleated growth mech-
anism. As IAA is formed locally at the site where atrial
natriuretic peptide (ANP) is synthesized, a high local
concentration of the precursor protein and enhanced syn-
thesis of ANP contribute to the formation of amyloid
deposits. Apart from an association with AF, the presence
of atrial amyloid correlated with age, gender, surface
P-wave duration, and valvular diseases. The highest prev-
alence of atrial amyloid were found in patients undergo-
ing mitral valve replacement (466). Of note, the amount of
amyloid is inversely related to the amount of interstitial
fibrosis, supporting the concept that amyloid itself pro-
duces an arrhythmogenic substrate (466, 644).
2. Changes in atrial singaling as a consequence of AF
A) IMPACT OF AF ON OXIDATIVE STRESS. Several lines of
evidence suggest an association between oxidative stress
and AF (92, 300, 379). Rapid atrial pacing models showed
that AF decreases tissue ascorbate levels and increases
protein nitration, a biomarker of oxidative and nitrosative
stress (92). Biochemical evidence for oxidation by per-
oxynitrite and hydroxyl (
OH) radicals, both downstream
products of O
2
·
generation, has also been demonstrated
in experimental models. Thus AF itself induces substan-
tial oxidative stress in fibrillating atrial tissue. Recently, it
was shown that AF caused left atrial endocardial dysfunc-
tion manifested as a 73% decrease in NO
·
production and
a 1.8-fold upregulation of plasminogen activator inhibi-
tor-1 (PAI-1) (82). One possible explanation for the reduc-
tion in NO
·
is increased oxidative degradation by O
2
·
.
Recent studies suggest that increased O
2
·
production is
at least partly the result of increased NAD(P)H oxidase
and xanthine oxidase (XO) activities. Interestingly, the
data indicate that the increase in O
2
·
production is
greater in the left than in the right atrium. Increased
NAD(P)H oxidase activity could be explained by an in-
crease in active Rac1 (151). The mechanism of increased
NAD(P)H oxidase activity appeared to be increased en-
zyme activation (486). A membrane-bound gp91
phox
con-
taining NAD(P)H oxidase in atrial myocytes was the main
source of atrial superoxide production in human atrial
tissue (82). In contrast to findings in sinus rhythm pa-
tients, NO synthases (NOSs) contributed significantly to
atrial superoxide production in fibrillating atria, suggest-
ing that increased oxidative stress in AF may lead to NOS
“uncoupling” (287). In addition, downregulation of eNOS
has also been described in rapid pacing models (75).
These findings indicate that a myocardial NAD(P)H oxi-
dase and, to a lesser extent, dysfunctional NOS contribute
significantly to superoxide production in the fibrillating
human atrial myocardium (287). Rapid pacing of atrial
tissue slices in vitro demonstrated that atrial tachycardia
is associated with mitochondrial dysfunction and oxida-
tive stress-activated signal transduction (75, 496). Thereby,
oxidative stress could contribute to metabolic and structural
atrial changes in AF. It has been shown that AF activates the
redox-sensitive NF-
B signaling pathway, which causes an
elevated expression of target genes, LOX-1 and ICAM-1.
Increased expression of these adhesion molecules may con-
tribute to an increased risk for platelet and leukocyte adhe-
sion to the atrial endocardium, which may initiate atrial
thrombogenesis (75). However, the true clinical impact of
these molecular changes on thrombogenesis remains to be
defined.
B) IMPACT OF AF ON CALCIUM-DEPENDENT PROTEASES AND
PHOSPHATASES.Ca
2
-dependent proteases like calpain and
phosphatases are activated during AF (66, 164, 198). Cal-
cineurin (Cn) is a calcium-activated serine-threonine
phosphatase composed of a catalytic A-subunit (59 63
kDa) and a regulatory B-subunit (19 kDa) (389). Three
catalytic genes (A-subunit) have been identified, of which
CnA
and CnA
are present in the heart (549). The in-
duction of cardiac hypertrophy has been associated with
an increase in CnA
(but not in CnA
) expression (72).
An elevation in the intracellular Ca
2
concentration leads
to calmodulin saturation and the subsequent activation of
calcineurin (130). Activated calcineurin dephosphorylates
nuclear factor of activated T cells (NFAT), allowing trans-
location of NFAT into the nucleus. NFATc3 plays a pivotal
role in regulating the hypertrophic pathways. Of note,
FK506, a Cn inhibitor, abolishes the hypertrophic re-
sponse induced by electrical pacing of atrial tissue slices
(74). Rapid pacing causes an upregulation of CnA activity,
leading to increased dephosphorylation of the transcrip-
tion factor NFATc3. This in turn increases the transcrip-
tion of genes responsible for the atrial hypertrophic cel-
lular response, characterized by an increased expression
of troponin I, ANP, and
-myosin heavy chain (74). The
presence of hypertrophied atrial myocytes during AF is a
consistent finding of several studies (30, 186). Increased
calcineurin enzyme activity was also shown in pigs with
AF (337). In this in vivo model, NFAT-c3 and NFAT-c4
were increased in the nuclei in AF tissue (337). In another
in vitro study, CnA activity increased during8hofrapid
pacing, but returned to baseline at 24 h (459). Interest-
ingly, CnA activity was associated with transcriptional
downregulation of calcium channel protein.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 287
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
288 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
The histopathology of fibrillating atria is thought to
be similar to chronically ischemic ventricular myocar-
dium (30, 186). Ischemia/reperfusion injury is mediated
by increased levels of cytosolic Ca
2
, which causes acti-
vation of the Ca
2
-dependent proteases calpain I and
calpain II (197). It had been proposed that AF-induced
atrial contractile dysfunction is at least partially due to
calpain I-dependent degradation of contractile proteins.
However, as will be explained in section VC, force gener-
ation and Ca
2
affinity of myofilaments are only modestly
altered in patients with AF. An in vitro study on HL-1
atrial myocytes showed that 24-h electrical field stimula-
tion at 5 Hz reduced plasmalemmal levels of L-type Ca
2
channel
1C
-subunit by 72% compared with controls,
whereas there was no change in amount of the potassium
channel subunits Kv4.3 and Kv1.5 (67). In that model,
rapid pacing induced marked changes in cellular struc-
ture; myolysis and nuclear condensation, paralleled by a
14-fold increase in calpain activity. Interestingly, inhibi-
tion of calpain, but not the calcium antagonist verapamil,
prevented these ultrastructural changes.
B. Ion Channel Remodeling and Shortening
of Refractoriness
In patients, paroxysms of AF often tend to become
longer with time, ultimately leading to persistent AF. An-
imal models of rapid atrial pacing have helped to under-
stand this progressive nature of the arrhythmia. AF itself
leads to changes which increase the stability of AF (614).
It is widely believed that “electrical remodeling” plays an
important role in this process. The rapid rates of AF cause
a shortening and a loss of rate adaptation of the AERP
(194, 614), as discussed in section IVB. In patients, the
existence of electrical remodeling is well established. Sev-
eral clinical studies have shown that atrial APD in AF
patients is shorter than in patients in sinus rhythm (26, 58,
132, 182, 646). In addition to AERP shortening, AF pa-
tients show a loss of rate adaptation of the AERP (26, 58,
182). Some studies have also reported an increased AERP
dispersion (378, 383).
The molecular mechanisms underlying APD shorten-
ing in human AF have been extensively investigated in
atrial myocytes isolated from the right atrial appendage of
patients undergoing open chest surgery. The properties of
these myocytes thus reflect not only adaptations to the
rhythm of the patients but also to underlying structural
heart disease, medication, and age (189). This shortcom-
ing can partly be overcome by matching respective con-
trol populations but remains a potential confounding fac-
tor in most of these experimental investigations.
The available data show that the molecular mecha-
nisms of electrical remodeling occur at the level of ex-
pression and/or phosphorylation of ion channels (Fig. 10,
expertly reviewed in Refs. 408, 626). Of the depolarizing
currents, I
Na
was unaltered (57) while I
CaL
density was
70% lower compared with patients in sinus rhythm (57,
68, 526, 576). The mRNA for the L-type channel
-subunit,
Cav1.2, shows a corresponding decrease both in RAP
animal models and AF patients (56, 68, 69, 573, 649).
However, some studies have reported a decrease of L-
type Ca
2
channel subunit protein expression (68, 69,
649), whereas others have found that protein levels were
unchanged (111, 506). Differences in patient populations
(underlying heart disease, antiarrhythmic drugs) may be
responsible for these discrepancies. Increased activity of
protein phosphatase 2A (PP2A) resulting in hypophos-
phorylation of the Ca
2
channel was found to cause I
CaL
downregulation in one study (111), although in another
study, L-type channel open probability was increased,
rather pointing towards decreased PP2A activity (294).
Finally, increased calpain activity may contribute to in-
creased proteolysis of L-type channel protein (66, 198).
Of the repolarizing potassium currents, I
to
(57, 209,
577) is strongly reduced in chronic AF. The resulting
slowing of phase 1 repolarization of the atrial action
potential has been found in human atrial action potentials
of patients with AF (112) but was not consistently found
in animal models. The inward rectifier potassium current
I
K1
(57, 146, 148, 577) and the corresponding subunits
Kir2.1 (mRNA and protein) were found to be increased in
AF patients (189). The reported increase in I
K1
probably
has a significant contribution in APD shortening (652).
AF decreases the mRNA and protein for the I
KACh
subunit Kir 3.1 and 3.4 (68, 69). In myocytes from AF
patients, the I
KACh
-mediated response to acetylcholine is
blunted (146), but I
KACh
is constitutively active due to
abnormal channel phosphorylation by protein kinase C
(PKC) (94, 145, 592), making this current an interesting
target for APD prolongation.
For I
Kur
, no change (57, 209, 628) or a decrease (62,
577) was reported. Christ et al. (112) recently reported
that amplitudes of both rapidly and slowly inactivating
components of I
Kur
were lower in AF patients. No
FIG. 10. Atrial action potentials in models of AF. Top panel: superposition of control action potentials with action potentials measured in atrial
myocytes from canine models of rapid atrial pacing (RAP), congestive heart failure (CHF), and aging and in myocytes from patients with chronic
AF and structural heart disease. The dotted line represents 0 mV. Action potentials are modified from the following references: RAP (647), CHF
(327), aging (16), human AF (57), and human SHD (302). Stimulation frequency was 1 Hz in all cases. The table provides an overview of alterations
in ionic currents in various animal models of AF and in myocytes from human patients with AF and structural heart disease. Changed are expressed
as “fold change.” Na, no data available. Note that the reported current changes due to RAP are derived from models of RAP with AV block.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 289
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
changes in the Na
/K
pump current were found in AF
patients (627). Apart from electrical remodeling caused
by AF, underlying heart disease in itself can also lead to
changes in ionic current that may also increase the like-
lihood for AF (see Fig. 10 for quantitative details) (626,
629).
Does the process of electrical remodeling explain the
progressive nature of AF in patients? Often, AF can be
cardioverted by agents that prolong the atrial APD. This
indicates that a prolongation of the AERP is antiarrhyth-
mic but does not show to which extent the AERP short-
ening contributes to AF stability. Importantly, the success
rate of chemical cardioversion is relatively high in recent-
onset AF, but its efficacy decreases with longer AF dura-
tion (15). The exact time course of electrical remodeling
during AF in humans is unknown, but in animal models it
is complete within at most a few days. The limited effi-
cacy of ion channel blockers in the treatment of chronic
AF indicates that other processes occurring more slowly
than electrical remodeling contribute to the stability of AF
in many patients.
Several studies reported that a short right atrial APD
directly after cardioversion was correlated to a higher AF
recurrence (124, 426). The question is how long this effect
persists. In most patients with recent onset of paroxysmal
AF, the time spent in sinus rhythm between two AF
episodes may be long enough to completely reverse elec-
trical remodeling. In animal models, complete reversal of
electrical remodeling takes place within 2–3 days of sinus
rhythm with a gradual prolongation of the AERP and
recovery of the rate adaptation response (504, 562, 615). A
number of clinical studies have demonstrated that after
cardioversion of persistent AF in patients, the AERP also
gradually increases and normal refractoriness is restored
within several days (357, 460, 646). Interestingly, several
other studies showed that increased vulnerability to AF
after cardioversion still exists 2– 4 wk after reversal of
electrical remodeling (561, 568). This discrepancy in time
course also indicates that the effect of other processes
than just electrical remodeling remain present after car-
dioversion of AF.
Expression and/or activity of the calmodulin-depen-
dent kinase II (CaMKII) has been shown to be enhanced
in fibrillating and dilated atria in animal models and pa-
tients with AF (409, 559). The most relevant target pro-
teins of CaMKII are ryanodine receptors, phospholamban,
L-type Ca
2
channels, and proteins involved in the epige-
netic regulation of atrial myocytes (e.g., histone-acety-
lases). In normal atrial myocytes, CaMKII activation plays
an important role in frequency adaptation of cell function.
In AF, CaMKII activation might contribute to generation
of abnormal impulse formation as will be described in
section VD.
C. Loss of Atrial Contractility and
Atrial Dilatation
Loss of atrial contractile function after cardioversion of
AF was first documented by Logan et al. in 1965 (Fig. 11A)
(344). Echocardiographic studies showed that this atrial
contractile dysfunction correlated with the duration of AF
and that it could take months before the atrial transport
function was fully recovered (358, 359). While after 2 wk
of AF, recovery of atrial contractile function was com-
plete within 24 h of sinus rhythm, it took more than 1 mo
to recover from AF lasting more than 6 wk (358). The
degree of atrial contractile function appears not to de-
pend on whether AF was cardioverted pharmacologically
or by direct-current shock (226). In cases with spontane-
ous termination of AF, a similar degree of atrial contrac-
tile dysfunction was demonstrated (171, 212).
The most important clinical consequence of loss of
atrial contractile dysfunction due to AF is the low blood
flow velocity in the atria following cardioversion, which
significantly contributes to the thromboembolic risk as-
sociated with AF (45). Most thromboembolic events oc-
cur shortly after cardioversion (90% within 10 days),
which suggests that preformed atrial thrombi are being
dislodged from the atrial wall due to restoration of vigor-
ous atrial contractions (49). However, transesophageal
echocardiography has shown that new atrial thrombi can
be formed after cardioversion (173), stressing a role of
prolonged depression of contractility as a factor promot-
ing thrombus formation after cardioversion (49). The loss
of atrial contractility also increases the compliance of the
fibrillating atria, which may enhance progressive dilata-
tion during AF and may contribute to further stabilization
of the arrhythmia (502). In contrast, restoration of sinus
rhythm has been shown to reduce atrial size (206, 574).
Finally, delayed recovery of atrial contraction after car-
dioversion might contribute to the delayed recovery of
exercise capacity (341).
The mechanisms responsible for atrial contractile
dysfunction following cardioversion are not completely
understood. In experimental and clinical studies, vera-
pamil was able to largely prevent the atrial dysfunction
after short periods of AF, indicating that atrial stunning is
mediated by Ca
2
overload (133, 319). While the altered
atrial function after short paroxysms of AF is likely to be
the result of changes in cellular metabolism, long-lasting
atrial tachyarrhythmias may induce additional changes
causing a more persistent atrial contractile dysfunction.
Some studies point towards an important role of I
CaL
downregulation as a cause of atrial contractile dysfunc-
tion. In dogs with sustained atrial tachycardia (6 wk), the
degree of shortening of isolated atrial myocytes was
shown to be reduced and associated with a pronounced
reduction of the Ca
2
transient (295, 547). In this model,
I
CaL
has been reported to be downregulated by 70% (647).
290 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
In goats, the time course of electrical remodeling (pre-
sumably related to I
CaL
downregulation) followed a time
course comparable to the progressive loss of atrial con-
tractility during the first days of AF, suggesting that both
phenomena are due to the same underlying mechanism
(504). Downregulation of I
CaL
directly reduces Ca
2
influx
into the cell but also shortens the atrial action potential
(211, 647), which further reduces Ca
2
entry into the cell.
Vice versa, compounds which prolong the atrial action
potential but do not directly interfere with I
CaL
have been
shown to enhance Ca
2
influx and atrial contractility
(136, 503, 608).
There are reasons to assume that in patients with
prolonged AF, other mechanisms beyond downregulation
of I
CaL
also contribute to loss of atrial contractility after
cardioversion. In such patients, recovery from electrical
remodeling occurs within a couple of days after cardio-
version, whereas restoration of atrial contraction takes
weeks to months. This indicates that I
CaL
downregulation
alone cannot explain the more persistent depression of
atrial contractile function after prolonged AF. Additional
mechanisms have been studied in right atrial trabeculae
prepared from the right atrial appendage of patients un-
dergoing mitral valve repair with and without long-stand-
ing AF (Fig. 11B) (501, 505). In patients with persistent
AF, the contractile force was reduced by 75%. Interest-
ingly, contractile reserve of these preparations and the
sarcomere content were hardly reduced (18%), indicat-
ing that the contribution of myolysis to the loss of atrial
contractile function is very limited. Also, the post-rest
potentiation was fully maintained, and diastolic proper-
ties were preserved, illustrating that the function of the
FIG. 11. Mechanisms of atrial contractile dysfunction due to AF. A: loss of the a-wave after electrical cardioversion of AF. [Modified from Logan
et al. (344), with permission from Elsevier.] B: depressed contractility in trabeculae isolated from right atrial appendage of patients undergoing mitral
valve surgery. [Modified from Schotten et al. (501).] C: confocal line scan of Ca
2
release in atrial myocytes isolated from rabbits after rapid atrial
pacing for 5 days (RAP) or sham. In RAP cells, central cellular Ca
2
release is blunted, indicating failure of intracellular Ca
2
wave propagation in
tachycardia-induced atrial remodeling. [Modified from Greiser et al. (210).] D: schematic illustration of Ca
2
handling alterations in AF. Down-
regulation of I
CaL
, upregulation of I
NCX
, and defective release of Ca
2
from the sarcoplasmic reticulum are the main mechanisms of atrial contractile
dysfunction (dark red labels) while alterations of SR Ca
2
load or myofilament function play a minor role (dark green labels).
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 291
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
sarcoplasmic reticulum was not largely altered. In con-
trast, the positive inotropic effect of isoproterenol was
markedly impaired, although the density of the
-adreno-
ceptors and the expression of the inhibitory and stimula-
tory G proteins were unaltered (501). The catecholamine-
stimulated adenylyl cyclase activity was not reduced, in-
dicating that the
-adrenergic signal transduction was not
desensitized (505). Thus, in contrast to ventricular tachy-
cardiomyopathy, which is due to a dysfunction of the
sarcoplasmic reticulum and
-adrenergic desensitization,
atrial contractile dysfunction after prolonged AF must be
due to other mechanisms.
Oxidative injury (379) as well as reduced phosphor-
ylation of myofibrillar proteins (164) have been suggested
to reduce the performance of the contractile apparatus in
patients with AF. Also, myosin binding protein C was
found to be dephosphorylated in patients with AF (164),
which would also be in line with reduced contractile
performance of the atrial myofibrils. Some studies on
skinned fibers showed that both the maximal force gen-
eration and also the Ca
2
sensitivity of the myofilaments
were unaltered (163, 406). A more recent report demon-
strated a reduction in maximum tension and the rate of
tension development, and an increase in myofilament
Ca
2
sensitivity in patients with AF (40). The extent to
which these changes contribute to loss of atrial contrac-
tility due to AF is difficult to assess but might be limited
in the light of preserved contractile reserve of intact
muscle preparations.
Recently, confocal imaging of Ca
2
release from the
sarcoplasmic reticulum showed that the centripetal Ca
2
wave propagation was significantly impaired in atrial
myocytes isolated from dogs (596a) and rabbits (210)
undergoing rapid atrial pacing for 5 days (Fig. 11C). These
studies indicate that impaired release of Ca
2
from the
sarcoplasmic reticulum might contribute to loss of atrial
contractility. The mechanism underlying impaired Ca
2
release from the sarcoplasmic reticulum is currently un-
known but not related to downregulation of I
CaL
(210).
Instead, impaired coupling between Ca
2
channels and
RYRs was recently reported in sheep with persistent AF
(322). In patients with AF, an upregulation of the Na
/
Ca
2
exchanger was reported that potentially reduces
Ca
2
load of atrial myocytes contributing to contractile
dysfunction in AF (505). The main mechanisms causing
atrial contractile dysfunction due to AF are summarized
in Figure 11D.
Atrial dilatation without concomitant AF can also
cause an atrial contractile dysfunction. In animal models
of heart failure, atrial emptying function is reduced (517),
and in patients, loss of atrioventricular synchrony due to
single chamber ventricular demand (VVI) pacing in-
creases left atrial diameter while markers of atrial con-
tractility decrease (537). In goats with atrial dilatation due
to AV block, atrial dysfunction was related to reduced
sarcoplasmic Ca
2
load due to phospholamban dephos-
phorylation and ryanodine receptor hyperphosphoryla-
tion (211). Reduced atrial contractile function was asso-
ciated with disruption of a t-tubular system in atrial myo-
cytes of sheep with heart failure (142). In dogs with
congestive heart failure, reduced atrial contractility is
associated with prolonged action potential duration, ele-
vated diastolic Ca
2
concentrations, and increased ampli-
tude of Ca
2
transients (643). These findings point to-
wards reduced force generation or Ca
2
sensitivity of the
contractile apparatus. Myosin binding protein C phos-
phorylation was reduced in both goats with atrial dilata-
tion due to AV block and in dogs with CHF. Whether this
also contributes to reduced contractile performance in
dilated atria is currently unclear.
AF may not only cause contractile dysfunction in the
atria but also in the ventricles. Persistent elevation of
ventricular rate above 130 beats/min can produce a tachy-
cardiomyopathy of the ventricles (286, 433). Reduction of
the heart rate might reverse normal pump function, em-
phasizing that ventricular rate control might not only
prevent deterioration of LV function but also restore
pump function that is already compromised (188).
D. Alterations of Atrial Ca
2
Handling and
Abnormal Impulse Formation
During the past years, alterations of intracellular
Ca
2
handling in dilated and fibrillating atria have at-
tracted the attention of many research groups. Many of
these studies indirectly suggest a role of triggered activity
in the generation of fibrillation wavefronts in patients
with AF. Vest et al. (589) demonstrated an enhanced open
probability of the RYRs in atrial myocardium of dogs
undergoing RAP as well as of patients with sustained AF,
which in both cases was due to hyperphosphorylation of
the RYR channel at the PKA site. The authors conclude
that these alterations might facilitate spontaneous Ca
2
release events from the sarcoplasmic reticulum, which in
turn might induce DADs and possibly trigger action po-
tentials. In agreement with this, Hove-Madson et al. (253)
reported an increase of the frequency of Ca
2
sparks
(elementary Ca
2
release events) in atrial myocytes iso-
lated from the right atria of patients with sustained AF.
The Ca
2
load of the sarcoplasmic reticulum was unal-
tered, which supports the hypothesis that the increase in
spontaneous Ca
2
release was due to a change of the
intrinsic properties of the RYRs. Recently, increased leak
of Ca
2
from the sarcoplasmic reticulum and elevated
diastolic Ca
2
concentrations have indeed been described
in right human atrial myocytes (409). In this study, hyper-
phosphorylation of RYRs was shown to be due to en-
hanced activity of CaMKII, as mentioned in section VA.
This interesting hypothesis, however, raises some impor-
292 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
tant questions. For instance, it is unclear why despite the
fact that in atrial myocytes isolated from the right atrial
appendages of patients with AF the Ca
2
spark frequency
is enhanced, ectopic activity originating in this region is a
rare phenomenon. Similarly, dogs after 2 mo of RAP do
not develop atrial ectopy, although the open probability of
the atrial ryanodine channels has been reported to be
increased (589). Also, increased open probability alone
only transiently enhances spontaneous Ca
2
release from
the sarcoplasmic reticulum, since increased release
would reduce the sarcoplasmic Ca
2
load and in turn
spontaneous Ca
2
release (580). However, in ventricular
myocytes of dogs with heart failure, enhanced diastolic
Ca
2
release has been observed even in the presence of
reduced Ca
2
load of the sarcoplasmic reticulum (306). In
this model, reduced Ca
2
release during the steady-state
Ca
2
transients might have prevented progressive empty-
ing of the Ca
2
stores. Another possibility is that stimu-
lation of the Ca
2
reuptake rate maintains Ca
2
load and
produces a sustained enhancement of Ca
2
release
events. Of note, phosphorylation levels of phospholam-
ban, an inhibitory protein controlling the reuptake of
Ca
2
into the sarcoplasmic reticulum by the sarcoplasmic
reticulum Ca
2
-ATPase (SERCA), is increased in atrial
myocardium of patients with AF (164). Enhanced phos-
phorylation of phospholamban dissociates the molecule
from SERCA and stimulates reuptake of Ca
2
into the
sarcoplasmic reticulum, possibly increasing Ca
2
load
and spontaneous Ca
2
release from the intracellular Ca
2
stores. So far, however, there is no evidence for enhanced
reuptake of Ca
2
into the sarcoplasmic reticulum in fib-
rillating atrial myocardium. Rather, experiments with iso-
lated right atrial trabeculae show that at any level of Ca
2
load, the diastolic properties of atrial myocardium of
patients with AF are unaltered (505). Recently, in mice
with a genetic gain-of-function defect in the RYRs, spon-
taneous AF was absent at rest but could be induced by
rapid pacing, which also resulted in increased hyperphos-
phorylation of RYRs at the CaMKII phosphorylation site
(100). Importantly, pharmacological and genetic inhibi-
tion of CaMKII prevented AF inducibility. Taken together,
these data suggest that enhanced Ca
2
leak through hy-
perphosphorylated or defective RYRs alone might not be
sufficient to produce relevant atrial ectopy. It is unknown
whether in patients, enhanced sympathetic stimulation or
simply the high rate during AF can sufficiently enhance
Ca
2
load of atrial myocytes to provoke proarrhythmic
Ca
2
spontaneous leak through hyperphosphorylated
RYRs.
The question of whether atrial pathology might en-
hance spontaneous electrical activity has been addressed
by several authors. In cats with atrial dilatation due to
spontaneously occurring cardiomyopathy, resting mem-
brane potentials have been reported to be more positive,
and action potentials showed lower amplitudes and up-
stroke velocities (60). In the presence of catecholamines,
automaticity and triggered activity were frequently found
in atrial preparations from dilated atria but only occasion-
ally occurred in normal atria. In dogs with heart failure
induced by rapid ventricular pacing, action potential du-
ration was prolonged, and frequently DADs occurred
(539). Left atrial myocytes showed increased diastolic
Ca
2
concentrations and sarcoplasmic Ca
2
loading.
Spontaneous action potentials occurred more frequently
in heart failure dogs compared with controls and were
suppressed by inhibition of SR Ca
2
release and the Na
/
Ca
2
exchanger (643). Of note, in none of the aforemen-
tioned studies (60, 539, 643) has ectopic activity or spon-
taneous onset of atrial tachyarrhythmias been reported.
The response to pacing can provide valuable infor-
mation about trigger mechanisms of AF. After treatment
with ryanodine, rabbit PVs showed pacemaker activity
that was enhanced when the preparations were tran-
siently paced at a higher rate (249). In canine PVs exposed
to pituitary adenylyl-cyclase activating polypeptide
(PACAP), which mimics neurohumoral dysbalance,
rapid atrial pacing caused ectopic beats that were cou-
pled with a shorter cycle length when the triggered
cycle length was also short (236). In dogs with CHF,
rapid pacing induced an atrial tachycardia with repetitive
radial spread of activation. The postpacing interval as
well as the cycle length of the tachycardia itself were
positively correlated with the pacing cycle lengths trigger-
ing the tachycardia (539). All these findings are consistent
with enhanced Ca
2
loading of atrial myocytes causing
spontaneous release of Ca
2
from the sarcoplasmic retic-
ulum and repetitive DADs. Reentrant mechanisms appear
less likely because here a high rate of the triggering
impulses results in a longer postpacing interval. It should
be noted, however, that DADs do not invariably react on
pacing protocols (267) and the specificity of such tests to
identify a specific proarrhythmic mechanism might be
limited (620).
Mapping of the activation pattern during or at the
onset of AF has also been used to support a role of
ectopic discharges in initiation and perpetuation of AF. Of
note, demonstration of radial spread of activation from a
localized area does not allow direct determination of the
mechanism of an arrhythmia, but to some degree sup-
ports the existence of a localized source of AF. For ex-
ample, activation mapping has revealed the onset of AF
triggered by a spontaneous beat originating from the PVs
in dogs during autonomic imbalance (236) or with con-
gestive heart failure (109, 423) and in rats during glyco-
lytic inhibition (427). In patients with paroxysmal AF,
endocardial mapping of the PVs demonstrated radial
spread of activation of spontaneous beats in the PVs
initiating runs of AF (440). It is important to note, how-
ever, that radial spread of activation can also be due to
“breakthrough” of wavefronts originating from deeper
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 293
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
layers of the atrial wall (159). Therefore, these studies do
not prove that in the studies mentioned localized sources
have induced episodes of AF.
A higher level of evidence for the presence of a
localized driver for AF comes from observations of radial
spread of activation occurring repetitively at the same
site. This conduction pattern has been described in dogs
with CHF and AF-induced by RAP (177, 653). In the study
by Fenelon et al. (177), repetitive radial spread of activa-
tion was documented in the endocardium and the epicar-
dium primarily at sites along the crista terminalis. Zhou et
al. (653) reported repetitive focal spread of activation in
the PVs of dogs with AF. Based on these activation pat-
terns, the authors of both studies cannot rule out micro-
reentry as a mechanism of repetitive radial spread of
activation. However, they did provide additional argu-
ments in favor of a cellular proarrhythmic mechanism
(see below).
Other studies interpreted an increased incidence of
radial spread of activation in structurally altered atria as
an argument for cellular proarrhythmic mechanisms in-
ducing AF. In dogs with pacing-induced heart failure,
acetylcholine increased the rate of “focal discharges”
from the PVs by factor 3 (109). Nitta el al. (419) observed
multiple “focal activations” with fibrillatory conduction in
patients with AF undergoing open chest surgery. The
mechanism of these events, however, remains unclear.
They might be due to cellular proarrhythmic activity, but
they might also be due to an increased incidence of
“breakthroughs” due to transmural conduction, which
more likely occurs in a complex substrate for AF as
explained in section VE(159).
Finally, the response of electrical activity to pharma-
cological compounds has been used as an argument in
favor of the presence of focal discharges during atrial
tachyarrhythmias. Spontaneously occurring arrhythmias
in PVs of dogs undergoing 6 8 wk of rapid atrial pacing
were suppressed by sodium channel blockers or magne-
sium (103). In canine PV, left atrial preparations exposed
to acetylcholine, thapsigargin, and ryanodine eliminated
focal discharges from the PVs (109). The atrial tachycar-
dia with repetitive radial spread of activation in dogs with
CHF was terminated by compounds reducing Ca
2
load-
ing like verapamil, flunarizine, and ryanodine (539). The
experiments with verapamil and ryanodine suggest that
Ca
2
loading might be critical for some forms of sponta-
neous electrical activity occurring in AF-related atrial
pathologies and support the role of DADs as a trigger
mechanism for spontaneous activity in the atrium. In
contrast, flunarizine has also been shown to terminate
reentrant rhythms like atrial flutter in the canine sterile
pericarditis model, illustrating that the response to fluna-
rizine cannot be taken as an argument for the existence of
Ca
2
-related cellular proarrhythmic mechanisms (591). In
sheep atria exposed to acute stretch and adrenocholin-
ergic stimulation (perfusion with high concentrations of
isoproterenol and acetylcholine), both caffeine and ryan-
odine inhibited breakthroughs showing radial spread of
activation during AF. The sensitivity of these break-
throughs to compounds which empty the sarcoplasmic
reticular Ca
2
stores might be taken as an argument in
favor of focal discharges as the underlying mechanism
under these specific conditions (639a).
In summary, the role of altered Ca
2
homeostasis for
initiation and perpetuation of AF is still obscure, mainly
due to technical limitations of existing experimental ap-
proaches. As confirmation of Ca
2
-dependent AF mecha-
nisms might have consequences for our understanding of
the working mechanisms of antiarrhythmic drugs (147), it
requires further clarification.
E. Atrial Structural Remodeling and
Conduction Disturbances
As explained in sections IV and VB, apart from elec-
trical remodeling, other factors must also contribute to
the high susceptibility to the arrhythmia in patients with
AF. This “second factor” may entail alterations in atrial
tissue structure. The relation between structural alter-
ations and AF has been studied in many different clinical
entities. Clearly, structural alterations of atria are not
exclusively related to AF but at least to the same degree
as the existence of structural heart disease present in an
individual patient. The kind of structural alterations also
shows a high diversity. For example, Frustaci and co-
workers (185, 187) have described evidence of occult
atrial pathology, such as myocyte necrosis, myocarditis,
and fibrosis, in patients with lone AF. However, the ma-
jority of chronic AF patients is of advanced age or suffers
from structural heart disease. Of the various aspects of
atrial structural remodeling in these patients, the conse-
quences of fibrosis, myocyte hypertrophy, and altered
connexin distribution have been studied most exten-
sively, as discussed below.
1. Atrial fibrosis
Atrial fibrosis is thought to be one of the most im-
portant factors in the formation of a substrate for AF (407,
529). Atrial fibrosis has been observed in biopsies from
patients with AF (301) as well as in patients with specific
risk factors for AF, such as valvular disease (14), rheu-
matic heart disease (352, 450), dilated and hypertrophic
cardiomyopathy (422), and advanced age (336).
Myocytes are organized in bundles, separated by per-
imysial fibrous tissue. Within these bundles, strands of
myocytes can be separated from each other by endomy-
sial fibrous tissue. Structural remodeling due to heart
disease is often associated with fibrosis and an increased
transverse fiber separation. In the atria, collagenous septa
294 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
between myofibers increase in volume during normal ag-
ing (304, 530). Similarly, the extracellular matrix volume
per myocyte increased in the goat of 4 mo of RAP (29).
Also in atria of animal models of atrial dilatation (59, 587)
and CHF (326), the amount of fibrosis is increased. How-
ever, in these models and in particular in the CHF model,
larger areas of fibrosis are observed, which are more
similar to “replacement fibrosis” secondary to tissue dam-
age and cell death. Various degrees and forms of atrial
fibrosis and the resulting conduction disturbances are
shown on Figure 12. Atrial fibrosis may in itself be suffi-
cient to increase AF vulnerability, as shown in mice with
selective atrial fibrosis due to overexpression of TGF-
1
(582).
Spach and Boineau (529) have demonstrated that the
nonuniform anisotropic arrangement in cell-to-cell con-
nections leads to discontinuous conduction at a micro-
scopic scale. While longitudinal propagation may still be
fast, transverse propagation may show “discontinuous
conduction,” i.e., discrete time delays in activation of
adjoining myocytes or myocyte bundles (529). During
such delays due to poor electrical coupling, propagation
FIG. 12. Relation between structural alterations and changes in conduction in animal models of AF. Fibrosis is stained in blue (Trichrome), red
(Sirius red), or light blue (Toluidine Bbue). Myocytes are stained red (Trichrome), green (Sirius red), or dark blue (Toluidine blue). A: in a canine
model of chronic atrial dilatation due to mitral insufficiency, inflammatory infiltrates and fibrosis were increased. During slow pacing, activation
wavefronts spread rapidly and homogeneously, but during extrastimulation, areas of slow conduction were observed. [Adapted from Verheule and
co-workers (586, 587).] B: cellular hypertrophy without an increase in fibrosis was reported in a goat model of chronic biatrial dilatation due to AV
block. Fast pacing (BCL 200 ms) revealed areas of slow conduction. [Adapted from Neuberger et al. (411).] C: pronounced increase in fibrosis in
a canine model of CHF due to rapid ventricular pacing was associated with areas of conduction heterogeneity in the left atrium. [Adapted from Li
et al. (326).] D: 6 mo of RAP led to increased interstitial fibrosis in the goat. During slow pacing, the conduction velocity in the left atrium was
decreased by 30%. E: mouse model of selective atrial fibrosis due to overexpression of TGF-
1. Left atrial activation maps show increased
heterogeneity of conduction during pacing at a cycle length of 150 ms (582).
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 295
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
may become increasingly dependent on I
CaL
rather than
I
Na
(514, 546). Discontinuities in conduction can become
more apparent at short cycle lengths or during extra-
stimulation with short coupling intervals due to incom-
plete recovery of Na
and Ca
2
channels (534, 546). This
may allow reentry to occur in very small circuits (529, 534,
535). Discontinuous conduction between poorly coupled
myocytes forms the basis of “electrical dissociation” dur-
ing AF that is reflected by fractionated electrograms.
Although there are strong indications from animal
models that atrial fibrosis can be proarrhythmic (80, 169),
some questions regarding the role of atrial fibrosis in the
development of a substrate of AF are so far unresolved.
First, the association between atrial fibrosis and AF
is quantitatively not very strong. Some, but not all, human
studies have found that atrial fibrosis is more pronounced
in chronic AF patients than in patients with sinus rhythm
(53, 301, 475). The degree of atrial fibrosis and fibrogenic
activity correlates with the persistence of AF (208, 635).
However, from these studies it is unclear whether in-
creased fibrosis is caused by underlying structural disease
or by AF itself. In some of these studies, the degree of the
underlying heart disease is not well documented. From a
careful comparison of structural heart disease patients
with and without AF, Anné et al. (14) have concluded that
AF itself is not associated with atrial fibrosis but is instead
related to the underlying structural heart disease.
Second, it is unclear what the quantitative relation
between atrial fibrosis and conduction disturbances is. As
discussed in section IVD, some animal models of atrial
dilation show both atrial fibrosis and conduction distur-
bances (586). However, other models show similar con-
duction disturbances in the absence of atrial fibrosis
(411).
Another question is how different types of fibrosis,
for example, large areas of replacement fibrosis (as in the
canine CHF model) or thin strands of interstitial fibrosis
(as in models of aging), affect atrial conduction. In ven-
tricular cardiomyopathy, Kawara et al. (281) have demon-
strated that conduction abnormalities strongly depend on
the pattern of fibrosis. In this study, diffuse fibrosis with
short fibrotic strands only marginally affected conduc-
tion. However, long fibrotic strands could cause pro-
nounced conduction slowing during extrastimulation, es-
pecially during transverse propagation.
Finally, the question remains how important atrial
fibrosis is as a causative factor for AF in humans. In
patients undergoing open heart surgery, the degree of
fibrosis does correlate with the occurrence of postop-
erative AF (199) and with the recurrence of AF (475).
However, both the degree of atrial fibrosis and the
occurrence of AF may reflect the severity of underlying
heart disease, without a strong, direct causal link be-
tween fibrosis and AF. Indeed, a significant degree of
atrial fibrosis can be present in patients without a
history of AF (199).
Despite the fact that the association between atrial
fibrosis and AF is surprisingly weak, experimental and
clinical studies show that prevention of atrial fibrosis
can delay the development of a substrate of AF. Several
compounds (statins, ACE inhibitors, AT
1
-receptor
blocker, fish oil, and glucocorticoids) have been proven
to effectively delay the structural remodeling process
and reduce AF stability in a variety of experimental
models (309, 318, 328, 380, 477, 478, 523). In patients,
several post hoc analyses of clinical trials and small-
scale proof-of-principle studies suggest that therapy
with ACE inhibitors, AT
1
-receptor blocker, statins, and
polyunsaturated fatty acids (PUFAs) are useful to pre-
vent the occurrence of AF. The details of these studies
are summarized in Tables 2–5. It is tempting to specu-
late that the beneficial effect reported in these studies
is due to an antifibrotic effect in the atria. However,
improvement of the patients’ hemodynamics with nor-
malization of atrial pressures might also have contrib-
uted to the beneficial effects of these compounds. The
current ESC guidelines for the management of AF con-
tain recommendations for the use of ACE inhibitors,
AT1-receptor blockers, and statins for primary and sec-
ondary prevention of AF (see section on upstream ther-
apy) (168a).
Unfortunately, compared with fibrosis, amyloidosis
(323, 466, 541) and fatty infiltrates (39), which can occur
widely in elderly patients and can potentially have a sim-
ilar impact on conduction, have received comparatively
little attention. Mechanistic studies on the role of these
structural changes in AF are so far lacking.
2. Altered connexin expression
Another relevant factor for atrial conduction may be
altered connexin expression. In the working myocardium,
conduction velocity is higher in the longitudinal than in
the transverse direction. In the transverse direction, a
propagating wavefront has to cross more cell-to-cell
boundaries within a given distance. In addition, the
smaller and sparser gap junctional plaques at side-to-side
connections represent a higher resistance than the larger
intercalated discs at end-to-end connections (531). With
age, gap junctions also become increasingly localized at
end-to-end connections between myocytes, thus further
increasing anisotropy (207, 304, 448).
Several studies have reported alterations in Cx40
and/or Cx43 in patients with AF, but the observations are
not consistent (152). In patients with chronic AF, both
higher (454) and lower (404, 617) levels of Cx40 were
reported. Another study reported lateralization of connex-
ins, with an increased heterogeneity in Cx40 distribution
and a reduction of Cx43 (301).
296 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
TABLE 2. Clinical trials of renin-angiotensin system inhibition to prevent new-onset AF in patients with CHF or AMI (primary prevention) and
patients with hypertension (primary prevention)
References Study Design Patients Intervention Primary Outcome
Pedersen et al. (445) TRACE study (post hoc analysis) 1,577 Patients with AMI and reduced
LVEF
Trandolapril versus placebo Reduced incidence of new-onset AF
Follow-up: 2–4 yr
Pizetti et al. (451) GISSI-3 study (post hoc analysis) 17,749 Patients with AMI Lisinopril versus nolisinopril No difference in new-onset AF
Follow-up: 4 yr
Vermes et al. (588) SOLVD study (post hoc analysis) 347 Patients with LV dysfunction Enalapril versus placebo Reduced incidence of new-onset AF
Follow-up: 2.9 yr
Alsheikh-Ali et al. (9) SOLVD study (post hoc analysis) 6,797 Patients with LV dysfunction Enalapril versus placebo Reduced incidence of hospitalisation
with AF
Follow-up: 34 mo
Maggioni et al. (355) Val-HeFT study (post hoc analysis) 4,395 Patients with chronic
symptomatic HF
Valsartan versus placebo Reduced incidence of new-onset AF
Follow-up: 23 mo
Ducharme et al. (150) CHARM study (prespezified secondary
end point)
6,379 Patients with symptomatic
CHF
Candesartan versus placebo Reduced incidence of new-onset AF
Follow-up: 37.7 mo
Hansson et al. (224) CAPPP study randomized, open label
(analysis on adverse event reports)
10,985 Patients with hypertension Captopril versus
diuretics beta-blocker
No difference in new-onset AF
Follow-up: 6.1 yr
Hansson et al. (223) STOP-2 study (randomized, open label) 6,628 Patients with hypertension Enalapril/lisinopril versus
CCB versus
diuretics beta-blocker
No difference in new-onset AF
Study duration: 4 yr
Lh’Allier et al. (312) Retrospective longitudinal cohort study
(from an administrative database of
8 million people in the USA)
10,926 Patients with hypertension ACEI versus CCB Reduced incidence of new-onset AF
Average follow-up: 4.5 yr
Wachtell et al. (594) LIFE study (randomized, double-blind) 8,851 Patients with
hypertensionECG LVH
Losartan versus atenolol Reduced incidence of new-onset AF
Follow-up: 4.8 yr
Salehian et al. (480) HOPE study (post hoc analysis) 8,835 Patients at high cardiovascular
risk without known heart failure
or LV systolic dysfunction
Ramipril versus placebo No difference in new-onset AF
Follow-up: 4.5 yr
Schmieder et al. (499) VALUE study (randomized,
double-blind)
13,760 Patients with hypertension at
high cardiovascular risk
Valsartan versus amlodipine Reduced incidence of new-onset AF
Follow-up: 5yr
ACEI, angiotensin-converting enzyme inhibitor; AF, atrial fibrillation; AMI, acute myocardial infarction; ARB, angiotensin receptor blockers; CCB, calcium channel blockers; CHF,
congestive heart failure; CI, confidence interval; HR, hazard ratio; LV, left ventricular; LVEF, left ventricular ejection fraction; LVH, left ventricular hypertrophy.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 297
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
In patients undergoing open-chest surgery, high Cx40
levels correlated with low conduction velocities (274) and
an increased incidence of postoperative AF (154). These
findings agree with measurements on strands of cultured
atrial myocytes from transgenic mice, where a reduction
in Cx40 led to an increased conduction velocity (38).
However, Kanagaratnam et al. (272) reported that in
chronic AF patients, lower relative Cx40 levels detected
by immunohistochemistry were associated with an in-
creased complexity of AF conduction patterns.
TABLE 3. Clinical trials of statins to prevent new-onset and progression of AF (primary and
secondary prevention)
References Study Design Patients Intervention Primary Outcome
Young-Xu et al.
(645)
Prospective observational
cohort
449 Patients with chronic stable
angina, without CHF
Statins versus no statins Reduced incidence of new-
onset AF
Average follow-up: 5 yr
Merckx et al. (377) Retrospective 218 Patients on statins and 449
matched controls in SR
without LV dysfunction
Statins versus no statins Reduced incidence of new-
onset AF
Mean follow-up: 6.5 yr
Dernellis et al. (140) Randomized, single-blind 80 Patients with proven
paroxysmal AF and
C-reactive protein levels
between 0.8 and 13 mg/l
Atorvastatin versus
placebo
Paroxysmal AF resolved in
65% of statin-treated
patients versus 10% with
placebo
Follow-up: 4–6 mo of
therapy
Amit et al. (11) Retrospective 264 Patients with permanent
pace makers
Statins versus no statins Reduced incidence of new-
onset AF
Median follow-up: 359 days
Hanna et al. (219) Retrospective 25,268 Patients with CHF Lipid-lowering drug use
versus no use
Reduced incidence of new-
onset AF
Adabag et al. (1) Retrospective 13,783 Patients with CHD Statins versus no statins No difference in new-onset
AF
Average follow-up: 4.8 yr
Ramani et al. (463) Retrospective 1,526 Patients with acute
coronary syndrome
Statins versus no statins at
time of admission
Reduced incidence of new-
onset AF
CHD, coronary heart diseases; CHF, congestive heart failure.
TABLE 4. Clinical Trials of statins to prevent recurrence of AF after thoracic surgery
References Study Design Patients Intervention Primary Outcome
Auer et al. (28) SPPAF study (prospective
cohort)
253 Patients with no CHF
or LV dysfunction
undergoing cardiac
surgery
Statin use before surgery
versus no statin use
Reduced incidence of
postoperative AF
Amar et al. (10) Prospective cohort 131 Patients undergoing
major lung or
esophageal surgery
Statin use before surgery
versus no statin use
Reduced incidence of
postoperative AF
Marin et al. (362) Prospective cohort 234 Patients undergoing
CABG
Statin use for median duration
of 31 days before surgery
versus no statin use
Reduced incidence of
postoperative AF
Patti et al. (443) ARMYDA-3 (randomized,
double–blind)
200 Patients undergoing
cardiac surgery
Atorvastatin versus placebo
starting 7 days before
surgery and continued until
hospital discharge
Reduced incidence of
postoperative AF
Chello et al. (99) Randomized, double-blind 40 Patients undergoing
CABG
Atorvastatin versus placebo
started 3 wk before surgery
Reduced incidence of
postoperative AF
Ozaydin et al. (431) Prospective cohort 362 Patients undergoing
first elective CABG
Statin use for a mean duration
of 2.7 mo before surgery
versus no statin use
Reduced incidence of
postoperative AF
Lertsburapa et al. (324) Prospective cohort from
the AFIST I, II, and III
studies
555 Patients undergoing
cardiothoracic surgery
Statin use before surgery
versus no statin use
Reduced incidence of
postoperative AF
Virani et al. (590) Retrospective study 4,044 Patients undergoing
cardiac surgery
Statin use before surgery
versus no statin use
No difference in incidence of
postoperative AF
CABG, coronary artery bypass graft; CHF, congestive heart failure; CI, confidence interval; LV, left ventricular.
298 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
Although altered connexin distribution patterns may play
a role in forming a substrate for (micro)-reentry in AF, their
exact contribution requires further investigation. Normal atrial
myocytes have a high degree of electrical coupling. The extent
to which electrical coupling has to decrease to affect wavefront
propagation is still a matter of discussion. Some information on
the relation between connexin expression and conduction ve-
locity has been obtained in transgenic mouse models. In Cx40
heterozygous mice, atrial conduction velocity was not affected,
but in Cx40 knockout mice, atrial conduction velocity was
reduced by 30% (583). The effect of reduced electrical coupling
on propagation has been investigated in detail in mathematical
models. “Coupling clamp” experiments, in which two individ-
ual atrial myocytes were coupled via a computer, have demon-
strated that a coupling conductance of 0.65 nS is sufficient for
action potential transfer between two cells (corresponding to
3– 6 gap junction channels) (605). In reality, the average cou-
pling conductance between atrial myocytes is 170 nS or larger
(584). Shaw and Rudy (514) have shown that a reduction in gap
junctional coupling can cause slow propagation with a high
safety factor (i.e., unlikely to block). Spach and Heidlage (531)
have presented a model of a two-dimensional sheet of myo-
cytes that incorporated a detailed topology of longitudinal and
transverse gap junctions between myocytes. In this model,
activation time delays were observed during transverse propa-
gation even with normal coupling conductances. The studies
on transgenic mice and mathematical models discussed above
have investigated propagation during slow pacing. It is conceiv-
able that reduced connexin expression has a more pronounced
effect during the high and irregular activation rates of AF.
In humans, Takeuchi et al. (552) have reported that
the levels of Cx40 and Cx43 were not altered in patients
with atrial dilatation or AF. However, confocal micros-
copy showed a redistribution of Cx43, with a shift toward
the periphery of intercalated discs in the hypertrophied
myocytes of dilated atria. In a modeling study, Wilders
and Jongsma (616) showed that the electrostatic interac-
tions between neighboring gap junction channels pro-
duces a decrease in the effective junctional conductance
in a gap junctional plaque. Thus the conductance of large
gap junctional plaques is lower than would be expected
based on the number of gap junction channels present.
3. Myocyte hypertrophy
Myocyte hypertrophy has been observed in animal
models of RAP (30), atrial dilatation (59, 411), and CHF
(326). In these settings, the contribution of cellular hyper-
trophy to alterations in atrial conduction is difficult to
assess. From cable theory, it might be expected that an
increase in myocyte width would lead to an increase in
conduction velocity. Indeed, in a mathematical model of
ventricular hypertrophy, an increase in macroscopic con-
duction velocity was observed (612). However, using a
more detailed model that took the nonuniform distribu-
tion of gap junction around myocytes into account, Spach
et al. (532, 533) calculated that an increase in cell size
would lead to more pronounced propagation delays be-
tween myocytes during transverse propagation. In fact,
cell size had a larger effect on the anisotropy of conduc-
tion in this model than the distribution pattern of gap
junctional plaques around the myocyte. This finding may
explain how myocyte hypertrophy in the absence of in-
creased fibrosis can also cause conduction disturbances,
as in the goat model of chronic AV block (411).
4. Atrial architecture
Some aspects of the atrial architecture itself, apart
from pathological structural changes, may play a role in
creating a substrate for AF. Based on anatomical and
TABLE 5. Clinical trials of n-3 (omega-3) polyunsaturated fatty acids to prevent AF
References Study Design Patients Intervention Primary Outcome
Mozaffarian et al. (394) Cardiovascular Health Study
(population-based,
prospective cohort)
4,815 Patients Dietary intake
assessment
Reduced incidence
of AF
Follow-up: 12 yr
Calo et al. (84) Randomized open-label 160 Patients undergoing
elective CABG
PUFAs for at least 5 days
before surgery and up
to discharge versus
control
Reduced incidence of
postoperative AF
Frost et al. (184) Danish Diet, Cancer,
and Health study
(population-based,
prospective cohort)
47,949 Patients Dietary intake
assessment
No difference in
incidence of AF
Mean follow-up: 5.4 yr
Brouwer et al. (64) Rotterdam study:
(population-based,
prospective cohort)
5,184 Patients Dietary intake
assessment
No difference in
incidence of AF
Follow-up: 6.4 yr
CABG, coronary artery bypass graft; PUFA, polyunsaturated fatty acid.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 299
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
histological studies of the atria, areas with strong prefer-
ential fiber orientation include the crista terminalis, the
bundle of Bachmann, and the area in between the PVs
(Fig. 13A). With the intrinsic anisotropy of these regions,
structural remodeling may readily lead to dissociated con-
duction patterns. Indeed, a high incidence of fractionated
electrograms has been reported in these areas (see sect.
VF). The bundle of Bachmann forms the major conduct-
ing pathway between the right and the left atrium and
consists of parallel muscle bundles (Fig. 13B). In goats
after 1 mo of AF, this structure showed a high incidence
of complex fractionated electrograms (513).
Another salient feature of atrial anatomy is the ex-
tensive trabecular network underlying a thin epicardial
layer in a major part of right and left atria (Fig. 13C).
Scheussler et al. (510) showed by endo- and epicardial
mapping in canine hearts that epicardial and endocardial
activation patterns can be markedly different and that the
epicardial layer plays a leading role in atrial wave propa-
gation during sinus rhythm. In a study on isolated sheep
right atria, Berenfeld et al. (44) have shown how the
endocardial network of trabeculae increases the com-
plexity of activation patterns during AF. Recently,
Houben et al. (252) have suggested that with a loss of
continuity in the thin epicardial layer of the atrial wall, the
trabeculated endocardial structure may become domi-
nant, resulting in a more disorganized and stable type of
AF (252). Indeed, using simulateneous endo-epicardial
high-density mapping, Eckstein et al. (159a) could show
that significant electrical dissociation occurs between the
thin epicardial layer and the endocardial bundle network,
resulting in a complex three-dimensional medium for
wavefront propagation. In this study, the degree of endo-
epicardial electrical dissociation increased with the com-
plexity of the AF substrate (159a). Also, the majority of
“breakthroughs” could be traced back to fibrillation
waves propagating on the contralateral side of the atrial
wall and therefore are likely to be due to transmural
conduction.
One of the challenges ahead is to link the various
conduction patterns during AF to the underlying tissue
architecture and pathological changes to elucidate the
electropathological substrate for perpetuation of AF.
F. Assessment of the AF Substrate by Fibrillation
Electrogram Analysis
As described earlier, alterations in myocyte electro-
physiology and tissue structure can create a substrate for
AF. From animal models and patient studies, it has become
clear that different pathological mechanisms with different
resulting substrates can cause AF. Thus more specific treat-
ment strategies require diagnostic tools that allow assess-
ment of the nature and severity of the AF substrate in
patients. A simple electrocardiographic measure, “coarse”
versus “fine” AF, does not correlate well with atrial size (54,
595), heart disease etiology (392), or average AF cycle length
(446), but may be indicative of thromboembolic risk (638).
However, the dominant atrial cycle length determined from
surface lead v1 by Fast Fourier Transform does reflect a
spatial average of the AF cycle length in the right atrium
(246, 260, 446). In patients with new-onset AF, the dominant
atrial cycle length was higher with increasing age, increased
FIG. 13. Aspects of atrial anatomy. A: the posterior left atrium,
showing the myocardial sleeves of the pulmonary veins in a human
heart. In the area between the pulmonary veins, fibers show a preferen-
tial inferior-superior orientation. [Photo modified from Saito et al. (476),
with permission from John Wiley and Sons.] B: epicardial aspect of the
bundle of Bachmann (BB) in a goat heart, the main connection between
the right and left atrium, consisting of a large number of parallel bun-
dles. Ao, aorta. (Photo by Sander Verheule.) C: endocardial aspect
anterior part of the left and right atrium in a goat heart, showing the
extensive network of endocardial trabeculae underlying the thin epicar-
dial layer. SVC, superior caval vein. (Photo by Sander Verheule.)
300 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
with AF duration, and was lower in patients who cardio-
verted spontaneously within the subsequent 24 h (259). A
new development is the use of tissue Doppler echocardiog-
raphy to determine the AF cycle length (156) and total atrial
conduction time (376). The latter parameter was predictive
of new-onset AF (139). At this point, the most extensive
information on the AF substrate in humans has come from
studies that have investigated atrial electrograms recorded
in patients with AF. The distribution of dominant frequen-
cies, fractionation of electrograms, and direct mapping of
conduction patterns are the main analysis techniques used,
as discussed below.
1. Distribution of dominant frequencies
Dominant frequency analysis is a relatively simple and
time-efficient method for determining the AF cycle length.
Recorded electrogram signals are broken down by Fast
Fourier Transformation into a number of constituent sinu-
soidal functions. Next, the magnitude of these sine waves is
plotted against the frequency in a “power spectrum.” The
dominant frequency is the highest peak in this power spec-
trum (412). In principle, a relatively high dominant fre-
quency may be caused by a rapid ectopic focus, disso-
ciated conduction, a rotor, or another form of reentrant
circuit or even an artifact. Dominant frequency analysis
often correlates well with the AF cycle length, but it is
sensitive to the signal quality, recording method (unipo-
lar vs. bipolar), and electrogram fractionation (414). In
fact, a recent study indicates that the overall correla-
tion between AF cycle length and dominant frequency
can be surprisingly poor, both for unipolar and bipolar
electrograms (168). A simulation study has shown that
spurious high dominant frequencies can be found when
the frequency and amplitude of deflections is variable
(413), as often occurs in fibrillation electrograms.
Newer forms of signal processing may improve robust-
ness of AF cycle length determination (113, 251).
Lazar et al. (314) reported that in paroxysmal AF
patients, dominant frequencies in the PV-left atrial junc-
tion were higher than in the coronary sinus and posterior
right atrium. Persistent AF patients did not show a con-
sistent dominant frequency gradients between these sites
FIG. 14. Substrate of AF in patients. A: schematic representation of high dominant frequency sites in patients with paroxysmal AF (left) and
permanent AF (right). Clustering of these sites at PV and PV-left atrial ostial sites was observed in patients with paroxysmal AF. In permanent AF
patients, high dominant frequency sites occurred throughout the atria. B: unipolar electrograms and underlying activation patterns during acute AF
in humans, recorded with a high-density array of electrodes from the right atrial free wall. Top left: four wavefronts (arrows) collide at the dashed
lines. Fractionated electrograms with short double potentials were recorded at sites of collision. Bottom left: long double potentials were recorded
along a functional line of block (thick black line). Complex fractionated potentials were recorded at pivot points (curved arrows, bottom right) and
in regions of slow conduction (crowded isochrones, top right). [Adapted from Konings et al. (299).]
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 301
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
(314). PV isolation reduced the left-to-right dominant fre-
quency gradients in paroxysmal AF (313). Similarly, Sand-
ers et al. (482) found in a more detailed analysis that high
dominant frequency sites in paroxysmal AF were prefer-
entially clustered around the PVs, whereas persistent AF
patients displayed a wider distribution of high dominant
frequency sites (Fig. 14A). Correspondingly, PV isolation
led to a significant decrease in the average atrial dominant
frequency in paroxysmal AF patients, but not in persistent
AF patients (484). Within another group of paroxysmal AF
patients, a higher degree of dominant frequency organi-
zation predicted the success of PV isolation (551). Lin et
al. (340) reported that a gradient in dominant frequency
was present from an arrhythmogenic PV or superior caval
vein site to the rest of the atrium in paroxysmal AF. Other
studies have indicated that a left to right gradient may
also be present in persistent AF patients (143, 474, 633).
Lazar et al. (313) have found that PV isolation was more
successful in persistent AF patients if a left to right dom-
inant frequency gradient was present. On the other
hand, a study by Schuessler et al. (511) has indicated
that the location of the area of highest dominant fre-
quency may be unstable, especially in persistent AF,
but also in paroxysmal AF. Nevertheless, an ablation
strategy specifically targeting high dominant frequency
sites was effective in 88% of paroxysmal and 56% of
persistent AF patients (24).
2. Relation between electrogram fractionation and
activation pattern
Structural alterations can disrupt electrical connec-
tions between muscle bundles. As a result of this, the
muscle bundles become activated out of phase during AF.
This dissociation of electrical activity is reflected by frac-
tionated electrograms. An electrogram is fractionated
when it shows more than one deflection per activation
cycle length. Fractionated electrograms may be relatively
simple with two deflections per cycle (short or long dou-
ble potentials) or more complex, in some cases displaying
“continuous electrical activity” throughout the whole cy-
cle. In all cases, fractionation reflects differences in acti-
vation time within the area sensed by the electrode. How-
ever, the degree of fractionation is not necessarily diag-
nostic for the severity of the AF substrate: long double
potentials can signify the presence of an arrhythmogenic
line of block that can cause reentrant conduction,
whereas complex fractionation may in many cases only
reflect local dyssynchroneities within broad fibrillation
waves, without effect on the overall activation pattern.
In recordings of acute AF from right atria in Wolff-
Parkinson-White patients, 23% of the fibrillation electro-
grams were fractionated, compared with 7% during sinus
rhythm (299). Of these fractionated electrograms, 0.2%
displayed complex fractionation (2 deflections) during
sinus rhythm, compared with 6% during AF. Fractionation
of unipolar fibrillation electrograms was observed in ar-
eas of slow conduction, at lines of conduction block,
around pivot points and in regions where wavefronts
collided (Fig. 14B) (299). The observation that during AF,
collision and slow conduction were associated with frac-
tionated electrograms indicates that these activation pat-
terns were associated with dyssynchronous activation in
the atrial wall. Lines of block can be either anatomical
(i.e., also present during slow pacing) or functional (i.e.,
absent at long cycle lengths, but present during fast pac-
ing). During slow pacing with propagation perpendicular
to an anatomical line of block formed by an incision,
electrograms displayed double deflections corresponding
to the activation time points at either side of the line of
block (135). However, fractionated electrograms ob-
served around functional lines of block were often com-
plex, showing numerous deflections. Slow, discontinuous
transverse conduction between fibers partially separated
by fibrotic tissue may explain why the recorded fraction-
ated electrograms were more complex than in the case of
a complete anatomical lesion.
In simulations with an atrial model, Jacquemet et al.
(263) showed that fractionated electrograms did not oc-
cur in homogeneous tissue, even with a high degree of
anisotropy. However, when abrupt changes in conductiv-
ity were introduced, fractionated electrograms were ob-
served. Similar to the correlation between activation pat-
tern and electrogram fractionation in patients (299), the
morphology and distribution of fractionated electrograms
in this model varied from beat to beat, depending on
variations in propagation direction. However, fraction-
ation may preferentially occur at certain anatomical loca-
tions. In recent clinical studies, the regional distribution
of sites with a high degree of electrogram fractionation
appeared to be stable over time (398, 470, 495).
3. Fractionated electrograms as indicators of the
AF substrate
In chronic AF or atria of hearts with underlying heart
disease, the slow process of structural remodeling may
increase transverse fiber separation and may reduce the
number of side-to-side connections throughout the free
walls. As a result, fractionated electrograms would be-
come more frequent and more widespread. Several stud-
ies indicate that the degree of electrogram fractionation
does indeed correlate with AF maintenance. During ex-
trastimulation, electrogram fractionation has a longer du-
ration in paroxysmal AF patients than in healthy control
subjects (548), and the prevalence of fractionated electro-
grams (mainly double potentials) in the coronary sinus
musculature (279) and PVs (234) is higher than in patients
without AF. In paroxysmal AF patients, a more extensive
distribution of fractionated electrograms is predictive for
302 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
a transition to persistent AF (403), and fractionation is
more localized to the PV region. In contrast, their occur-
rence is more widespread in persistent AF (339, 632).
Some known risk factors for AF are also associated with
increased electrogram fractionation. Compared with age-
matched controls, CHF patients showed both a higher
incidence of double potentials along the crista terminalis
and an increased AF vulnerability and stability (483). In
addition, the incidence of electrogram fractionation in-
creases with age (93, 293).
A high degree of fractionation may also be a direct
result of a high activation rate. Rostock et al. (469) dem-
onstrated that local AF cycle length shortening preceded
the development of electrogram fractionation. In parox-
ysmal AF, fractionated electrograms tend to be clustered
at the PVs, which often act as high-frequency drivers of
AF in those patients, whereas fractionated electrograms
show a wider distribution in chronic AF patients (339,
543, 632). In addition, electrogram fractionation is ob-
served in close proximity to sites with a high dominant
frequency (543).
Several studies indicate that fractionation is linked to
parasympathetic activity. In paroxysmal AF patients, elec-
trogram fractionation was associated with vagal activity,
an effect that could be mimicked by adenosine adminis-
tration (320). In addition, fractionation preferentially oc-
curred near ganglionated plexi (278). In dogs, acetylcho-
line injection into ganglionated plexi led to local high
activation frequencies with irregular fractionated electro-
grams (350). As noted above, the occurrence of fraction-
ation may be correlated with the activation frequency.
Therefore, the association between vagal activity and
electrogram fractionation may be caused by a (local)
acceleration of the fibrillatory rate.
4. Ablation of fractionated electrograms
Because of its association with complex activation
patterns mapping of electrogram fractionation during AF
may allow identification of regions with structural con-
duction disturbances that are involved in perpetuation of
AF. Nademanee and co-workers (398, 399) have studied
fractionated electrograms in patients with paroxysmal
and chronic AF and frequently localized “complex frac-
tionated atrial electrograms” (CFAE) in the interatrial
septum, PVs and the left atrial roof (398, 399). Ablation of
areas with these CFAEs resulted in restoration of sinus
rhythm in 91% of the patients after a follow-up of 1 yr.
Other groups have used similar electrogram character-
istics to search for underlying local left atrial parasym-
pathetic innervation (“ganlionated plexus”), and asso-
ciated the perceived benefit of CFAE ablation with
parasympathetic denervation (250, 338). More recent
controlled trials have not been able to reproduce the
findings of Nademanee, and even suggest that CFAE
ablation does not alter the natural time course of AF
(428). Also, the benefit of ablation of CFAEs alone or as
an adjunct to PV isolation appears to be limited (428).
However, more specific electrogram-guided strategies,
such as ablation of electrograms with continuous activity
or a high gradient in activation rate, may improve efficacy
(550). Recent advances of the field of ablation of fraction-
ated electrograms have been expertly reviewed elsewhere
(397, 630) and are beyond the focus of this article.
5. Mapping of AF conduction patterns
Although the distribution of dominant frequencies
and fractionated electrograms can provide valuable infor-
mation on the AF substrate, direct mapping of activation
patterns using contact electrograms provides the most
direct electrophysiological information on the AF sub-
strate. Bipolar electrograms represent a convenient way
to reduce noise and far-field (ventricular) potentials, but
the morphology of a bipolar electrogram is sensitive to
the propagation direction and is more difficult to interpret
in case of dissociated or fractionated signals. Unipolar
electrograms allow a more accurate determination of lo-
cal activation times and are therefore preferable for a
detailed reconstruction of activation patterns. One of the
first studies to describe AF conduction patterns investi-
gated the right atrium during electrically induced AF in
patients undergoing surgery for WPW syndrome (298).
During sinus rhythm and atrial pacing, right atrial conduc-
tion in WPW patients is mostly homogeneous and isotro-
pic (222). During electrically induced AF, right atrial con-
duction showed varying degrees of complexity (298).
With increasing complexity, the number of wavefronts
simultaneously present underneath the mapping elec-
trode frequency and irregularity of the AF cycle length
increased. Clearly, complex AF was driven by more and
narrower waves. Cox et al. (129) studied WPW patients
with paroxysmal AF and described conduction of multi-
ple wavefronts around areas of conduction block in both
atria, in some cases with a single reentrant circuit be-
tween the caval veins in the RA. In chronic AF patients
undergoing mitral valve surgery, several studies reported
rapid repetitive left atrial activation with more complex
fibrillatory conduction in the right atrium, consistent with
left atrial driver regions (225, 419, 544). Similar rapid
repetitive activation was recorded in the left atrial poste-
rior wall of a varied group of chronic AF patients with
structural heart disease, but with more uniform and
broader wavefronts in the right atrium (633). Kanagarat-
nam and co-workers (272, 273) did not detect differences
in the complexity (again quantified by the number of
wavefronts simultaneously present) of right atrial con-
duction during acute and chronic AF. However, in an-
other study in patients with persistent AF, a high degree
of variability in AF wavefront propagation was observed
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 303
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
throughout the right atrium (464). Recently, Allessie et al.
(7a) have described a new, more quantitative, method for
analyzing fibrillation waves, showing an increase in lon-
gitudinal dissociation between waves in chronic AF com-
pared with acute AF. The PV region showed a higher
degree of dissociation than the RA free wall. Longitudinal
dissociation resulted in narrower wavefronts and a higher
number of simultaneous fibrillation waves. A variety of
right atrial activation patterns in chronic AF were re-
ported by Holm et al. (245), with multiple coexisting
fibrillation waves in some patients and repetitive focal spread
of activation in others, most likely reflecting transmural con-
duction of fibrillation waves (breakthroughs). De Groot et al.
(135a) have recently demonstrated that the incidence of “epi-
cardial breakthroughs” is higher in chronic than in acute AF.
The origin of these waves was nonrepetitive in time and wide-
spread throughout the atria. Based on these and other charac-
teristics, the authors have argued that most breakthroughs are
caused by transmural conduction.
Several studies have investigated conduction pat-
terns in arrhythmogenic thoracic veins of paroxysmal AF
patients. In atrial tachycardia originating in the superior
caval vein, Shah et al. (512) observed slow and anisotro-
pic conduction within the myocardial sleeve, compatible
with a tendency for local reentry. Similarly, mapping with
a 64-electrode basket catheter in the PVs found relatively
fast, narrow activation waves along the length axis of the
vein and slow conduction perpendicular to it (481). The
presence of this dissociated conduction patterns was pre-
dictive for the efficacy of ablation at that vein. In contrast,
using similar recording methods, Arentz et al. (19) ob-
served predominantly radial spread of activation without
indications for local reentrant circuits in the PV myocar-
dial sleeves. However, breakthrough patterns at the PV-
left atrial junction, leading to focal spread of activation
were also observed during sinus rhythm and atrial pacing
(137). In the latter study, mapping during ectopic activity
often showed a “multifocal” pattern of activation within
the PV area mapped with a basket catheter.
Overall, detailed studies on atrial conduction pat-
terns at various stages of the development of an AF
substrate are disappointingly rare, probably due to the
limited opportunity for mapping studies in patients and
the complexity of the time-consuming techniques in-
volved. Automated and rapid analysis of fibrillation
electrograms might enable objective and prompt anal-
ysis of activation patterns during AF in the future (251).
Another interesting new possibility is the recording of
atrial activation patterns by ECG imaging (ECGI). This
approach addresses the inverse problem by integrating
body surface mapping with detailed anatomical data
from magnetic resonance imaging (461). ECGI can de-
lineate atrial activation patterns during sinus rhythm
(462), atrial pacing (385), atrial flutter (604), and focal
atrial tachycardia (603). Recently, the first study char-
acterizing epicardial activation patterns during AF us-
ing ECGI was published (130a). The results show a
large variability in the complexity of AF increasing with
longer duration of the arrhythmia. It will be important
to validate the ability of this technique to image the
fibrillation process up to a level of complexity which
enables a clinically meaningful classification of AF.
VI. SPECIFIC FORMS OF ATRIAL
FIBRILLATION
A. Postoperative AF
In up to 30% of patients who undergo open heart
surgery, usually associated with the use of cardiopulmo-
nary bypass, AF occurs during the first days after the
operation (538, 609). The most important factor that de-
termines the incidence of postoperative AF is the type of
surgery (8, 27). The highest incidence of postoperative AF
occurs with mitral valve repair in single cardiac surgery
and with coronary artery bypass graft surgery plus mitral
valve repair in combined cardiac surgery.
Postoperative AF differs in some respects from other clin-
ical forms of the arrhythmia. The insertion of a cannula in the
right atrium and, depending on the type of operation, other cuts
in the atrial walls, and compromised hemodynamics are poten-
tial external stressors that may promote postoperative AF.
Importantly, the use of anti-inflammatory agents may prevent
postoperative AF (34, 218, 611), suggesting an important role of
systemic or local inflammatory processes. It is known that
cardiac surgery causes a biphasic complement activation that
is associated with a higher incidence of the arrhythmia (65).
Furthermore, during cardiopulmonary bypass, free radicals are
produced (116) and atrial ischemia due to inadequate atrial
protection during cardiac arrest occurs. However, several stud-
ies failed to show a lower incidence of the arrhythmia with
off-pump versus on-pump surgery (1, 513, 547, 636), question-
ing the importance of cardiopulmonary bypass. Finally, sympa-
thetic activation appears to promote postoperative AF. In af-
flicted patients, norepinephrine levels are elevated, and preop-
erative and postoperative use of
-adrenoceptor blockers
decreases the incidence of postoperative AF (271, 610).
On the other hand, postoperative AF appears to share
some pathophysiological mechanisms with other forms of
AF. For example, the duration of the preoperative P-wave
on the surface ECG does predict the occurrence of post-
operative AF (540). Interestingly, the P-wave duration
correlates with the amount of interstitial fibrosis, which
predicts the occurrence of postoperative AF supporting
the concept that fibrosis is also involved in the pathogen-
esis of this type of AF (199). Advanced age is also a
known risk factor for the development of postoperative
AF, which can be explained by age-related structural
changes like fibrosis and atrial dilatation. Like in other
304 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
forms of AF, there is also emerging evidence that oxida-
tive stress contributes to the vulnerability to postopera-
tive AF (91, 248). Also, complex changes in ionic channels
and gap junctions might enhance the propensity for post-
operative AF (154, 273, 348, 576).
B. Inherited Cardiomyopathies and Genetic
Defects Associated With AF
Numerous inherited syndromes associated with AF have
been identified during the past years (Table 6). The mechanistic
link to AF is not always obvious and sometimes surprising. For
example, patients with an inherited prolongation of the atrial
action potential carry a risk for AF (41, 266, 351, 651). Most
patients show short episodes of paroxysmal AF, and some data
may suggest that these arrhythmias can be initiated by EADs
(290, 291). Both short and long QT syndromes are associated
with AF, indicating that channelopathies can predispose to the
arrhythmia by multiple mechanisms. Also the Brugada syn-
drome is associated with supraventricular arrhythmias, often
including AF (158, 311, 497). Whether Na
channel mutations,
usually loss-of-function mutations, provoke AF in these pa-
tients is not known. Because of the large genetic heterogene-
ity of Brugada syndrome, a clear cause and effect anal-
ysis is challenging. Finally, in some, but not all, patients
with catecholaminergic polymorphic ventricular tachy-
cardia caused by genetic defects of RYR2, sinus node
and AV-nodal dysfunction and AF occur (47).
In addition to these channelopathies, several other
inherited cardiomyopathies are strongly associated with
AF of early onset (Table 6). Hypertrophic cardiomyopathy
is a well-known example for this association. A familial
form of ventricular preexcitation and abnormal LV hypertrophy
associated with mutations in the PRKAG gene is also often
associated with AF. These associations may suggest that ab-
normal ventricular hypertrophy can promote AF, possibly
through cardiomyocyte dysarray, diastolic dysfunction, or met-
abolic deficiency (345). Other genetic defects associated with
AF comprise mutations in the atrial natriuretic peptide and
changes in transcription factors (AF in Holt-Oram syndrome
and the association of PITX2 and AF) (213, 456). The patho-
physiological role of these common genetic variants for the
initiation and perpetuation of AF certainly warrants further
investigation.
VII. SUMMARY AND FUTURE PERSPECTIVES
A. Summary
Figure 15 provides an overview of the main mecha-
nisms involved in initiation and perpetuation of AF as
described in sections IV–VI. Figure 15 emphasizes the
TABLE 6. Genetic abnormalities associated with AF identified in patients with inherited cardiomyopathies
carrying a high risk for AF and genetic defects found in association with AF
Cardiac Abnormality/Type of AF Genetic Defect
AF Prevalence (Estimate)/Associated
With AF In
Inherited cardiomyopathies (366) associated with AF
Brugada syndrome Loss-of-function SCN5A mutations (10–15% of
patients)
10–20% (158)
Long QT syndrome Late gain-of-function SCN5A and loss-of-function
K channel mutations, among others
5–10% (266, 290, 291, 651)
Short QT syndrome Gain-of-function K channel mutations 70% (191, 196)
Catecholaminergic VT Loss-of-function ryanodine receptor mutation Rare families (47)
Hypertrophic cardiomyopathy Sarcomeric proteins 5–15% (317, 347, 365)
Wolff-Parkinson-White syndrome
and abnormal LVH
PRKAG mutations Rare familial forms (18, 204)
Holt-Oram syndrome with AF TBX5 mutations (regulatory gene) Family clusters (456)
Gene defects associated with AF
“Lone” AF Loss-of-function SCN5A mutations 5% of “lone” AF patients (134, 167)
AF and heart failure SCN5A mutation Rare forms of AF (425)
“Lone” AF Gain-of-function K channel mutations Rare families with AF and short QT
interval (102)
“Lone” AF Loss-of-function K channel polymorphisms Rare families, associated with long QT
syndrome (524)
“Lone” AF Loss-of-function KV1.5 mutation (I
Kur
)Rare patients (524)
“Lone” AF Somatic connexin40 mutations “Lone AF” patients (203) (requires
atrial tissue for testing)
“Lone” AF Frameshift (loss-of-function) ANP mutation Large families (242)
All types of AF PITX2 polymorphism (involved in pulmonary
and cardiac development)
Populations in Iceland (213)
Reference numbers are given in parentheses. [From Kirchhoff et al. (289).]
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 305
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
dynamic character of the process and the diversity of the
contributors to the atrial remodeling process. The dia-
gram mainly consists of four positive-feedback loops. The
central element in all loops is the arrhythmia itself, which at the
same time represents both trigger and effect of the four circular
processes. The circular positive-feedback enhancement of
these pathophysiological changes explains the general ten-
dency of AF to become more stable with time. Also, many of
the mechanisms synergistically interact once AF has become
manifest in a patient. Since the contribution of these different
factors to AF varies from patient to patient, the progression of
AF shows a high inter-individual variability. Initially, pathophys-
iological factors like structural heart diseases, arrhythmias, ag-
ing, or inherited diseases are required to drive the positive
feedback loops, but once the pathophysiological alterations in
the atria have reached a certain threshold, the process will
sustain itself and AF will become more stable over time.
B. Current Challenges and Future Perspectives
The increase in life expectancy and recent improve-
ments in treatment of acute heart disease have resulted in
FIG. 15 . Overview of mechanisms of AF. Four different positive-feedback loops are proposed as the main driving forces for the atrial remodeling process.
Enhanced Ca
2
loading during AF is believed to underlie most of the cellular proarrhythmic mechanisms (trigger loop). The main process in the electrical loop
is an altered contribution of ion channels to the action potential configuration that protects atrial myocytes against excessive Ca
2
loading. Abbreviation of the
action potential facilitates reentry and thereby promotes AF. In the structural loop, chronic atrial stretch activates numerous signaling cascades that produce
alterations of the extracellular matrix and conduction disturbances, also facilitating reentrant mechanisms. The main changes of the contractile properties of the
heart are loss of atrial contractility which increases atrial compliance and the development of a ventricular tachycardiomyopathy, both of which increase stretch
in the atrial wall. The circular positive-feedback enhancement of these pathophysiological changes explains the general tendency of AF to become more stable
with time. It should be noted that the different loops are interconnected by mechanisms that are part of more than one loop. For example, increased Ca
2
loading
enhances trigger activity (trigger loop) and also results in a change in the ion channel population and activity (electrical loop). Reentrant mechanisms are
promoted by both shortening of refractoriness (electrical loop) as well as by conduction disturbances resulting from tissue fibrosis (structural loop). Like in a
system of meshing gear wheels, one loop will drive the other, leading to progression of the arrhythmia. However, the proposed system of gear wheels does not
start to move spontaneously. Structural heart diseases, arrhythmias, aging, or inherited diseases are required to initiate movement of one or more of these wheels.
When the pathophysiological alterations eventually reach a certain threshold, AF will ensue.
306 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
a major increase in number of patients living with heart
failure and/or AF. Although the socioeconomic burden of
AF is growing steadily and significant progress has been
made in understanding the pathophysiology of this ar-
rhythmia, treatment of AF patients is still far from satis-
factory. The success rate of pharmacological cardiover-
sion is still limited, and antiarrhythmic drugs are unable
to prevent recurrences of AF (86, 122, 488). Prevention of
thromboembolic events still requires potentially harmful
anticoagulation therapy (256, 599). Radiofrequency abla-
tion, originally developed for treatment of paroxysmal
AF, varies in its efficacy to cure persistent AF and is
afflicted with a number of potentially serious side effects
(180, 400, 457, 458, 650).
In the authors’ view, addressing the following re-
search objectives might help to improve therapeutic ap-
proaches for AF.
Mechanisms contributing to the initiation and perpet-
uation of AF show a high diversity and interindividual
variability. Functional and structural changes in the atria
can promote AF by a variety of processes that may in-
volve ion channel remodeling, atrial fibrosis, inflamma-
tion, apoptosis, loss of cell-cell contacts, altered auto-
nomic tone, cellular hypertrophy, and deposition of amy-
loid. Better understanding of the factors that initiate and
maintain AF in specific patients or patient groups will not
only allow to chose appropriate therapeutic approaches,
but also help to limit therapy duration to periods when
these therapies are really beneficial.
To develop individualized therapy for AF, a classifi-
cation of the arrhythmia based on the pathophysiological
changes in the atria needs to be developed. Current ther-
apeutic regimes are most often chosen based on clinical
symptoms and the duration of AF (paroxysmal or persis-
tent AF). These categories, though helpful, do not neces-
sarily reflect the nature and degree of electrophysiologi-
cal changes resulting in AF. New diagnostic tools for AF
classification might include invasive and noninvasive
electrophysiological measurements, biochemical mark-
ers, and imaging techniques.
The structural alterations that progressively increase AF
susceptibility occur relatively early during atrial remodeling, in
most cases before the first onset of AF, and certainly before the
arrhythmia becomes persistent. Identification of patients in
early stages of this remodeling process might enable timely and
effective preventive therapy.
The development of new upstream therapy targets
requires the identification of relevant microstructural de-
terminants of conduction disturbances. Most importantly,
the exact qualitative and quantitative relation between
atrial fibrosis and conduction disturbances needs to be
determined.
The role of ectopic focal discharges in the perpetuation of
persistent AF is largely unknown. Clarification of mechanisms
and contribution of abnormal impulse formation to the perpet-
uation of persistent AF might offer the opportunity to identify
new pharmacological targets for AF therapy.
Given the complexity of the mechanisms causing AF
and the significant contribution of nonmodifiable factors
such as aging and genetic predispositions, we believe that
even an early, aggressive, and individualized therapy will
not prevent or “cure” AF in the foreseeable future. How-
ever, addressing the above-mentioned research objectives
might allow the postponing of the time point to accept AF
in an individual patient. In other cases, we might, based
on pathophysiological insights, prefer to refrain from po-
tentially harmful interventions.
ACKNOWLEDGMENTS
U. Schotten and S. Verheule contributed equally to this work.
Address for reprint requests and other correspondence:
U. Schotten, Dept of Physiology, Univ. of Maastricht, PO Box
616, 6200 MD Maastricht, The Netherlands (e-mail: Schotten
@maastrichtuniversity.nl).
GRANTS
This work was supported by the Dutch Research Organi-
zation (NWO, VIDI-Grant 016.086.379), the Foundation Leducq
(07 CVD 03), and the German Federal Ministry of Education and
Research through the Atrial Fibrillation Competence Network
(01Gi0204), and the European Union (European Network for
Translational Network in Atrial Fibrillation).
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared
by the author(s).
REFERENCES
1. Adabag ASND, Bloomfield HE. Effects of statin therapy on pre-
venting atrial fibrillation in coronary disease and heart failure. Am
Heart J 154: 1140 –1145, 2007.
2. Adam ONH, Böhm M, Laufs U. Prevention of atrial fibrillation
with 3-hydroxy-3-methylglutaryl coenzyme A reductase inhibitors.
Circulation 118: 1285–1293, 2008.
3. AFFIRM I. A comparison of rate control and rhythm control in
patients with atrial fibrillation. N Engl J Med 347: 1825–1833, 2002.
4. Alessi R, Nusynowitz M, Abildskov JA, Moe GK. Nonuniform
distribution of vagal effects on the atrial refractory period. Am J
Physiol 194: 406– 410, 1958.
5. Allessie MA, Bonke FI, Schopman FJ. Circus movement in
rabbit atrial muscle as a mechanism of tachycardia. III. The “lead-
ing circle” concept: a new model of circus movement in cardiac
tissue without the involvement of an anatomical obstacle. Circ Res
41: 9 –18, 1977.
6. Allessie MA, Bonke FI, Schopman FJ. Circus movement in
rabbit atrial muscle as a mechanism of trachycardia. Circ Res 33:
54 – 62, 1973.
7. Allessie MA, Lammers WJEP, Bonke FIM, Hollen J. Experi-
mental evaluation of Moe’s multiple wavelet hypothesis of atrial
fibrillation. In: Cardiac Arrhythmias, edited by DP Zipes, Jalife J.
New York: Grune & Stratton, 1985, p. 265–276.
7a.Allessie MA, de Groot NM, Houben RP, Schotten U, Boersma
E, Smeets JL, Crijns HJ. The electropathological substrate of
longstanding persistent atrial fibrillation in patients with structural
heart disease: longitudinal dissociation. Circ Arrhythm Electro-
physiol. 2010 Aug 18. [Epub ahead of print]
8. Almassi GH, Schowalter T, Nicolosi AC, Aggarwal A, Moritz
TE, Henderson WG, Tarazi R, Shroyer AL, Sethi GK, Grover
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 307
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
FL, Hammermeister KE. Atrial fibrillation after cardiac surgery: a
major morbid event? Ann Surg 226: 501–511, 1997.
9. Alsheikh-Ali AAWP, Rand W, Konstam MA, Homoud MK, Link
MS, Estes NA 3rd, Salem DN, Al-Ahmad AM. Enalapril treat-
ment and hospitalization with atrial tachyarrhythmias in patients
with left ventricular dysfunction.. Am Heart J 147: 1061–1065, 2004.
10. Amar DZH, Heerdt PM, Park B, Fleisher M, Thaler HT. Statin
use is associated with a reduction in atrial fibrillation after noncar-
diac thoracic surgery independent of C-reactive protein. Chest 128:
3421–3427, 2005.
11. Amit GKA, Bar-On S, Gilutz H, Wagshal A, Ilia R, Henkin Y.
Association of statin therapy and the risk of atrial fibrillation in
patients with a permanent pacemaker. Clin Cardiol 29: 249–252,
2006.
12. Anderson RH, Ho SY. “Specialized” conducting cells in the pul-
monary veins. J Cardiovasc Electrophysiol 15: 121–123, 2004.
13. Anne W, Willems R, Holemans P, Beckers F, Roskams T,
Lenaerts I, Ector H, Heidbuchel H. Self-terminating AF depends
on electrical remodeling while persistent AF depends on additional
structural changes in a rapid atrially paced sheep model. J Mol Cell
Cardiol 43: 148 –158, 2007.
14. Anne W, Willems R, Roskams T, Sergeant P, Herijgers P,
Holemans P, Ector H, Heidbuchel H. Matrix metalloproteinases
and atrial remodeling in patients with mitral valve disease and
atrial fibrillation. Cardiovasc Res 67: 655– 666, 2005.
15. Antman EM, Beamer AD, Cantillon C, McGowan N, Goldman
L, Friedman PL. Long-term oral propafenone therapy for suppres-
sion of refractory symptomatic atrial fibrillation and atrial flutter. J
Am Coll Cardiol 12: 1005–1011, 1988.
16. Anyukhovsky EP, Sosunov EA, Chandra P, Rosen TS, Boyden
PA, Danilo P, Rosen MR. Age-associated changes in electrophysi-
ologic remodeling: a potential contributor to initiation of atrial
fibrillation. Cardiovasc Res 66: 353–363, 2005.
17. Anyukhovsky EP, Sosunov EA, Plotnikov A, Gainullin RZ,
Jhang JS, Marboe CC, Rosen MR. Cellular electrophysiologic
properties of old canine atria provide a substrate for arrhythmo-
genesis. Cardiovasc Res 54: 462– 469, 2002.
18. Arad M, Moskowitz IP, Patel VV, Ahmad F, Perez-Atayde AR,
Sawyer DB, Walter M, Li GH, Burgon PG, Maguire CT, Staple-
ton D, Schmitt JP, Guo XX, Pizard A, Kupershmidt S, Roden
DM, Berul CI, Seidman CE, Seidman JG. Transgenic mice over-
expressing mutant PRKAG2 define the cause of Wolff-Parkinson-
White syndrome in glycogen storage cardiomyopathy. Circulation
107: 2850 –2856, 2003.
19. Arentz T, Haegeli L, Sanders P, Weber R, Neumann FJ,
Kalusche D, Haissaguerre M. High-density mapping of sponta-
neous pulmonary vein activity initiating atrial fibrillation in hu-
mans. J Cardiovasc Electrophysiol 18: 31–38, 2007.
20. Arndt MLU, Röcken C, Nepple K, Zahn C, Huth C, Ansorge S,
Klein HU, Goette A. Altered expression of ADAMs (a disintegrin
and metalloproteinase) in fibrillating human atria. Circulation 105:
720 –725, 2002.
21. Aronson RS, Cranefield PF, Wit AL. The effects of caffeine and
ryanodine on the electrical activity of the canine coronary sinus. J
Physiol 368: 593– 610, 1985.
22. Arora R, Ulphani JS, Villuendas R, Ng J, Harvey L, Thordson
S, Inderyas F, Lu Y, Gordon D, Denes P, Greene R, Crawford
S, Decker R, Morris A, Goldberger J, Kadish AH. Neural sub-
strate for atrial fibrillation: implications for targeted parasympa-
thetic blockade in the posterior left atrium. Am J Physiol Heart
Circ Physiol 294: H134 –H144, 2008.
23. Arora R, Verheule S, Scott L, Navarrete A, Katari V, Wilson E,
Vaz D, Olgin JE. Arrhythmogenic substrate of the pulmonary
veins assessed by high-resolution optical mapping. Circulation
107: 1816 –1821, 2003.
24. Atienza F, Almendral J, Jalife J, Zlochiver S, Ploutz-Snyder
R, Torrecilla EG, Arenal A, Kalifa J, Fernandez-Aviles F,
Berenfeld O. Real-time dominant frequency mapping and ablation
of dominant frequency sites in atrial fibrillation with left-to-right
frequency gradients predicts long-term maintenance of sinus
rhythm. Heart Rhythm 6: 33– 40, 2009.
25. Atienza F, Almendral J, Moreno J, Vaidyanathan R, Talka-
chou A, Kalifa J, Arenal A, Villacastin JP, Torrecilla EG,
Sanchez A, Ploutz-Snyder R, Jalife J, Berenfeld O. Activation
of inward rectifier potassium channels accelerates atrial fibrillation
in humans: evidence for a reentrant mechanism. Circulation 114:
2434 –2442, 2006.
26. Attuel P, Childers R, Cauchemez B, Poveda J, Mugica J,
Coumel P. Failure in the rate adaptation of the atrial refractory
period: its relationship to vulnerability. Int J Cardiol 2: 179–197,
1982.
27. Auer J, Weber T, Berent R, Ng CK, Lamm G, Eber B. Risk
factors of postoperative atrial fibrillation after cardiac surgery. J
Cardiac Surg 20: 425– 431, 2005.
28. Auer JWT, Berent R, Lamm G, Ng CK, Hartl P, Strasser U,
Eber B. Use of HMG-coenzyme a-reductase inhibitor (statin) and
risk reduction of atrial fibrillation after cardiac surgery: results of
the SPPAF study, a randomised placebo-controlled trial (Abstract).
Eur heart J 25 Suppl: 353, 2004.
29. Ausma J, van der Velden HM, Lenders MH, van Ankeren EP,
Jongsma HJ, Ramaekers F, Borgers M, Allessie M. Reverse
structural and gap-junctional remodeling after prolonged atrial fi-
brillation in the goat. Circulation 107: 2051–2058, 2003.
30. Ausma J, Wijffels M, Thone F, Wouters L, Allessie M, Borgers
M. Structural changes of atrial myocardium due to sustained atrial
fibrillation in the goat. Circulation 96: 3157–3163, 1997.
31. Avitall B, Bi J, Mykytsey A, Chicos A. Atrial and ventricular
fibrosis induced by atrial fibrillation: evidence to support early
rhythm control. Heart Rhythm 5: 839– 845, 2008.
32. Ayettey AS, Navaratnam V. The T-tubule system in the special-
ized and general myocardium of the rat. J Anat 127: 125–140, 1978.
33. Baba S, Dun W, Hirose M, Boyden PA. Sodium current function
in adult and aged canine atrial cells. Am J Physiol Heart Circ
Physiol 291: H756 –H761, 2006.
34. Baker WL, White CM, Kluger J, Denowitz A, Konecny CP,
Coleman CI. Effect of perioperative corticosteroid use on the
incidence of postcardiothoracic surgery atrial fibrillation and
length of stay. Heart Rhythm 4: 461– 468, 2007.
35. Balana B, Dobrev D, Wettwer E, Christ T, Knaut M, Ravens U.
Decreased ATP-sensitive K
current density during chronic human
atrial fibrillation. J Mol Cell Cardiol 35: 1399 –1405, 2003.
36. Bassani JW, Bassani RA, Bers DM. Relaxation in rabbit and rat
cardiac cells: species-dependent differences in cellular mecha-
nisms. J Physiol 476: 279 –293, 1994.
37. Bassett AL, Fenoglio JJ Jr, Wit AL, Myerburg RJ, Gelband H.
Electrophysiologic and ultrastructural characteristics of the canine
tricuspid valve. Am J Physiol 230: 1366 –1373, 1976.
38. Beauchamp P, Yamada KA, Baertschi AJ, Green K, Kanter
EM, Saffitz JE, Kleber AG. Relative contributions of connexins
40 and 43 to atrial impulse propagation in synthetic strands of
neonatal and fetal murine cardiomyocytes. Circ Res 99: 1216 –1224,
2006.
39. Becker AE. How structurally normal are human atria in patients
with atrial fibrillation? Heart Rhythm 1: 627– 631, 2004.
40. Belus A, Piroddi N, Ferrantini C, Tesi C, Cazorla O, Toniolo
L, Drost M, Mearini G, Carrier L, Rossi A, Mugelli A, Cerbai
E, van der Velden J, Poggesi C. Effects of chronic atrial fibrilla-
tion on active and passive force generation in human atrial myofi-
brils. Circ Res 107: 144 –152, 2010.
41. Benito B, Brugada R, Perich RM, Lizotte E, Cinca J, Mont L,
Berruezo A, Tolosana JM, Freixa X, Brugada P, Brugada J. A
mutation in the sodium channel is responsible for the association
of long QT syndrome and familial atrial fibrillation. Heart Rhythm
5: 1434 –1440, 2008.
42. Benjamin EJ, Levy D, Vaziri SM, D’Agostino RB, Belanger AJ,
Wolf PA. Independent risk factors for atrial fibrillation in a popu-
lation-based cohort. The Framingham Heart Study. JAMA 271:
840 – 844, 1994.
43. Benjamin EJ, Wolf PA, D’Agostino RB, Silbershatz H, Kannel
WB, Levy D. Impact of atrial fibrillation on the risk of death: the
Framingham Heart Study. Circulation 98: 946 –952, 1998.
44. Berenfeld O, Zaitsev A, Mironov S, Pertsov AM, Jalife J.
Frequency-dependent breakdown of wave propagation into fibril-
latory conduction across the pectinate muscle network in the
isolated sheep right atrium. Circ Res 90: 1173–1180, 2002.
308 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
45. Berger M, Schweitzer P. Timing of thromboembolic events after
electrical cardioversion of atrial fibrillation or flutter: a retrospec-
tive analysis. Am J Cardiol 82: 1545–1547, 1998.
46. Bers DM. Calcium cycling and signaling in cardiac myocytes.
Annu Rev Physiol 70: 23– 49, 2008.
47. Bhuiyan ZA, van den Berg MP, van Tintelen JP, Bink-Boelk-
ens MT, Wiesfeld AC, Alders M, Postma AV, van Langen I,
Mannens MM, Wilde AA. Expanding spectrum of human RYR2-
related disease: new electrocardiographic, structural, and genetic
features. Circulation 116: 1569 –1576, 2007.
48. Blaauw Y, Schotten U, van Hunnik A, Neuberger HR, Allessie
MA. Cardioversion of persistent atrial fibrillation by a combination
of atrial specific and non-specific class III drugs in the goat. Car-
diovasc Res 75: 89 –98, 2007.
49. Black IW, Fatkin D, Sagar KB, Khandheria BK, Leung DY,
Galloway JM, Feneley MP, Walsh WF, Grimm RA, Stollberger
C. Exclusion of atrial thrombus by transesophageal echocardiog-
raphy does not preclude embolism after cardioversion of atrial
fibrillation. A multicenter study. Circulation 89: 2509 –2513, 1994.
50. Blatter LA, Kockskamper J, Sheehan KA, Zima AV, Huser J,
Lipsius SL. Local calcium gradients during excitation-contraction
coupling and alternans in atrial myocytes. J Physiol 546: 19 –31,
2003.
51. Blom NA, Gittenberger-de Groot AC, DeRuiter MC, Poel-
mann RE, Mentink MM, Ottenkamp J. Development of the
cardiac conduction tissue in human embryos using HNK-1 antigen
expression: possible relevance for understanding of abnormal
atrial automaticity. Circulation 99: 800 – 806, 1999.
52. Bode F, Sachs F, Franz M. Tarantula peptide inhibits atrial fibril-
lation. Nature 409: 6818 – 6819, 2001.
53. Boldt A, Wetzel U, Lauschke J, Weigl J, Gummert J, Hindricks
G, Kottkamp H, Dhein S. Fibrosis in left atrial tissue of patients
with atrial fibrillation with and without underlying mitral valve
disease. Heart 90: 400 – 405, 2004.
54. Bollmann A, Binias KH, Sonne K, Grothues F, Esperer HD,
Nikutta P, Klein HU. Electrocardiographic characteristics in pa-
tients with nonrheumatic atrial fibrillation and their relation to
echocardiographic parameters. Pacing Clin Electrophysiol 24:
1507–1513, 2001.
55. Bootman MD, Higazi DR, Coombes S, Roderick HL. Calcium
signalling during excitation-contraction coupling in mammalian
atrial myocytes. J Cell Sci 119: 3915–3925, 2006.
56. Bosch RF, Scherer CR, Rub N, Wohrl S, Steinmeyer K, Haase
H, Busch AE, Seipel L, Kuhlkamp V. Molecular mechanisms of
early electrical remodeling: transcriptional downregulation of ion
channel subunits reduces I(Ca,L) and I(to) in rapid atrial pacing in
rabbits. J Am Coll Cardiol 41: 858 – 869, 2003.
57. Bosch RF, Zeng X, Grammer JB, Popovic K, Mewis C, Ku-
hlkamp V. Ionic mechanisms of electrical remodeling in human
atrial fibrillation. Cardiovasc Res 44: 121–131, 1999.
58. Boutjdir M, Le Heuzey JY, Lavergne T, Chauvaud S, Guize L,
Carpentier A, Peronneau P. Inhomogeneity of cellular refracto-
riness in human atrium: factor of arrhythmia? Pacing Clin Electro-
physiol 9: 1095–1100, 1986.
59. Boyden PA, Hoffman BF. The effects on atrial electrophysiology
and structure of surgically induced right atrial enlargement in dogs.
Circ Res 49: 1319 –1331, 1981.
60. Boyden PA, Tilley LP, Albala A, Liu SK, Fenoglio JJ Jr, Wit
AL. Mechanisms for atrial arrhythmias associated with cardiomy-
opathy: a study of feline hearts with primary myocardial disease.
Circulation 69: 1036 –1047, 1984.
61. Brand TSM. Transforming growth factor-
signal transduction.
Circ Res 78: 173–179, 1996.
62. Brandt MC, Priebe L, Bohle T, Sudkamp M, Beuckelmann DJ.
The ultrarapid and the transient outward K
current in human
atrial fibrillation. Their possible role in postoperative atrial fibril-
lation. J Mol Cell Cardiol 32: 1885–1896, 2000.
63. Braun AP, Fedida D, Giles WR. Activation of alpha 1-adrenocep-
tors modulates the inwardly rectifying potassium currents of mam-
malian atrial myocytes. Pflügers Arch 421: 431– 439, 1992.
64. Brouwer IAHJ, Geleijnse JM, Zock PL, Witteman JC. Intake of
very long-chain n-3 fatty acids from fish and incidence of atrial
fibrillation. The Rotterdam Study. Am Heart J 151: 857– 862, 2006.
65. Bruins P, te Velthuis H, Yazdanbakhsh AP, Jansen PG, van
Hardevelt FW, de Beaumont EM, Wildevuur CR, Eijsman L,
Trouwborst A, Hack CE. Activation of the complement system
during and after cardiopulmonary bypass surgery: postsurgery ac-
tivation involves C-reactive protein and is associated with postop-
erative arrhythmia. Circulation 96: 3542–3548, 1997.
66. Brundel BJ, Ausma J, van Gelder IC, Van der Want JJ, van
Gilst WH, Crijns HJ, Henning RH. Activation of proteolysis by
calpains and structural changes in human paroxysmal and persis-
tent atrial fibrillation. Cardiovasc Res 54: 380 –389, 2002.
67. Brundel BJ, Henning RH. Calpain inhibition prevents pacing-
induced cellular remodeling in a HL-1 myocyte model for atrial
fibrillation. Cardiovasc Res 62: 521–528, 2004.
68. Brundel BJ, Van Gelder IC, Henning RH, Tieleman R, Tuinen-
burg AE, Wietses M, Grandjean J, Van Gilst WH, Crijns H. Ion
channel remodeling is related to intraoperative atrial effective re-
fractory periods in patients with paroxysmal and persistent atrial
fibrillation. Circulation 103: 684 – 690, 2001.
69. Brundel BJ, Van Gelder IC, Henning RH, Tuinenburg AE,
Wietses M, Grandjean JG, Wilde AA, Van Gilst WH, Crijns HJ.
Alterations in potassium channel gene expression in atria of pa-
tients with persistent and paroxysmal atrial fibrillation: differential
regulation of protein and mRNA levels for K
channels. J Am Coll
Cardiol 37: 926 –932, 2001.
70. Brunton TL, Fayer J. Note on independent pulsation of the
pulmonary veins and vena cava. Proc R Soc Lond B Biol Sci 25:
174 –176, 1876.
71. Buchanan JW. Spontaneous arrhythmias and conduction distur-
bances in domestic animals. Ann NY Acad Sci 127: 224 –238, 1965.
72. Bueno OFBE, Rothenberg ME, Molkentin JD. Defective T cell
development and function in calcineurinA beta-deficient mice. Proc
Natl Acad Sci USA 99: 9398 –9403, 2002.
73. Bui HM, Khrestian CM, Ryu K, Sahadevan J, Waldo AL. Fixed
intercaval block in the setting of atrial fibrillation promotes the
development of atrial flutter. Heart Rhythm 5: 1745–1752, 2008.
74. Bukowska A, Hirte D, Wolke C, Striggow F, Röhnert P, Huth
C, Klein HU, Goette A. Activation of the calcineurin signaling
pathway induces atrial hypertrophy during atrial fibrillation. Cell
Mol Life Sci 63: 333–342, 2006.
75. Bukowska A, Keilhoff G, Hirte D, Neumann M, Gardemann A,
Neumann KH, Röhl FW, Huth C, Goette A, Lendeckel U.
Mitochondrial dysfunction and redox signaling in atrial tachyar-
rhythmia. Exp Biol Med 233: 558 –574, 2008.
76. Burashnikov A, Antzelevitch C. Late-phase 3 EAD: a unique
mechanism contributing to initiation of atrial fibrillation. Pacing
Clin Electrophysiol 29: 290 –295, 2006.
77. Burashnikov A, Antzelevitch C. Reinduction of atrial fibrillation
immediately after termination of the arrhythmia is mediated by late
phase 3 early afterdepolarization-induced triggered activity. Circu-
lation 107: 2355–2360, 2003.
78. Burashnikov A, Mannava S, Antzelevitch C. Transmembrane
action potential heterogeneity in the canine isolated arterially per-
fused right atrium: effect of I
Kr
and I
Kur
/I
to
block. Am J Physiol
Heart Circ Physiol 286: H2393–H2400, 2004.
79. Burstein BLE, Calderone A, Nattel S. Differential behaviors of
atrial versus ventricular fibroblasts: a potential role for platelet-
derived growth factor in atrial-ventricular remodeling differences.
Circulation 117: 1630 –1641, 2008.
80. Burstein B, Nattel S. Atrial fibrosis: mechanisms and clinical
relevance in atrial fibrillation. J Am Coll Cardiol 51: 802– 809, 2008.
81. Burstein B, Qi XY, Yeh YH, Calderone A, Nattel S. Atrial
cardiomyocyte tachycardia alters cardiac fibroblast function: a
novel consideration in atrial remodeling. Cardiovasc Res 76: 442–
452, 2007.
82. Cai HLZ, Goette A, Mera F, Honeycutt C, Feterik K, Wilcox
JN, Dudley SC Jr, Harrison DG, Langberg JJ. Downregulation
of endocardial nitric oxide synthase expression and nitric oxide
production in atrial fibrillation: potential mechanisms for atrial
thrombosis and stroke. Circulation 106: 2854 –2858, 2002.
83. Calkins H, Brugada J, Packer DL, Cappato R, Chen SA, Crijns
HJ, Damiano RJ Jr, Davies DW, Haines DE, Haissaguerre M,
Iesaka Y, Jackman W, Jais P, Kottkamp H, Kuck KH, Lindsay
BD, Marchlinski FE, McCarthy PM, Mont JL, Morady F,
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 309
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
Nademanee K, Natale A, Pappone C, Prystowsky E, Raviele A,
Ruskin JN, Shemin RJ, Calkins H, Brugada J, Chen SA, Prys-
towsky EN, Kuck KH, Natale A, Haines DE, Marchlinski FE,
Calkins H, Davies DW, Lindsay BD, McCarthy PM, Packer DL,
Cappato R, Crijns HJ, Damiano RJ Jr, Haissaguerre M, Jack-
man WM, Jais P, Iesaka Y, Kottkamp H, Mont L, Morady F,
Nademanee K, Pappone C, Raviele A, Ruskin JN, Shemin RJ.
HRS/EHRA/ECAS Expert Consensus Statement on Catheter and
Surgical Ablation of Atrial Fibrillation: Recommendations for Per-
sonnel, Policy, Procedures and Follow-Up: A report of the Heart
Rhythm Society (HRS) Task Force on Catheter and Surgical Abla-
tion of Atrial Fibrillation Developed in partnership with the Euro-
pean Heart Rhythm Association (EHRA) and the European Cardiac
Arrhythmia Society (ECAS); in collaboration with the American
College of Cardiology (ACC), American Heart Association (AHA),
the Society of Thoracic Surgeons (STS) Endorsed and Approved by
the governing bodies of the American College of Cardiology, the
American Heart Association, the European Cardiac Arrhythmia
Society, the European Heart Rhythm Association, the Society of
Thoracic Surgeons, and the Heart Rhythm Society Europace. Am
Coll Cardiol 9: 335–379, 2007.
84. Calò LBL, Colivicchi F, Lamberti F, Loricchio ML, de Ruvo E,
Meo A, Pandozi C, Staibano M, Santini M. N-3 fatty acids for the
prevention of atrial fibrillation after coronary artery bypass sur-
gery: a randomized, controlled trial. J Am Coll Cardiol 45: 1723–
1728, 2005.
85. Camm AJ, Kirchhof P, Lip G, Savelieva I, Ernst S. Atrial
fibrillation. In: The ESC Textbook of Cardiology, edited by Camm
AJ, Serruys PW, Luscher T2009.
86. Camm J. Medical management of atrial fibrillation: state of the art.
J Cardiovasc Electrophysiol 17 Suppl 2: S2– 6, 2006.
87. Capucci A, Villani GQ, Aschieri D, Piepoli M. Effects of Class III
drugs on atrial fibrillation. J Cardiovasc Electrophysiol 9: S109 –
120, 1998.
88. Cardin S, Libby E, Pelletier P, Le Bouter S, Shiroshita-
Takeshita A, Le Meur N, Léger J, Demolombe S, Ponton A,
Glass L, Nattel S. Contrasting gene expression profiles in two
canine models of atrial fibrillation. Circ Res 100: 425– 433, 2007.
89. Cardin S, Pelletier P, Libby E, Le Bouter S, Xiao L, Kääb S,
Demolombe S, Glass L, Nattel S. Marked differences between
atrial and ventricular gene-expression remodeling in dogs with
experimental heart failure. J Mol Cell Cardiol 45: 821– 831, 2008.
90. Carl SL, Felix K, Caswell AH, Brandt NR, Ball WJ Jr, Vaghy
PL, Meissner G, Ferguson DG. Immunolocalization of sarcolem-
mal dihydropyridine receptor and sarcoplasmic reticular triadin
and ryanodine receptor in rabbit ventricle and atrium. J Cell Biol
129: 672– 682, 1995.
91. Carnes CA, Chung MK, Nakayama T, Nakayama H, Baliga RS,
Piao S, Kanderian A, Pavia S, Hamlin RL, McCarthy PM,
Bauer JA, Van Wagoner DR. Ascorbate attenuates atrial pacing-
induced peroxynitrite formation and electrical remodeling and de-
creases the incidence of postoperative atrial fibrillation. Circ Res
89: E32–E38, 2001.
92. Carnes CACM, Nakayama T, Nakayama N, Baliga RS, Piao S.
Ascorbate attenuates atrial pacing-induced peroxynitrite formation
and electrical remodeling and decreases the incidence of postop-
erative atrial fibrillation. Circ Res 89: e32– e38, 2001.
93. Centurion OA, Shimizu A, Isomoto S, Konoe A, Kaibara M,
Hayano M, Yano K. Influence of advancing age on fractionated
right atrial endocardial electrograms. Am J Cardiol 96: 239 –242,
2005.
94. Cha TJ, Ehrlich JR, Chartier D, Qi XY, Xiao L, Nattel S.
Kir3-based inward rectifier potassium current: potential role in
atrial tachycardia remodeling effects on atrial repolarization and
arrhythmias. Circulation 113: 1730 –1737, 2006.
95. Cha TJ, Ehrlich JR, Zhang L, Chartier D, Leung TK, Nattel S.
Atrial tachycardia remodeling of pulmonary vein cardiomyocytes:
comparison with left atrium and potential relation to arrhythmo-
genesis. Circulation 111: 728 –735, 2005.
96. Cha TJ, Ehrlich JR, Zhang L, Nattel S. Atrial ionic remodeling
induced by atrial tachycardia in the presence of congestive heart
failure. Circulation 110: 1520 –1526, 2004.
97. Cha TJ, Ehrlich JR, Zhang L, Shi Y, Tardif JC, Leung TK,
Nattel S. Dissociation between ionic remodeling and ability to
sustain atrial fibrillation during recovery from experimental con-
gestive heart failure. Circulation 109: 412– 418, 2004.
98. Cha YMDP, Shen WK, Jahangir A, Hart CY, Terzic A, Redfield
MM. Failing atrial myocardium: energetic deficits accompany
structural remodeling and electrical instability. Am J Physiol Heart
Circ Physiol 284: H1313–H1320, 2003.
99. Chello MPG, Candura D, Mastrobuoni S, Di Sciascio G, Agrò
F, Carassiti M, Covino E. Effects of atorvastatin on systemic
inflammatory response after coronary bypass surgery. Crit Care
Med 34: 660 – 667, 2006.
100. Chelu M, Sarma S, Sood S, Wang S, van Oort R, Skapura D, Li
N, Santonastasi M, Mueller F, Schmitz W, Schotten U, Ander-
son M, Valderrabano M, Dobrev D, Wehrens XH. Calmodulin
kinase-II-mediated sarcoplasmic reticulum Ca
2
leak promotes
atrial fibrillation in mice. J Clin Invest 119: 1940 –1951, 2009.
101. Chen Y, Chen SA, Chen Y, Yeh HI, Chan P, Chang MS, Lin CI.
Effects of rapid atrial pacing on the arrhythmogenic activity of
single cardiomyocytes from pulmonary veins: implication in initia-
tion of atrial fibrillation. Circulation 104: 2849 –2854, 2001.
102. Chen YH, Xu SJ, Bendahhou S, Wang XL, Wang Y, Xu WY, Jin
HW, Sun H, Su XY, Zhuang QN, Yang YQ, Li YB, Liu Y, Xu HJ,
Li XF, Ma N, Mou CP, Chen Z, Barhanin J, Huang W. KCNQ1
gain-of-function mutation in familial atrial fibrillation. Science 299:
251–254, 2003.
103. Chen YJ, Chen SA, Chang MS, Lin CI. Arrhythmogenic activity
of cardiac muscle in pulmonary veins of the dog: implication for the
genesis of atrial fibrillation. Cardiovasc Res 48: 265–273, 2000.
104. Chen YJ, Chen SA, Chen YC, Yeh HI, Chang MS, Lin CI.
Electrophysiology of single cardiomyocytes isolated from rabbit
pulmonary veins: implication in initiation of focal atrial fibrillation.
Basic Res Cardiol 97: 26 –34, 2002.
105. Cheng TH, Lee FY, Wei J, Lin CI. Comparison of calcium-current
in isolated atrial myocytes from failing and nonfailing human
hearts. Mol Cell Biochem 157: 157–162, 1996.
106. Cheung DW. Electrical activity of the pulmonary vein and its
interaction with the right atrium in the guinea-pig. J Physiol 314:
445– 456, 1981.
107. Chiu YT, Wu TJ, Wei HJ, Cheng CC, Lin NN, Chen YT, Ting
CT. Increased extracellular collagen matrix in myocardial sleeves
of pulmonary veins: an additional mechanism facilitating repetitive
rapid activities in chronic pacing-induced sustained atrial fibrilla-
tion. J Cardiovasc Electrophysiol 16: 753–759, 2005.
108. Choisy SC, Arberry LA, Hancox JC, James AF. Increased sus-
ceptibility to atrial tachyarrhythmia in spontaneously hypertensive
rat hearts. Hypertension 49: 498 –505, 2007.
109. Chou CC, Nguyen BL, Tan AY, Chang PC, Lee HL, Lin FC, Yeh
SJ, Fishbein MC, Lin SF, Wu D, Wen MS, Chen PS. Intracellular
calcium dynamics and acetylcholine-induced triggered activity in
the pulmonary veins of dogs with pacing-induced heart failure.
Heart Rhythm 5: 1170 –1177, 2008.
110. Chou CC, Nihei M, Zhou S, Tan A, Kawase A, Macias ES,
Fishbein MC, Lin SF, Chen PS. Intracellular calcium dynamics
and anisotropic reentry in isolated canine pulmonary veins and left
atrium. Circulation 111: 2889 –2897, 2005.
111. Christ T, Boknik P, Wöhrl S, Wettwer E, Graf EM, Bosch RF,
Knaut M, Schmitz W, Ravens U, Dobrev D. L-type Ca
2
current
downregulation in chronic human atrial fibrillation is associated
with increased activity of protein phosphatases. Circulation 110:
2651–2657, 2004.
112. Christ T, Wettwer E, Voigt N, Hála O, Radicke S, Matschke K,
Várro A, Dobrev D, Ravens U. Pathology-specific effects of the
I
Kur
/I
to
/I
K,ACh
blocker AVE0118 on ion channels in human chronic
atrial fibrillation. Br J Pharmacol 154: 1619 –1630, 2008.
113. Ciaccio EJ, Biviano AB, Whang W, Wit AL, Garan H, Coromi-
las J. New methods for estimating local electrical activation rate
during atrial fibrillation. Heart Rhythm 6: 21–32, 2009.
114. Clapham DE, Lechleiter JD, Girard S. Intracellular waves ob-
served by confocal microscopy from Xenopus oocytes. Adv Second
Messenger Phosphoprotein Res 28: 161–165, 1993.
115. Cleland JG, Swedberg K, Follath F, Komajda M, Cohen-Solal
A, Aguilar JC, Dietz R, Gavazzi A, Hobbs R, Korewicki J,
310 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
Madeira HC, Moiseyev VS, Preda I, van Gilst WH, Widimsky J,
Freemantle N, Eastaugh J, Mason J. The EuroHeart Failure
survey programme: a survey on the quality of care among patients
with heart failure in Europe. Part 1: patient characteristics and
diagnosis. Eur Heart J 24: 442– 463, 2003.
116. Clermont G, Vergely C, Jazayeri S, Lahet JJ, Goudeau JJ,
Lecour S, David M, Rochette L, Girard C. Systemic free radical
activation is a major event involved in myocardial oxidative stress
related to cardiopulmonary bypass. Anesthesiology 96: 80 – 87,
2002.
117. Coetzee WA, Opie LH. Effects of components of ischemia and
metabolic inhibition on delayed afterdepolarizations in guinea pig
papillary muscle. Circ Res 61: 157–165, 1987.
118. Cohen HC, Arbel ER. Tachycardias and electrical pacing. Med
Clin North Am 60: 343–367, 1976.
119. Comtois P, Kneller J, Nattel S. Of circles and spirals: bridging
the gap between the leading circle and spiral wave concepts of
cardiac reentry. Europace 7Suppl 2: 10 –20, 2005.
120. Connolly SJ, Ezekowitz MD, Yusuf S, Eikelboom J, Oldgren J,
Parekh A, Pogue J, Reilly PA, Themeles E, Varrone J, Wang S,
Alings M, Xavier D, Zhu J, Diaz R, Lewis BS, Darius H, Diener
HC, Joyner CD, Wallentin L. Dabigatran versus warfarin in
patients with atrial fibrillation. N Engl J Med 361: 1139 –1151, 2009.
121. Cooradi DCS, Benussi S, Nascimbene S, Pastori P, Calvi S,
Maestri R, Astorri E, Pappone C, Alfieri O. Regional left atrial
interstitial remodeling in patients with chronic atrial fibrillation
undergoing mitral-valve surgery. Virchows Arch 445: 498 –505,
2004.
122. Cordina J, Mead G. Pharmacological cardioversion for atrial
fibrillation and flutter. Indian Pacing Electrophysiol J 6: 31, 2006.
123. Corley SD, Epstein AE, DiMarco JP, Domanski MJ, Geller N,
Greene HL, Josephson RA, Kellen JC, Klein RC, Krahn AD,
Mickel M, Mitchell LB, Nelson JD, Rosenberg Y, Schron E,
Shemanski L, Waldo AL, Wyse DG. Relationships between sinus
rhythm, treatment, and survival in the Atrial Fibrillation Follow-Up
Investigation of Rhythm Management (AFFIRM) Study. Circula-
tion 109: 1509 –1513, 2004.
124. Cotoi S, Gavrilescu S, Pop T, Vicas E. The prognostic value of
right astrium monophasic action potential after conversion of atrial
fibrillation. Eur J Clin Invest 2: 472– 474, 1972.
125. Coumel P. Autonomic influences in atrial tachyarrhythmias. J
Cardiovasc Electrophysiol 7: 999 –1007, 1996.
126. Courtemanche M, Ramirez RJ, Nattel S. Ionic mechanisms
underlying human atrial action potential properties: insights from a
mathematical model. Am J Physiol Heart Circ Physiol 275: H301–
H321, 1998.
127. Coutu P, Chartier D, Nattel S. Comparison of Ca
2
-handling
properties of canine pulmonary vein and left atrial cardiomyocytes.
Am J Physiol Heart Circ Physiol 291: H2290 –H2300, 2006.
128. Cox JL, Boineau JP, Schuessler RB, Kater KM, Lappas DG.
Five-year experience with the maze procedure for atrial fibrillation.
Ann Thorac Surg 56: 814 – 823, 1993.
129. Cox JL, Canavan TE, Schuessler RB, Cain ME, Lindsay BD,
Stone C, Smith PK, Corr PB, Boineau JP. The surgical treat-
ment of atrial fibrillation. II. Intraoperative electrophysiologic map-
ping and description of the electrophysiologic basis of atrial flutter
and atrial fibrillation. J Thorac Cardiovasc Surg 101: 406 – 426,
1991.
130. Crabtree GROE. NFAT signalling: choreographing the social lives
of cells. Cell Mol Life Sci 109 Suppl: S67–79, 2002.
130a.Cuculich PS, Wang Y, Lindsay BD, Faddis MN, Schuessler RB,
Damiano RJ Jr, Li L, Rudy Y. Noninvasive characterization of
epicardial activation in humans with diverse atrial fibrillation pat-
terns. Circulation 122: 1364 –1372, 2010.
131. Cummings JE, Gill I, Akhrass R, Dery M, Biblo LA, Quan KJ.
Preservation of the anterior fat pad paradoxically decreases the
incidence of postoperative atrial fibrillation in humans. J Am Coll
Cardiol 43: 994 –1000, 2004.
132. Daoud EG, Bogun F, Goyal R, Harvey M, Man KC, Strickber-
ger SA, Morady F. Effect of atrial fibrillation on atrial refractori-
ness in humans. Circulation 94: 1600 –1606, 1996.
133. Daoud EG, Marcovitz P, Knight BP, Goyal R, Man KC, Strick-
berger SA, Armstrong WF, Morady F. Short-term effect of atrial
fibrillation on atrial contractile function in humans. Circulation 99:
3024 –3027, 1999.
134. Darbar D, Kannankeril PJ, Donahue BS, Kucera G, Stubble-
field T, Haines JL, George AL Jr, Roden DM. Cardiac sodium
channel (SCN5A) variants associated with atrial fibrillation. Circu-
lation 117: 1927–1935, 2008.
135. De Bakker JM, van Capelle FJ, Janse MJ, Tasseron S, Ver-
meulen JT, de Jonge N, Lahpor JR. Fractionated electrograms
in dilated cardiomyopathy: origin and relation to abnormal conduc-
tion. J Am Coll Cardiol 27: 1071–1078, 1996.
135a.De Groot NM, Houben RP, Smeets JL, Boersma E, Schotten U,
Schalij MJ, Crijns H, Allessie MA. Electropathogical substrate
of longstanding persistent atrial fibrillation in patients with struc-
tural heart disease: epicardial breakthrough. Circulation 122:
1674 –1682, 2010.
136. De Haan S, Greiser M, Harks E, Blaauw Y, van Hunnik A,
Verheule S, Allessie M, Schotten U. AVE0118, blocker of the
transient outward current [I(to)] and ultrarapid delayed rectifier
current [I(Kur)], fully restores atrial contractility after cardiover-
sion of atrial fibrillation in the goat. Circulation 114: 1234–1242,
2006.
137. De Ponti R, Tritto M, Lanzotti ME, Spadacini G, Marazzi R,
Moretti P, Salerno-Uriarte J. A Computerized high-density map-
ping of the pulmonary veins: new insights into their electrical
activation in patients with atrial fibrillation. Europace 6: 97–108,
2004.
138. De Vos CB, Nieuwlaat R, Crijns HJ, Camm AJ, LeHeuzey JY,
Kirchhof CJ, Capucci A, Breithardt G, Vardas PE, Pisters R,
Tieleman RG. Autonomic trigger patterns and anti-arrhythmic
treatment of paroxysmal atrial fibrillation: data from the Euro
Heart Survey. Eur Heart J 29: 632– 639, 2008.
139. De Vos CB, Weijs B, Crijns HJ, Cheriex EC, Palmans A,
Habets J, Prins MH, Pisters R, Nieuwlaat R, Tieleman RG.
Atrial tissue doppler imaging for prediction of new-onset atrial
fibrillation. Heart 95: 835– 840, 2009.
140. Dernellis JPM. Relationship between C-reactive protein concen-
trations during glucocorticoid therapy and recurrent atrial fibrilla-
tion. Eur Heart J 25: 1100 –1107, 2004.
141. Dhamoon AS, Jalife J. The inward rectifier current (I
K1
) controls
cardiac excitability and is involved in arrhythmogenesis. Heart
Rhythm 2: 316 –324, 2005.
142. Dibb KM, Clarke JD, Horn MA, Richards MA, Graham HK,
Eisner DA, Trafford AW. Characterization of an extensive trans-
verse tubular network in sheep atrial myocytes and its depletion in
heart failure. Circ Heart Fail 2: 482– 489, 2009.
143. Dibs SR, Ng J, Arora R, Passman RS, Kadish AH, Goldberger
JJ. Spatiotemporal characterization of atrial activation in persis-
tent human atrial fibrillation: multisite electrogram analysis and
surface electrocardiographic correlations–a pilot study. Heart
Rhythm 5: 686 – 693, 2008.
144. Diness JG, Sorensen US, Nissen JD, Al-Shahib B, Jespersen
T, Grunnet M, Hansen RS. Inhibition of small conductance Ca
2
-
activated K
channels terminates and protects against atrial fibril-
lation. Circ Arrhythm Electrophysiol 3: 380 –390, 2010.
145. Dobrev D, Friedrich A, Voigt N, Jost N, Wettwer E, Christ T,
Knaut M, Ravens U. The G protein-gated potassium current
I(K,ACh) is constitutively active in patients with chronic atrial
fibrillation. Circulation 112: 3697–3706, 2005.
146. Dobrev D, Graf E, Wettwer E, Himmel HM, Hála O, Doerfel C,
Christ T, Schüler S, Ravens U. Molecular basis of downregula-
tion of G-protein-coupled inward rectifying K
current [I(K,ACh)]
in chronic human atrial fibrillation: decrease in GIRK4 mRNA
correlates with reduced I(K,ACh) and muscarinic receptor-medi-
ated shortening of action potentials. Circulation 104: 2551–2557,
2001.
147. Dobrev D, Nattel S. New antiarrhythmic drugs for treatment of
atrial fibrillation. Lancet 375: 1212–1223, 2010.
148. Dobrev D, Wettwer E, Kortner A, Knaut M, Schuler S, Ravens
U. Human inward rectifier potassium channels in chronic and
postoperative atrial fibrillation. Cardiovasc Res 54: 397– 404, 2002.
149. Dobrzynski H, Marples DD, Musa H, Yamanushi TT, Hender-
son Z, Takagishi Y, Honjo H, Kodama I, Boyett MR. Distribu-
tion of the muscarinic K
channel proteins Kir3.1 and Kir3 4 in the
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 311
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
ventricle, atrium, and sinoatrial node of heart. J Histochem Cyto-
chem 49: 1221–1234, 2001.
150. Ducharme ASK, Pfeffer MA, Cohen-Solal A, Granger CB,
Maggioni AP, Michelson EL, McMurray JJ, Olsson L, Rouleau
JL, Young JB, Olofsson B, Puu M, Yusuf S, Investigators
CHARM. Prevention of atrial fibrillation in patients with symptom-
atic chronic heart failure by candesartan in the Candesartan in
Heart failure: Assessment of Reduction in Mortality and morbidity
(CHARM) program. Am Heart J 152: 86 –92, 2006.
151. Dudley SC Jr, McCann LA, Honeycutt C, Diamandopoulos L,
Fukai T. Atrial fibrillation increases production of superoxide by
the left atrium and left atrial appendage: role of the NADPH and
xanthine oxidases. Circulation 112: 1266 –1273, 2005.
152. Duffy HS, Wit AL. Is there a role for remodeled connexins in AF?
No simple answers. J Mol Cell Cardiol 44: 4 –13, 2008.
153. Dun W, Yagi T, Rosen MR, Boyden PA. Calcium and potassium
currents in cells from adult and aged canine right atria. Cardiovasc
Res 58: 526 –534, 2003.
154. Dupont E, Ko Y, Rothery S, Coppen SR, Baghai M, Haw M,
Severs NJ. The gap-junctional protein connexin40 is elevated in
patients susceptible to postoperative atrial fibrillation. Circulation
103: 842– 849, 2001.
155. Duytschaever M, Danse P, Allessie M. Supervulnerable phase
immediately after termination of atrial fibrillation. J Cardiovasc
Electrophysiol 13: 267–275, 2002.
156. Duytschaever M, Heyse A, de Sutter J, Crijns H, Gillebert T,
Tavernier R, Tieleman R. Transthoracic tissue Doppler imaging
of the atria: a novel method to determine the atrial fibrillation cycle
length. J Cardiovasc Electrophysiol 17: 1202–1209, 2006.
157. Echahidi N, Pibarot P, O’Hara G, Mathieu P. Mechanisms,
prevention, and treatment of atrial fibrillation after cardiac surgery.
J Am Coll Cardiol 51: 793– 801, 2008.
158. Eckardt L, Kirchhof P, Loh P, Schulze-Bahr E, Johna R, Wich-
ter T, Breithardt G, Haverkamp W, Borggrefe M. Brugada
syndrome and supraventricular tachyarrhythmias: a novel associa-
tion? J Cardiovasc Electrophysiol 12: 680 – 685, 2001.
159. Eckstein J, Verheule S, de Groot N, Allessie M, Schotten U.
Mechanisms of perpetuation of atrial fibrillation in chronically
dilated atria. Prog Biophys Mol Biol 97: 435– 451, 2008.
159a.Eckstein J, Maesen B, Linz D, Zeemering S, van Hunnik A,
Verheule S, Allessie MA, Schotten U. Time course and mecha-
nisms of endo-epicardial electrical dissociation during atrial fibril-
lation in the goat. Cardiovasc Res 2010 Nov 23. [Epub ahead of
print]
160. Ehrlich JR, Biliczki P, Hohnloser SH, Nattel S. Atrial-selective
approaches for the treatment of atrial fibrillation. J Am Coll Car-
diol 51: 787–792, 2008.
161. 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.
162. Ehrlich JR, Cha TJ, Zhang L, Chartier D, Villeneuve L, Hébert
TE, Nattel S. Characterization of a hyperpolarization-activated
time-dependent potassium current in canine cardiomyocytes from
pulmonary vein myocardial sleeves and left atrium. J Physiol 557:
583–597, 2004.
163. Eiras S, Narolska NA, van Loon RB, Boontje NM, Zaremba R,
Jimenez CR, Visser FC, Stooker W, van der Velden J, Stienen
GJ. Alterations in contractile protein composition and function in
human atrial dilatation and atrial fibrillation. J Mol Cell Cardiol 41:
467– 477, 2006.
164. El-Armouche A, Boknik P, Eschenhagen T, Carrier L, Knaut
M, Ravens U, Dobrev D. Molecular determinants of altered Ca
2
handling in human chronic atrial fibrillation. Circulation 114: 670 –
680, 2006.
165. Elenes S, Rubart M, Moreno AP. Junctional communication
between isolated pairs of canine atrial cells is mediated by homo-
geneous and heterogeneous gap junction channels. J Cardiovasc
Electrophysiol 10: 990 –1004, 1999.
166. Ellinor PT, Lunetta KL, Glazer NL, Pfeufer A, Alonso A,
Chung MK, Sinner MF, de Bakker PI, Mueller M, Lubitz SA,
Fox E, Darbar D, Smith NL, Smith JD, Schnabel RB, Soliman
EZ, Rice KM, Van Wagoner DR, Beckmann BM, van Noord C,
Wang K, Ehret GB, Rotter JI, Hazen SL, Steinbeck G, Smith
AV, Launer LJ, Harris TB, Makino S, Nelis M, Milan DJ, Perz
S, Esko T, Kottgen A, Moebus S, Newton-Cheh C, Li M,
Mohlenkamp S, Wang TJ, Kao WH, Vasan RS, Nothen MM,
MacRae CA, Stricker BH, Hofman A, Uitterlinden AG, Levy D,
Boerwinkle E, Metspalu A, Topol EJ, Chakravarti A, Gudna-
son V, Psaty BM, Roden DM, Meitinger T, Wichmann HE,
Witteman JC, Barnard J, Arking DE, Benjamin EJ, Heckbert
SR, Kaab S. Common variants in KCNN3 are associated with lone
atrial fibrillation. Nat Genet 42: 240 –244.
167. Ellinor PT, Nam EG, Shea MA, Milan DJ, Ruskin JN, MacRae
CA. Cardiac sodium channel mutation in atrial fibrillation. Heart
Rhythm 5: 99 –105, 2008.
168. Elvan A, Linnenbank AC, van Bemmel MW, Misier AR, Delnoy
PP, Beukema WP, de Bakker JM. Dominant frequency of atrial
fibrillation correlates poorly with atrial fibrillation cycle length.
Circ Arrhythm Electrophysiol 2: 634 – 644, 2009.
168a.European Heart Rhythm Association, European Association
for Cardio-Thoracic Surgery, Camm AJ, Kirchhof P, Lip GY,
Schotten U, Savelieva I, Ernst S, Van Gelder IC, Al-Attr N,
Hindricks G, Prendergast B, Heidbuchel H, Alfieri O, Angelini
A, Atar D, Colonna P, De Caternia R, De Sutter J, Goette A,
Gorenek B, Heldal M, Hohloser SH, Kolh P, Le Heuzey JY,
Ponikowski P, Rutten FH. Guidelines for the management of
atrial fibrillation: the Task Force for the Management of Atrial
Fibrillation of the European Society of Cardiology (ESC). Eur
Heart J 31: 2369 –2429, 2010.
169. Everett TH, Olgin JE. Atrial fibrosis and the mechanisms of atrial
fibrillation. Heart Rhythm: Official J Heart Rhythm Soc 4: S24 –27,
2007.
170. Everett THt Li H, Mangrum JM, McRury ID, Mitchell MA,
Redick JA, Haines DE. Electrical, morphological, and ultrastruc-
tural remodeling and reverse remodeling in a canine model of
chronic atrial fibrillation. Circulation 102: 1454 –1460, 2000.
171. Falk RH, Decara J, Abascal V. Is pharmacologic cardioversion of
atrial fibrillation really preferable to electrical cardioversion? JAm
Coll Cardiol 31: 1446 –1447, 1998.
172. Fareh S, Villemaire C, Nattel S. Importance of refractoriness
heterogeneity in the enhanced vulnerability to atrial fibrillation
induction caused by tachycardia-induced atrial electrical remodel-
ing. Circulation 98: 2202–2209, 1998.
173. Fatkin D, Kuchar DL, Thorburn CW, Feneley MP. Transesoph-
ageal echocardiography before and during direct current cardio-
version of atrial fibrillation: evidence for “atrial stunning” as a
mechanism of thromboembolic complications. J Am Coll Cardiol
23: 307–316, 1994.
174. Fedida D, Braun AP, Giles WR. Alpha 1-adrenoceptors reduce
background K
current in rabbit ventricular myocytes. J Physiol
441: 673– 684, 1991.
175. Fedida D, Shimoni Y, Giles WR. Alpha-adrenergic modulation of
the transient outward current in rabbit atrial myocytes. J Physiol
423: 257–277, 1990.
176. Feinberg WM, Blackshear JL, Laupacis A, Kronmal R, Hart
RG. Prevalence, age distribution, and gender of patients with atrial
fibrillation. Analysis and implications. Arch Intern Med 155: 469–
473, 1995.
177. Fenelon G, Shepard RK, Stambler BS. Focal origin of atrial
tachycardia in dogs with rapid ventricular pacing-induced heart
failure. J Cardiovasc Electrophysiol 14: 1093–1102, 2003.
178. Feng J, Yue L, Wang Z, Nattel S. Ionic mechanisms of regional
action potential heterogeneity in the canine right atrium. Circ Res
83: 541–551, 1998.
179. Ferrier GR, Moffat MP, Lukas A. Possible mechanisms of ven-
tricular arrhythmias elicited by ischemia followed by reperfusion.
Studies on isolated canine ventricular tissues. Circ Res 56: 184 –
194, 1985.
180. Fiala M, Chovancik J, Nevralova R, Neuwirth R, Jiravsky O,
Januska J, Branny M. Termination of long-lasting persistent ver-
sus short-lasting persistent and paroxysmal atrial fibrillation by
ablation. Pacing Clin Electrophysiol 31: 985–997, 2008.
181. Fioranelli M, Piccoli M, Mileto GM, Sgreccia F, Azzolini P,
Risa MP, Francardelli RL, Venturini E, Puglisi A. Analysis of
312 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
heart rate variability five minutes before the onset of paroxysmal
atrial fibrillation. Pacing Clin Electrophysiol 22: 743–749, 1999.
182. Franz MR, Karasik PL, Li C, Moubarak J, Chavez M. Electrical
remodeling of the human atrium: similar effects in patients with
chronic atrial fibrillation and atrial flutter. J Am Coll Cardiol 30:
1785–1792, 1997.
183. Fraser HR, Turner R. Electrocardiography in mitral valvular dis-
ease. Br Heart J 17: 459 – 483, 1955.
184. Frost LVP. n-3 Fatty acids consumed from fish and risk of atrial
fibrillation or flutter: the Danish Diet, Cancer, and Health Study.
Am J Clin Nutr 81: 50 –54, 2005.
185. Frustaci A, Caldarulo M, Buffon A, Bellocci F, Fenici R,
Melina D. Cardiac biopsy in patients with “primary” atrial fibrilla-
tion. Histologic evidence of occult myocardial diseases. Chest 100:
303–306, 1991.
186. Frustaci ACC, Bellocci F, Morgante E, Russo MA, Maseri A.
Histological substrate of atrial biopsies in patients with lone atrial
fibrillation. Circulation 96: 1180 –1184, 1997.
187. Frustaci A, Chimenti C, Bellocci F, Morgante E, Russo MA,
Maseri A. Histological substrate of atrial biopsies in patients with
lone atrial fibrillation. Circulation 96: 1180 –1184, 1997.
188. Fuster V, Ryden LE, Cannom DS, Crijns HJ, Curtis AB, El-
lenbogen KA, Halperin JL, Le Heuzey JY, Kay GN, Lowe JE,
Olsson SB, Prystowsky EN, Tamargo JL, Wann S, Priori SG,
Blanc JJ, Budaj A, Camm AJ, Dean V, Deckers JW, Despres
C, Dickstein K, Lekakis J, McGregor K, Metra M, Morais J,
Osterspey A, Tamargo JL, Zamorano JL, Smith SC Jr, Jacobs
AK, Adams CD, Anderson JL, Antman EM, Halperin JL, Hunt
SA, Nishimura R, Ornato JP, Page RL, Riegel B. ACC/AHA/ESC
2006 guidelines for the management of patients with atrial fibrilla-
tion-executive summary: a report of the American College of Car-
diology/American Heart Association Task Force on practice guide-
lines and the European Society of Cardiology Committee for Prac-
tice Guidelines (Writing Committee to Revise the 2001 Guidelines
for the Management of Patients with Atrial Fibrillation) Developed
in collaboration with the European Heart Rhythm Association and
the Heart Rhythm Society. Eur Heart J 27: 1979 –2030, 2006.
189. Gaborit N, Steenman M, Lamirault G, Le Meur N, Le Bouter S,
Lande G, Léger J, Charpentier F, Christ T, Dobrev D, Escande
D, Nattel S, Demolombe S. Human atrial ion channel and trans-
porter subunit gene-expression remodeling associated with valvu-
lar heart disease and atrial fibrillation. Circulation 112: 471–481,
2005.
190. Gage BF, Waterman AD, Shannon W, Boechler M, Rich MW,
Radford MJ. Validation of clinical classification schemes for pre-
dicting stroke: results from the National Registry of Atrial Fibrilla-
tion. JAMA 285: 2864 –2870, 2001.
191. Gaita F, Giustetto C, Bianchi F, Wolpert C, Schimpf R, Ric-
cardi R, Grossi S, Richiardi E, Borggrefe M. Short QT Syn-
drome: a familial cause of sudden death. Circulation 108: 965–970,
2003.
192. Garrey W. Auricular fibrillation. Physiol Rev 4: 215–250, 1924.
193. Gaspo R, Bosch RF, Bou-Abboud E, Nattel S. Tachycardia-
induced changes in Na
current in a chronic dog model of atrial
fibrillation. Circ Res 81: 1045–1052, 1997.
194. Gaspo R, Bosch RF, Talajic M, Nattel S. Functional mechanisms
underlying tachycardia-induced sustained atrial fibrillation in a
chronic dog model. Circulation 96: 4027– 4035, 1997.
195. Giles WR, Imaizumi Y. Comparison of potassium currents in
rabbit atrial and ventricular cells. J Physiol 405: 123–145, 1988.
196. Giustetto C, Di Monte F, Wolpert C, Borggrefe M, Schimpf R,
Sbragia P, Leone G, Maury P, Anttonen O, Haissaguerre M,
Gaita F. Short QT syndrome: clinical findings and diagnostic-
therapeutic implications. Eur Heart J 27: 2440 –2447, 2006.
197. Goette AAM, Röcken C, Staack T, Bechtloff R, Reinhold R,
Huth C, Ansorge S, Klein HU, Lendeckel U. Calpains and cyto-
kines in fibrillating human atria. Am J Physiol Heart Circ Physiol
283: H264 –H272, 2002.
198. Goette A, Arndt M, Rocken C, Staack T, Bechtloff R, Reinhold
D, Huth C, Ansorge S, Klein HU, Lendeckel U. Calpains and
cytokines in fibrillating human atria. Am J Physiol Heart Circ
Physiol 283: H264 –H272, 2002.
199. Goette A, Juenemann G, Peters B, Klein H, Roessner A, Huth
C, Rocken C. Determinants and consequences of atrial fibrosis in
patients undergoing open heart surgery. Cardiovasc Res 54: 390–
396, 2002.
200. Goette ARC, Nepple K, Lendeckel U. Proteases and arrhyth-
mias. In: Proteases in Tissue Remodelling of Lung and Heart,
edited by Lendeckel U. New York: Plenum, 2003, p. 191–390 –218.
201. Goette AST, Arndt M, Röcken C, Geller C, Huth C, Ansorge S,
Klein HU, Lendeckel U. Increased expression of extracellular-
signal regulated kinase and angiotensin-converting enzyme in hu-
man atria during atrial fibrillation. J Am Coll Cardiol 35: 1669 –
1677, 2000.
202. Goldstein RN, Ryu K, Khrestian C, Van Wagoner DR, Waldo
AL. Prednisone prevents inducible atrial flutter in the canine sterile
pericarditis model. J Cardiovasc Electrophysiol 19: 74 – 81, 2008.
203. Gollob MH, Jones DL, Krahn AD, Danis L, Gong XQ, Shao Q,
Liu X, Veinot JP, Tang AS, Stewart AF, Tesson F, Klein GJ,
Yee R, Skanes AC, Guiraudon GM, Ebihara L, Bai D. Somatic
mutations in the connexin 40 gene (GJA5) in atrial fibrillation. N
Engl J Med 354: 2677–2688, 2006.
204. Gollob MH, Seger JJ, Gollob TN, Tapscott T, Gonzales O,
Bachinski L, Roberts R. Novel PRKAG2 mutation responsible for
the genetic syndrome of ventricular preexcitation and conduction
system disease with childhood onset and absence of cardiac hy-
pertrophy. Circulation 104: 3030 –3033, 2001.
205. Golod DA, Kumar R, Joyner RW. Determinants of action poten-
tial initiation in isolated rabbit atrial and ventricular myocytes. Am
J Physiol Heart Circ Physiol 274: H1902–H1913, 1998.
206. Gosselink AT, Crijns HJ, Hamer HP, Hillege H, Lie KI.
Changes in left and right atrial size after cardioversion of atrial
fibrillation: role of mitral valve disease. J Am Coll Cardiol 22:
1666 –1672, 1993.
207. Gourdie RG, Green CR, Severs NJ, Thompson RP. Immunola-
belling patterns of gap junction connexins in the developing and
mature rat heart. Anat Embryol 185: 363–378, 1992.
208. Gramley F, Lorenzen J, Plisiene J, Rakauskas M, Benetis R,
Schmid M, Autschbach R, Knackstedt C, Schimpf T, Mischke
K, Gressner A, Hanrath P, Kelm M, Schauerte P. Decreased
plasminogen activator inhibitor and tissue metalloproteinase inhib-
itor expression may promote increased metalloproteinase activity
with increasing duration of human atrial fibrillation. J Cardiovasc
Electrophysiol 18: 1076 –1082, 2007.
209. Grammer JB, Bosch RF, Kuhlkamp V, Seipel L. Molecular
remodeling of Kv4.3 potassium channels in human atrial fibrilla-
tion. J Cardiovasc Electrophysiol 11: 626 – 633, 2000.
210. Greiser M, Harks E, Verheule S, Allessie M, Schotten U.
Functional but not structural changes underlie failure of intracel-
lular Ca wave propagation in tachycardia-induced atrial remodel-
ing. Circulation 118: S438, 2008.
211. Greiser M, Neuberger HR, Harks E, El-Armouche A, Boknik
P, de Haan S, Verheyen F, Verheule S, Schmitz W, Ravens U,
Nattel S, Allessie MA, Dobrev D, Schotten U. Distinct contrac-
tile and molecular differences between two goat models of atrial
dysfunction: AV block-induced atrial dilatation and atrial fibrilla-
tion. J Mol Cell Cardiol 46: 385–394, 2009.
212. Grimm RA, Leung DY, Black IW, Stewart WJ, Thomas JD,
Klein AL. Left atrial appendage “stunning” after spontaneous con-
version of atrial fibrillation demonstrated by transesophageal
Doppler echocardiography. Am Heart J 130: 174 –176, 1995.
213. Gudbjartsson DF, Arnar DO, Helgadottir A, Gretarsdottir S,
Holm H, Sigurdsson A, Jonasdottir A, Baker A, Thorleifsson
G, Kristjansson K, Palsson A, Blondal T, Sulem P, Backman
VM, Hardarson GA, Palsdottir E, Helgason A, Sigurjonsdottir
R, Sverrisson JT, Kostulas K, Ng MC, Baum L, So WY, Wong
KS, Chan JC, Furie KL, Greenberg SM, Sale M, Kelly P,
MacRae CA, Smith EE, Rosand J, Hillert J, Ma RC, Ellinor PT,
Thorgeirsson G, Gulcher JR, Kong A, Thorsteinsdottir U,
Stefansson K. Variants conferring risk of atrial fibrillation on
chromosome 4q25. Nature 448: 353–357, 2007.
214. Guerra JM, Everett THt Lee KW, Wilson E, Olgin J. Effects of
the gap junction modifier rotigaptide (ZP123) on atrial conduction
and vulnerability to atrial fibrillation. Circulation 114: 110 –118,
2006.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 313
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
215. Guerra PG, Thibault B, Dubuc M, Talajic M, Roy D, Crepeau
J, Nattel S, Tardif JC. Identification of atrial tissue in pulmonary
veins using intravascular ultrasound. J Am Soc Echocardiogr 16:
982–987, 2003.
216. Hagendorff A, Schumacher B, Kirchhoff S, Luderitz B, Wil-
lecke K. Conduction disturbances and increased atrial vulnerabil-
ity in Connexin40-deficient mice analyzed by transesophageal stim-
ulation. Circulation 99: 1508 –1515, 1999.
217. Haissaguerre M, Jais P, Shah DC, Takahashi A, Hocini M,
Quiniou G, Garrigue S, LeMouroux A, LeMetayer P, Clementy
J. Spontaneous initiation of atrial fibrillation by ectopic beats
originating in the pulmonary veins. N Engl J Med 339: 659 – 666,
1998.
218. Halonen J, Halonen P, Jarvinen O, Taskinen P, Auvinen T,
Tarkka M, Hippelainen M, Juvonen T, Hartikainen J, Hakala
T. Corticosteroids for the prevention of atrial fibrillation after
cardiac surgery: a randomized controlled trial. JAMA 297: 1562–
1567, 2007.
219. Hanna IRHB, Bush H, Brosius L, King-Hageman D, Dudley SC
Jr, Beshai JF, Langberg JJ. Lipid-lowering drug use is associated
with reduced prevalence of atrial fibrillation in patients with left
ventricular systolic dysfunction. Heart Rhythm 3: 881– 886, 2006.
220. Hanna N, Cardin S, Leung TK, Nattel S. Differences in atrial
versus ventricular remodeling in dogs with ventricular tachypac-
ing-induced congestive heart failure. Cardiovasc Res 63: 236 –244,
2004.
221. Hanna NCS, Leung TK, Nattel S. Differences in atrial versus
ventricular remodeling in dogs with ventricular tachypacing-in-
duced congestive heart failure. Cardiovasc Res 63(2): ,
2004236 –244.
222. Hansson A, Holm M, Blomstrom P, Johansson R, Luhrs C,
Brandt J, Olsson SB. Right atrial free wall conduction velocity
and degree of anisotropy in patients with stable sinus rhythm
studied during open heart surgery. Eur Heart J 19: 293–300, 1998.
223. Hansson LLL, Ekbom T, Dahlöf B, Lanke J, Scherstén B,
Wester PO, Hedner T, de Faire U. Randomised trial of old, new
antihypertensive drugs in elderly patients: cardiovascular mortality
and morbidity the Swedish Trial in Old Patients with Hyperten-
sion-2 study. Lancet 354: 1751–1756, 1999.
224. Hansson LLL, Niskanen L, Lanke J, Hedner T, Niklason A,
Luomanmäki K, Dahlöf B, de Faire U, Mörlin C, Karlberg BE,
Wester PO, Björck JE. Effect of angiotensin-converting-enzyme
inhibition compared with conventional therapy on cardiovascular
morbidity and mortality in hypertension: the Captopril Prevention
Project (CAPPP) randomised trial. Lancet 353: 611– 616, 1999.
225. Harada A, Sasaki K, Fukushima T, Ikeshita M, Asano T, Yam-
auchi S, Tanaka S, Shoji T. Atrial activation during chronic atrial
fibrillation in patients with isolated mitral valve disease. Ann Tho-
rac Surg 61: 104 –112, 1996.
226. Harjai KJ, Mobarek SK, Cheirif J, Boulos LM, Murgo JP,
Abi-Samra F. Clinical variables affecting recovery of left atrial
mechanical function after cardioversion from atrial fibrillation. J
Am Coll Cardiol 30: 481– 486, 1997.
227. Hart RG, Benavente O, McBride R, Pearce LA. Antithrombotic
therapy to prevent stroke in patients with atrial fibrillation: a
meta-analysis. Ann Intern Med 131: 492–501, 1999.
228. Hassink RJ, Aretz HT, Ruskin J, Keane D. Morphology of atrial
myocardium in human pulmonary veins: a postmortem analysis in
patients with and without atrial fibrillation. J Am Coll Cardiol 42:
1108 –1114, 2003.
229. Hatem SN, Benardeau A, Rucker-Martin C, Marty I, de
Chamisso P, Villaz M, Mercadier JJ. Different compartments of
sarcoplasmic reticulum participate in the excitation-contraction
coupling process in human atrial myocytes. Circ Res 80: 345–353,
1997.
230. Haugan K, Miyamoto T, Takeishi Y, Kubota I, Nakayama J,
Shimojo H, Hirose M. Rotigaptide (ZP123) improves atrial con-
duction slowing in chronic volume overload-induced dilated atria.
Basic Clin Pharmacol Toxicol 99: 71–79, 2006.
231. Hayashi H, Wang C, Miyauchi Y, Omichi C, Pak HN, Zhou S,
Ohara T, Mandel WJ, Lin SF, Fishbein MC, Chen PS, Kara-
gueuzian HS. Aging-related increase to inducible atrial fibrillation
in the rat model. J Cardiovasc Electrophysiol 13: 801– 808, 2002.
232. Henning B, Wit AL. The time course of action potential repolar-
ization affects delayed afterdepolarization amplitude in atrial fibers
of the canine coronary sinus. Circ Res 55: 110 –115, 1984.
233. Henry WL, Morganroth J, Pearlman AS, Clark CE, Redwood
DR, Itscoitz SB, Epstein SE. Relation between echocardio-
graphically determined left atrial size and atrial fibrillation. Circu-
lation 53: 273–279, 1976.
234. Hertervig E, Kongstad O, Ljungstrom E, Olsson B, Yuan S.
Pulmonary vein potentials in patients with and without atrial fibril-
lation. Europace 10: 692– 697, 2008.
235. Hinescu ME, Gherghiceanu M, Mandache E, Ciontea SM,
Popescu LM. Interstitial Cajal-like cells (ICLC) in atrial myocar-
dium: ultrastructural and immunohistochemical characterization. J
Cell Mol Med 10: 243–257, 2006.
236. Hirose M, Laurita KR. Calcium-mediated triggered activity is an
underlying cellular mechanism of ectopy originating from the pul-
monary vein in dogs. Am J Physiol Heart Circ Physiol 292: H1861–
H1867, 2007.
237. Hirose M, Leatmanoratn Z, Laurita KR, Carlson MD. Partial
vagal denervation increases vulnerability to vagally induced atrial
fibrillation. J Cardiovasc Electrophysiol 13: 1272–1279, 2002.
238. Hirose M, Takeishi Y, Miyamoto T, Kubota I, Laurita KR,
Chiba S. Mechanism for atrial tachyarrhythmia in chronic volume
overload-induced dilated atria. J Cardiovasc Electrophysiol 16:
760 –769, 2005.
239. Hirose M, Takeishi Y, Niizeki T, Shimojo H, Nakada T, Kubota
I, Nakayama J, Mende U, Yamada M. Diacylglycerol kinase zeta
inhibits G
q-induced atrial remodeling in transgenic mice. Heart
Rhythm: Official J Heart Rhythm Soc 6: 78 – 84, 2009.
240. Hobbs FD, Fitzmaurice DA, Mant J, Murray E, Jowett S,
Bryan S, Raftery J, Davies M, Lip G. A randomised controlled
trial and cost-effectiveness study of systematic screening (targeted
and total population screening) versus routine practice for the
detection of atrial fibrillation in people aged 65 and over. The SAFE
study Health Technol Assess 9: iii-iv, ix-x, 1–74, 2005.
241. Hocini M, Ho SY, Kawara T, Linnenbank AC, Potse M, Shah D,
Jais P, Janse MJ, Haissaguerre M, De Bakker JM. Electrical
conduction in canine pulmonary veins: electrophysiological and
anatomic correlation. Circulation 105: 2442–2448, 2002.
242. Hodgson-Zingman DM, Karst ML, Zingman LV, Heublein DM,
Darbar D, Herron KJ, Ballew JD, de Andrade M, Burnett JC
Jr, Olson TM. Atrial natriuretic peptide frameshift mutation in
familial atrial fibrillation. N Engl J Med 359: 158 –165, 2008.
243. Hogan PM, Davis LD. Electrophysiological characteristics of ca-
nine atrial plateau fibers. Circ Res 28: 62–73, 1971.
244. Hohnloser SH, Kuck KH, Lilienthal J. Rhythm or rate control in
atrial fibrillation–Pharmacological Intervention in Atrial Fibrilla-
tion (PIAF): a randomised trial. Lancet 356: 1789 –1794, 2000.
245. Holm M, Johansson R, Brandt J, Luhrs C, Olsson SB. Epicar-
dial right atrial free wall mapping in chronic atrial fibrillation.
Documentation of repetitive activation with a focal spread–a hith-
erto unrecognised phenomenon in man. Eur Heart J 18: 290 –310,
1997.
246. Holm M, Pehrson S, Ingemansson M, Sornmo L, Johansson R,
Sandhall L, Sunemark M, Smideberg B, Olsson C, Olsson SB.
Non-invasive assessment of the atrial cycle length during atrial
fibrillation in man: introducing, validating and illustrating a new
ECG method. Cardiovasc Res 38: 69 – 81, 1998.
247. Hong CS, Cho MC, Kwak YG, Song CH, Lee YH, Lim JS, Kwon
YK, Chae SW, Kim do H. Cardiac remodeling and atrial fibrillation
in transgenic mice overexpressing junctin. FASEB J 16: 1310 –1312,
2002.
248. Hong CS, Kwon SJ, Cho MC, Kwak YG, Ha KC, Hong B, Li H,
Chae SW, Chai OH, Song CH, Li Y, Kim JC, Woo SH, Lee SY,
Lee CO, Kim do H. Overexpression of junctate induces cardiac
hypertrophy and arrhythmia via altered calcium handling. J Mol
Cell Cardiol 44: 672– 682, 2008.
249. 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.
250. Hou Y, Scherlag BJ, Lin J, Zhang Y, Lu Z, Truong K, Patterson
E, Lazzara R, Jackman WM, Po SS. Ganglionated plexi modulate
314 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
extrinsic cardiac autonomic nerve input: effects on sinus rate,
atrioventricular conduction, refractoriness, and inducibility of
atrial fibrillation. J Am Coll Cardiol 50: 61– 68, 2007.
251. Houben R, de Groot N, Allessie M. Analysis of fractionated atrial
fibrillation electrograms by wavelet decomposition. IEEE Trans
Biomed Eng 57: 1388 –1398, 2010.
252. Houben RP, de Groot N, Smeets J, Becker AE, Lindemans FW,
Allessie M. S-wave predominance of epicardial electrograms dur-
ing atrial fibrillation in humans: indirect evidence for a role of the
thin subepicardial layer. Heart Rhythm 1: 639 – 647, 2004.
253. Hove-Madsen L, Llach A, Bayes-Genis A, Roura S, Rodriguez
Font E, Aris A, Cinca J. Atrial fibrillation is associated with
increased spontaneous calcium release from the sarcoplasmic re-
ticulum in human atrial myocytes. Circulation 110: 1358 –1363,
2004.
254. Hsu LF, Jais P, Sanders P, Garrigue S, Hocini M, Sacher F,
Takahashi Y, Rotter M, Pasquie JL, Scavee C, Bordachar P,
Clementy J, Haissaguerre M. Catheter ablation for atrial fibrilla-
tion in congestive heart failure. N Engl J Med 351: 2373–2383, 2004.
255. Hughes M, Lip GY. Risk factors for anticoagulation-related bleeding
complications in patients with atrial fibrillation: a systematic review.
QJM 100: 599 – 607, 2007.
256. Hughes M, Lip GY. Stroke and thromboembolism in atrial fibrillation:
a systematic review of stroke risk factors, risk stratification schema
and cost effectiveness data. Thromb Haemost 99: 295–304, 2008.
257. Hunt MJ, Aru GM, Hayden MR, Moore CK, Hoit BD, Tyagi SC.
Induction of oxidative stress and disintegrin metalloproteinase in
human heart end-stage failure. Am J Physiol Lung Cell Mol Physiol
283: L239 –L245, 2002.
258. Huser J, Lipsius SL, Blatter LA. Calcium gradients during exci-
tation-contraction coupling in cat atrial myocytes. J Physiol 494:
641– 651, 1996.
259. Husser D, Cannom DS, Bhandari AK, Stridh M, Sornmo L,
Olsson SB, Bollmann A. Electrocardiographic characteristics of
fibrillatory waves in new-onset atrial fibrillation. Europace 9: 638 –
642, 2007.
260. Husser D, Stridh M, Cannom DS, Bhandari AK, Girsky MJ,
Kang S, Sornmo L, Bertil Olsson S, Bollmann A. Validation and
clinical application of time-frequency analysis of atrial fibrillation
electrocardiograms. J Cardiovasc Electrophysiol 18: 41– 46, 2007.
261. Hwang C, Karagueuzian HS, Chen PS. Idiopathic paroxysmal
atrial fibrillation induced by a focal discharge mechanism in the left
superior pulmonary vein: possible roles of the ligament of Marshall.
J Cardiovasc Electrophysiol 10: 636 – 648, 1999.
262. Israel CW, Gronefeld G, Ehrlich JR, Li YG, Hohnloser SH.
Long-term risk of recurrent atrial fibrillation as documented by an
implantable monitoring device: implications for optimal patient
care. J Am Coll Cardiol 43: 47–52, 2004.
263. Jacquemet V, Virag N, Ihara Z, Dang L, Blanc O, Zozor S,
Vesin JM, Kappenberger L, Henriquez C. Study of unipolar
electrogram morphology in a computer model of atrial fibrillation.
J Cardiovasc Electrophysiol 14: S172–179, 2003.
264. January CT, Riddle JM. Early afterdepolarizations: mechanism
of induction and block. A role for L-type Ca
2
current. Circ Res 64:
977–990, 1989.
265. Jin W, Lu Z. Synthesis of a stable form of tertiapin: a high-affinity
inhibitor for inward-rectifier K
channels. Biochemistry 38: 14286 –
14293, 1999.
266. Johnson JN, Tester DJ, Perry J, Salisbury BA, Reed CR,
Ackerman MJ. Prevalence of early-onset atrial fibrillation in con-
genital long QT syndrome. Heart Rhythm 5: 704 –709, 2008.
267. Johnson N, Danilo P Jr, Wit AL, Rosen MR. Characteristics of
initiation and termination of catecholamine-induced triggered ac-
tivity in atrial fibers of the coronary sinus. Circulation 74: 1168–
1179, 1986.
268. Jongbloed MR, Schalij MJ, Poelmann RE, Blom NA, Fekkes
ML, Wang Z, Fishman GI, Gittenberger-De Groot AC. Embry-
onic conduction tissue: a spatial correlation with adult arrhythmo-
genic areas. J Cardiovasc Electrophysiol 15: 349 –355, 2004.
269. Joyner RW, Wang YG, Wilders R, Golod DA, Wagner MB,
Kumar R, Goolsby WN. A spontaneously active focus drives a
model atrial sheet more easily than a model ventricular sheet. Am
J Physiol Heart Circ Physiol 279: H752–H763, 2000.
270. Joyner RW, Wilders R, Wagner MB. Propagation of pacemaker
activity. Med Biol Eng Comput 45: 177–187, 2007.
271. Kalman JM, Munawar M, Howes LG, Louis WJ, Buxton BF,
Gutteridge G, Tonkin AM. Atrial fibrillation after coronary artery
bypass grafting is associated with sympathetic activation. Ann
Thoracic Surg 60: 1709 –1715, 1995.
272. Kanagaratnam P, Cherian A, Stanbridge RD, Glenville B, Sev-
ers NJ, Peters NS. Relationship between connexins and atrial
activation during human atrial fibrillation. J Cardiovasc Electro-
physiol 15: 206 –216, 2004.
273. Kanagaratnam P, Kojodjojo P, Peters NS. Electrophysiological
abnormalities occur prior to the development of clinical episodes
of atrial fibrillation: observations from human epicardial mapping.
Pacing Clin Electrophysiol 31: 443– 453, 2008.
274. Kanagaratnam P, Rothery S, Patel P, Severs NJ, Peters NS.
Relative expression of immunolocalized connexins 40 and 43 cor-
relates with human atrial conduction properties. J Am Coll Cardiol
39: 116 –123, 2002.
275. Kannel WB, Wolf PA, Benjamin EJ, Levy D. Prevalence, inci-
dence, prognosis, and predisposing conditions for atrial fibrillation:
population-based estimates. Am J Cardiol 82: 2N–9N, 1998.
276. Kato T, Yamashita T, Sagara K, Iinuma H, Fu LT. Progressive
nature of paroxysmal atrial fibrillation. Observations from a 14-year
follow-up study. Circ J 68: 568 –572, 2004.
277. Katra RP, Laurita KR. Cellular mechanism of calcium-mediated
triggered activity in the heart. Circ Res 96: 535–542, 2005.
278. Katritsis D, Giazitzoglou E, Sougiannis D, Voridis E, Po SS.
Complex fractionated atrial electrograms at anatomic sites of gan-
glionated plexi in atrial fibrillation. Europace 11: 308 –315, 2009.
279. Katritsis D, Ioannidis JP, Giazitzoglou E, Korovesis S, Anag-
nostopoulos CE, Camm AJ. Conduction delay within the coro-
nary sinus in humans: implications for atrial arrhythmias. J Car-
diovasc Electrophysiol 13: 859 – 862, 2002.
280. Katritsis DG. The coronary sinus: passive bystander or source of
arrhythmia? Heart Rhythm 1: 113–116, 2004.
281. Kawara T, Derksen R, de Groot JR, Coronel R, Tasseron S,
Linnenbank AC, Hauer RN, Kirkels H, Janse MJ, de Bakker
JM. Activation delay after premature stimulation in chronically
diseased human myocardium relates to the architecture of inter-
stitial fibrosis. Circulation 104: 3069 –3075, 2001.
282. Kawashima T. The autonomic nervous system of the human heart
with special reference to its origin, course, and peripheral distri-
bution. Anat Embryol 209: 425– 438, 2005.
283. Kerr CR, Humphries KH, Talajic M, Klein G, Connolly SJ,
Green M, Boone J, Sheldon R, Dorian P, Newman D. Progres-
sion to chronic atrial fibrillation after the initial diagnosis of par-
oxysmal atrial fibrillation: results from the Canadian Registry of
Atrial Fibrillation. Am Heart J 149: 489 – 496, 2005.
284. Khan MN, Jais P, Cummings J, Di Biase L, Sanders P, Martin
DO, Kautzner J, Hao S, Themistoclakis S, Fanelli R, Potenza
D, Massaro R, Wazni O, Schweikert R, Saliba W, Wang P,
Al-Ahmad A, Beheiry S, Santarelli P, Starling RC, Dello Russo
A, Pelargonio G, Brachmann J, Schibgilla V, Bonso A, Casella
M, Raviele A, Haissaguerre M, Natale A. Pulmonary-vein isola-
tion for atrial fibrillation in patients with heart failure. N Engl J Med
359: 1778 –1785, 2008.
285. Kholova I, Kautzner J. Morphology of atrial myocardial exten-
sions into human caval veins: a postmortem study in patients with
and without atrial fibrillation. Circulation 110: 483– 488, 2004.
286. Kieny JR, Sacrez A, Facello A, Arbogast R, Bareiss P, Roul G,
Demangeat JL, Brunot B, Constantinesco A. Increase in radio-
nuclide left ventricular ejection fraction after cardioversion of
chronic atrial fibrillation in idiopathic dilated cardiomyopathy. Eur
Heart J 13: 1290 –1295, 1992.
287. Kim YM, Guzik TJ, Zhang YH, Zhang MH, Kattach H, Ratna-
tunga C, Pillai R, Channon KM, Casadei B. A myocardial Nox2
containing NAD(P)H oxidase contributes to oxidative stress in
human atrial fibrillation. Circ Res 97: 629 – 636, 2005.
288. Kirchhof C, Chorro F, Scheffer GJ, Brugada J, Konings K,
Zetelaki Z, Allessie M. Regional entrainment of atrial fibrillation
studied by high-resolution mapping in open-chest dogs. Circula-
tion 88: 736 –749, 1993.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 315
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
289. Kirchhof P, Bax J, Blomstrom-Lundquist C, Calkins H, Camm
AJ, Cappato R, Cosio F, Crijns H, Diener HC, Goette A, Israel
CW, Kuck KH, Lip GY, Nattel S, Page RL, Ravens U, Schotten
U, Steinbeck G, Vardas P, Waldo A, Wegscheider K, Willems S,
Breithardt G. Early and comprehensive management of atrial
fibrillation: executive summary of the proceedings from the 2nd
AFNET-EHRA consensus conference “research perspectives in
AF.” Eur Heart J 30: 2969 –2977, 2009.
290. Kirchhof P, Eckardt L, Franz MR, Mönnig G, Loh P, Wedekind
H, Schulze-Bahr E, Breithardt G, Haverkamp W. Prolonged
atrial action potential durations and polymorphic atrial tachyar-
rhythmias in patients with long QT syndrome. J Cardiovasc Elec-
trophysiol 14: 1027–1033, 2003.
291. Kirchhof P, Eckardt L, Monnig G, Johna R, Loh P, Schulze-
Bahr E, Breithardt G, Borggrefe M, Haverkamp W. A patient
with “atrial torsades de pointes.” J Cardiovasc Electrophysiol 11:
806 – 811, 2000.
292. Kistler PM, Sanders P, Dodic M, Spence SJ, Samuel CS, Zhao
C, Charles JA, Edwards GA, Kalman JM. Atrial electrical and
structural abnormalities in an ovine model of chronic blood pres-
sure elevation after prenatal corticosteroid exposure: implications
for development of atrial fibrillation. Eur Heart J 27: 3045–3056,
2006.
293. Kistler PM, Sanders P, Fynn SP, Stevenson IH, Spence SJ,
Vohra JK, Sparks PB, Kalman JM. Electrophysiologic and elec-
troanatomic changes in the human atrium associated with age. J
Am Coll Cardiol 44: 109 –116, 2004.
294. Klein G, Schroder F, Vogler D, Schaefer A, Haverich A,
Schieffer B, Korte T, Drexler H. Increased open probability of
single cardiac L-type calcium channels in patients with chronic
atrial fibrillation. Role of phosphatase 2A. Cardiovasc Res 59:
37– 45, 2003.
295. Kneller J, Sun H, Leblanc N, Nattel S. Remodeling of Ca
2
-
handling by atrial tachycardia: evidence for a role in loss of rate-
adaptation. Cardiovasc Res 54: 416 – 426, 2002.
296. Kneller J, Zou R, Vigmond EJ, Wang Z, Leon LJ, Nattel S.
Cholinergic atrial fibrillation in a computer model of a two-dimen-
sional sheet of canine atrial cells with realistic ionic properties.
Circ Res 90: E73–E87, 2002.
297. Kockskamper J, Sheehan KA, Bare DJ, Lipsius SL, Mignery
GA, Blatter LA. Activation and propagation of Ca
2
release dur-
ing excitation-contraction coupling in atrial myocytes. Biophys J
81: 2590 –2605, 2001.
298. Konings K, Kirchhof C, Smeets J, Wellens HJ, Penn OC,
Allessie M. High-density mapping of electrically induced atrial
fibrillation in humans. Circulation 89: 1665–1680, 1994.
299. Konings K, Smeets J, Penn OC, Wellens HJ, Allessie M. Con-
figuration of unipolar atrial electrograms during electrically in-
duced atrial fibrillation in humans. Circulation 95: 1231–1241,
1997.
300. Korantzopoulos PKT, Galaris D, Goudevenos JA. The role of
oxidative stress in the pathogenesis and perpetuation of atrial
fibrillation. Int J Cardiol 115: 135–143, 2001.
301. Kostin S, Klein G, Szalay Z, Hein S, Bauer EP, Schaper J.
Structural correlate of atrial fibrillation in human patients. Cardio-
vasc Res 54: 361–379, 2002.
302. Koumi S, Arentzen CE, Backer CL, Wasserstrom JA. Alter-
ations in muscarinic K
channel response to acetylcholine and to G
protein-mediated activation in atrial myocytes isolated from failing
human hearts. Circulation 90: 2213–2224, 1994.
303. Koumi SI, Martin RL, Sato R. Alterations in ATP-sensitive po-
tassium channel sensitivity to ATP in failing human hearts. Am J
Physiol Heart Circ Physiol 272: H1656 –H1665, 1997.
304. Koura T, Hara M, Takeuchi S, Ota K, Okada Y, Miyoshi S,
Watanabe A, Shiraiwa K, Mitamura H, Kodama I, Ogawa S.
Anisotropic conduction properties in canine atria analyzed by high-
resolution optical mapping. Circulation 105: 2092–2098, 2002.
305. Kovoor P, Wickman K, Maguire CT, Pu W, Gehrmann J, Berul
CI, Clapham DE. Evaluation of the role of I(KACh) in atrial
fibrillation using a mouse knockout model. J Am Coll Cardiol 37:
2136 –2143, 2001.
306. Kubalova Z, Terentyev D, Viatchenko-Karpinski S, Nishijima
Y, Gyorke I, Terentyeva R, da Cunha DN, Sridhar A, Feldman
DS, Hamlin RL, Carnes CA, Gyorke S. Abnormal intrastore
calcium signaling in chronic heart failure. Proc Natl Acad Sci USA
102: 14104 –14109, 2005.
307. Kumagai K, Khrestian C, Waldo AL. Simultaneous multisite
mapping studies during induced atrial fibrillation in the sterile
pericarditis model. Insights into the mechanism of its maintenance.
Circulation 95: 511–521, 1997.
308. Kumagai K, Nakashima H, Saku K. The HMG-CoA reductase
inhibitor atorvastatin prevents atrial fibrillation by inhibiting in-
flammation in a canine sterile pericarditis model. Cardiovasc Res
62: 105–111, 2004.
309. Kumagai K, Nakashima H, Urata H, Gondo N, Arakawa K,
Saku K. Effects of angiotensin II type 1 receptor antagonist on
electrical and structural remodeling in atrial fibrillation. J Am Coll
Cardiol 41: 2197–2204, 2003.
310. Kumagai K, Ogawa M, Noguchi H, Yasuda T, Nakashima H,
Saku K. Electrophysiologic properties of pulmonary veins as-
sessed using a multielectrode basket catheter. J Am Coll Cardiol
43: 2281–2289, 2004.
311. Kusano KF, Taniyama M, Nakamura K, Miura D, Banba K,
Nagase S, Morita H, Nishii N, Watanabe A, Tada T, Murakami
M, Miyaji K, Hiramatsu S, Nakagawa K, Tanaka M, Miura A,
Kimura H, Fuke S, Sumita W, Sakuragi S, Urakawa S, Iwasaki
J, Ohe T. Atrial fibrillation in patients with Brugada syndrome
relationships of gene mutation, electrophysiology, and clinical
backgrounds. J Am Coll Cardiol 51: 1169 –1175, 2008.
312. L’Allier PLDA, Keller PF, Yu H, Guertin MC, Tardif JC. An-
giotensin-converting enzyme inhibition in hypertensive patients is
associated with a reduction in the occurrence of atrial fibrillation.
J Am Coll Cardiol 44: 159 –164, 2004.
313. Lazar S, Dixit S, Callans DJ, Lin D, Marchlinski FE, Gersten-
feld EP. Effect of pulmonary vein isolation on the left-to-right
atrial dominant frequency gradient in human atrial fibrillation.
Heart Rhythm 3: 889 – 895, 2006.
314. Lazar S, Dixit S, Marchlinski FE, Callans DJ, Gerstenfeld EP.
Presence of left-to-right atrial frequency gradient in paroxysmal but
not persistent atrial fibrillation in humans. Circulation 110: 3181–
3186, 2004.
315. Le Grand BL, Hatem S, Deroubaix E, Couetil JP, Coraboeuf
E. Depressed transient outward and calcium currents in dilated
human atria. Cardiovasc Res 28: 548 –556, 1994.
316. Lechleiter JD, John LM, Camacho P. Ca
2
wave dispersion and
spiral wave entrainment in Xenopus laevis oocytes overexpressing
Ca
2
ATPases. Biophys Chem 72: 123–129, 1998.
317. Lee CH, Liu PY, Lin LJ, Chen JH, Tsai LM. Clinical character-
istics and outcomes of hypertrophic cardiomyopathy in Taiwan: a
tertiary center experience. Clin Cardiol 30: 177–182, 2007.
318. Lee KW, Everett TH, Rahmutula D, Guerra JM, Wilson E,
Ding C, Olgin JE. Pirfenidone prevents the development of a
vulnerable substrate for atrial fibrillation in a canine model of heart
failure. Circulation 114: 1703–1712, 2006.
319. Leistad E, Aksnes G, Verburg E, Christensen G. Atrial contrac-
tile dysfunction after short-term atrial fibrillation is reduced by
verapamil but increased by BAY K8644. Circulation 93: 1747–1754,
1996.
320. Lellouche N, Buch E, Celigoj A, Siegerman C, Cesario D, De
Diego C, Mahajan A, Boyle NG, Wiener I, Garfinkel A, Shiv-
kumar K. Functional characterization of atrial electrograms in
sinus rhythm delineates sites of parasympathetic innervation in
patients with paroxysmal atrial fibrillation. J Am Coll Cardiol 50:
1324 –1331, 2007.
321. Lemola K, Chartier D, Yeh YH, Dubuc M, Cartier R, Armour A,
Ting M, Sakabe M, Shiroshita-Takeshita A, Comtois P, Nattel
S. Pulmonary vein region ablation in experimental vagal atrial
fibrillation: role of pulmonary veins versus autonomic ganglia. Cir-
culation 117: 470 – 477, 2008.
322. Lenaerts I, Bito V, Heinzel FR, Driesen RB, Holemans P,
D’Hooge J, Heidbuchel H, Sipido KR, Willems R. Ultrastruc-
tural and functional remodeling of the coupling between Ca
2
influx and sarcoplasmic reticulum Ca
2
release in right atrial myo-
cytes from experimental persistent atrial fibrillation. Circ Res 105:
876 – 885, 2009.
316 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
323. Leone O, Boriani G, Chiappini B, Pacini D, Cenacchi G, Mar-
tin Suarez S, Rapezzi C, Bacchi Reggiani ML, Marinelli G.
Amyloid deposition as a cause of atrial remodelling in persistent
valvular atrial fibrillation. Eur Heart J 25: 1237–1241, 2004.
324. Lertsburapa KWC, Kluger J, Faheem O, Hammond J,
Coleman CI. Preoperative statins for the prevention of atrial
fibrillation after cardiothoracic surgery. J Thorac Cardiovasc Surg
135: 405– 411, 2008.
325. Levy S. Epidemiology and classification of atrial fibrillation. J
Cardiovasc Electrophysiol 9: S78 – 82, 1998.
326. Li D, Fareh S, Leung TK, Nattel S. Promotion of atrial fibrillation
by heart failure in dogs: atrial remodeling of a different sort.
Circulation 100: 87–95, 1999.
327. Li D, Melnyk P, Feng J, Wang Z, Petrecca K, Shrier A, Nattel
S. Effects of experimental heart failure on atrial cellular and ionic
electrophysiology. Circulation 101: 2631–2638, 2000.
328. Li D, Shinagawa K, Pang L, Leung TK, Cardin S, Wang Z,
Nattel S. Effects of angiotensin-converting enzyme inhibition on
the development of the atrial fibrillation substrate in dogs with
ventricular tachypacing-induced congestive heart failure. Circula-
tion 104: 2608 –2614, 2001.
330. Li GR, Feng J, Wang Z, Fermini B, Nattel S. Adrenergic mod-
ulation of ultrarapid delayed rectifier K
current in human atrial
myocytes. Circ Res 78: 903–915, 1996.
331. Li GR, Nattel S. Properties of human atrial I
Ca
at physiological
temperatures and relevance to action potential. Am J Physiol
Heart Circ Physiol 272: H227–H235, 1997.
332. Li J, McLerie M, Lopatin AN. Transgenic upregulation of IK1 in
the mouse heart leads to multiple abnormalities of cardiac excit-
ability. Am J Physiol Heart Circ Physiol 287: H2790 –H2802, 2004.
333. Li N, Timofeyev V, Tuteja D, Xu D, Lu L, Zhang Q, Zhang Z,
Singapuri A, Albert TR, Rajagopal AV, Bond CT, Periasamy
M, Adelman J, Chiamvimonvat N. Ablation of a Ca
2
-activated
K
channel (SK2 channel) results in action potential prolongation
in atrial myocytes and atrial fibrillation. J Physiol 587: 1087–1100,
2009.
334. Liang BT, Frame LH, Molinoff PB. Beta 2-adrenergic receptors
contribute to catecholamine-stimulated shortening of action poten-
tial duration in dog atrial muscle. Proc Natl Acad Sci USA 82:
4521– 4525, 1985.
335. Liang L, Pan Q, Liu Y, Chen H, Li J, Brugada R, Brugada P,
Hong K, Perez GJ, Zhao C, Qi J, Zhang Y, Peng L, Li L, Chen
YH. High sensitivity of the sheep pulmonary vein antrum to ace-
tylcholine stimulation. J Appl Physiol 105: 293–298, 2008.
336. Lie JT, Hammond PI. Pathology of the senescent heart: anatomic
observations on 237 autopsy studies of patients 90 to 105 years old.
Mayo Clin Proc 63: 552–564, 1988.
337. Lin CCLJ, Lin CS, Tsai MC, Su MJ, Lai LP, Huang SK. Activa-
tion of the calcineurin-nuclear factor of activated T-cell signal
transduction pathway in atrial fibrillation. Chest 126: 1926 –1932,
2004.
338. Lin J, Scherlag BJ, Zhou J, Lu Z, Patterson E, Jackman WM,
Lazzara R, Po SS. Autonomic mechanism to explain complex
fractionated atrial electrograms (CFAE). J Cardiovasc Electro-
physiol 18: 1197–1205, 2007.
339. Lin YJ, Tai CT, Kao T, Chang SL, Wongcharoen W, Lo LW,
Tuan TC, Udyavar AR, Chen YJ, Higa S, Ueng KC, Chen SA.
Consistency of complex fractionated atrial electrograms during
atrial fibrillation. Heart Rhythm 5: 406 – 412, 2008.
340. Lin YJ, Tai CT, Kao T, Tso HW, Higa S, Tsao HM, Chang SL,
Hsieh MH, Chen SA. Frequency analysis in different types of
paroxysmal atrial fibrillation. J Am Coll Cardiol 47: 1401–1407,
2006.
341. Lipkin DP, Frenneaux M, Stewart R, Joshi J, Lowe T, McK-
enna WJ. Delayed improvement in exercise capacity after cardio-
version of atrial fibrillation to sinus rhythm. Br Heart J 59: 572–577,
1988.
342. Lipp P, Pott L, Callewaert G, Carmeliet E. Simultaneous re-
cording of Indo-1 fluorescence and Na
/Ca
2
exchange current
reveals two components of Ca
2
-release from sarcoplasmic retic-
ulum of cardiac atrial myocytes. FEBS Lett 275: 181–184, 1990.
343. Liu L, Nattel S. Differing sympathetic and vagal effects on atrial
fibrillation in dogs: role of refractoriness heterogeneity. Am J
Physiol Heart Circ Physiol 273: H805–H816, 1997.
344. Logan WF, Rowlands DJ, Howitt G, Holmes AM. Left atrial
activity following cardioversion. Lancet 2: 471– 473, 1965.
345. Lombardi R, Rodriguez G, Chen SN, Ripplinger CM, Li W,
Chen J, Willerson JT, Betocchi S, Wickline SA, Efimov IR,
Marian AJ. Resolution of established cardiac hypertrophy and
fibrosis and prevention of systolic dysfunction in a transgenic
rabbit model of human cardiomyopathy through thiol-sensitive
mechanisms. Circulation 119: 1398 –1407, 2009.
346. Lopatin AN, Nichols CG. Inward rectifiers in the heart: an update
on I(K1). J Mol Cell Cardiol 33: 625– 638, 2001.
347. Losi MA, Betocchi S, Aversa M, Lombardi R, Miranda M,
D’Alessandro G, Cacace A, Tocchetti CG, Barbati G, Chiari-
ello M. Determinants of atrial fibrillation development in patients
with hypertrophic cardiomyopathy. Am J Cardiol 94: 895–900,
2004.
348. Lowe JE, Hendry PJ, Hendrickson SC, Wells R. Intraoperative
identification of cardiac patients at risk to develop postoperative
atrial fibrillation. Ann Surg 213: 388 –391, 1991.
349. Lu L, Zhang Q, Timofeyev V, Zhang Z, Young JN, Shin HS,
Knowlton AA, Chiamvimonvat N. Molecular coupling of a Ca
2
-
activated K
channel to L-type Ca
2
channels via alpha-actinin2.
Circ Res 100: 112–120, 2007.
350. Lu Z, Scherlag BJ, Lin J, Niu G, Ghias M, Jackman WM,
Lazzara R, Jiang H, Po SS. Autonomic mechanism for complex
fractionated atrial electrograms: evidence by fast fourier transform
analysis. J Cardiovasc Electrophysiol 19: 835– 842, 2008.
351. Lundby A, Ravn LS, Svendsen JH, Olesen SP, Schmitt N.
KCNQ1 mutation Q147R is associated with atrial fibrillation and
prolonged QT interval. Heart Rhythm 4: 1532–1541, 2007.
352. Luo MH, Li Y, Yang KP. Fibrosis of collagen i and remodeling of
connexin 43 in atrial myocardium of patients with atrial fibrillation.
Cardiology 107: 248 –253, 2006.
353. Mackenzie L, Bootman MD, Berridge MJ, Lipp P. Predeter-
mined recruitment of calcium release sites underlies excitation-
contraction coupling in rat atrial myocytes. J Physiol 530: 417– 429,
2001.
354. Mackenzie L, Roderick HL, Berridge MJ, Conway SJ, Boot-
man MD. The spatial pattern of atrial cardiomyocyte calcium
signalling modulates contraction. J Cell Sci 117: 6327– 6337, 2004.
355. Maggioni APLR, Carson PE, Singh SN, Barlera S, Glazer R, Masson
S, Cerè E, Tognoni G, Cohn JN, Val-HeFT Investigators. Valsartan
reduces the incidence of atrial fibrillation in patients with heart failure:
results from the Valsartan Heart Failure Trial (Val-HeFT). Am Heart J
149: 548 –557, 2005.
356. Mandapati R, Skanes A, Chen J, Berenfeld O, Jalife J. Stable
microreentrant sources as a mechanism of atrial fibrillation in the
isolated sheep heart. Circulation 101: 194 –199, 2000.
357. Manios EG, Kanoupakis EM, Chlouverakis GI, Kaleboubas
MD, Mavrakis HE, Vardas PE. Changes in atrial electrical prop-
erties following cardioversion of chronic atrial fibrillation: relation
with recurrence. Cardiovasc Res 47: 244 –253, 2000.
358. Manning WJ, Silverman DI, Katz SE, Riley MF, Come PC,
Doherty RM, Munson JT, Douglas PS. Impaired left atrial me-
chanical function after cardioversion: relation to the duration of
atrial fibrillation. J Am Coll Cardiol 23: 1535–1540, 1994.
359. Manning WJ, Silverman DI, Katz SE, Riley MF, Doherty RM,
Munson JT, Douglas PS. Temporal dependence of the return of
atrial mechanical function on the mode of cardioversion of atrial
fibrillation to sinus rhythm. Am J Cardiol 75: 624 – 626, 1995.
360. Mansour M, Mandapati R, Berenfeld O, Chen J, Samie FH,
Jalife J. Left-to-right gradient of atrial frequencies during acute
atrial fibrillation in the isolated sheep heart. Circulation 103: 2631–
2636, 2001.
361. Mansourati J, Le Grand B. Transient outward current in young
and adult diseased human atria. Am J Physiol Heart Circ Physiol
265: H1466 –H1470, 1993.
362. Marín FPD, Roldán V, Arribas JM, Ahumada M, Tornel PL,
Oliver C, Gómez-Plana J, Lip GY, Valdés M. Statins and post-
operative risk of atrial fibrillation following coronary artery bypass
grafting. Am J Cardiol 97: 55– 60, 2006.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 317
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
363. Marin FRV, Climent V, Garcia A, Marco P, Lip GYH. Is throm-
bogenesis in atrial fibrillation related to matrix metalloproteinase-1
and ist inhibitor, TIMP-1? Stroke 34: 1181–1186, 2003.
364. Markides V, Schilling RJ, Ho SY, Chow AW, Davies DW, Pe-
ters NS. Characterization of left atrial activation in the intact
human heart. Circulation 107: 733–739, 2003.
365. Maron BJ, Olivotto I, Bellone P, Conte MR, Cecchi F, Flygen-
ring BP, Casey SA, Gohman TE, Bongioanni S, Spirito P.
Clinical profile of stroke in 900 patients with hypertrophic cardio-
myopathy. J Am Coll Cardiol 39: 301–307, 2002.
366. Maron BJ, Towbin JA, Thiene G, Antzelevitch C, Corrado D, Ar-
nett D, Moss AJ, Seidman CE, Young JB. Contemporary definitions
and classification of the cardiomyopathies: an American Heart Associa-
tion Scientific Statement from the Council on Clinical Cardiology, Heart
Failure and Transplantation Committee; Quality of Care and Outcomes
Research and Functional Genomics and Translational Biology Interdisci-
plinary Working Groups; and Council on Epidemiology and Prevention.
Circulation 113: 1807–1816, 2006.
367. Marrero MBSB, Paxton WG. Direct stimulation of the Jak/STAT
pathway by the angiotensin II AT1 receptor. Nature 375: 247–250,
1995.
368. Masani F. Node-like cells in the myocardial layer of the pulmonary
vein of rats: an ultrastructural study. J Anat 145: 133–142, 1986.
369. Matsuda T, Masumiya H, Tanaka N, Yamashita T, Tsuruzoe N,
Tanaka Y, Tanaka H, Shigenoba K. Inhibition by a novel anti-
arrhythmic agent, NIP-142, of cloned human cardiac K
channel
Kv1.5 current. Life Sci 68: 2017–2024, 2001.
370. Matsuo K, Uno K, Khrestian CM, Waldo AL. Conduction left-
to-right and right-to-left across the crista terminalis. Am J Physiol
Heart Circ Physiol 280: H1683–H1691, 2001.
371. Maupoil V, Bronquard C, Freslon JL, Cosnay P, Findlay I.
Ectopic activity in the rat pulmonary vein can arise from simulta-
neous activation of alpha1- and beta1-adrenoceptors. Br J Phar-
macol 150: 899 –905, 2007.
372. Mayer AG. Rhythmical Pulsation in Scyphomedusae. Washing-
ton, DC: Carnegie Institute of Washington, 1906, p. 1– 62.
373. Mays DJ, Foose JM, Philipson LH, Tamkun MM. Localization of
the Kv1.5 K
channel protein in explanted cardiac tissue. J Clin
Invest 96: 282–292, 1995.
374. Maytin O, Castillo C, Castellanos A Jr. Marked sinus tachycar-
dia and overdrive suppression. Chest 58: 528 –530, 1970.
375. Melnyk P, Ehrlich JR, Pourrier M, Villeneuve L, Cha TJ,
Nattel S. Comparison of ion channel distribution and expression
in cardiomyocytes of canine pulmonary veins versus left atrium.
Cardiovasc Res 65: 104 –116, 2005.
376. Merckx KL, De Vos CB, Palmans A, Habets J, Cheriex EC,
Crijns HJ, Tieleman RG. Atrial activation time determined by
transthoracic Doppler tissue imaging can be used as an estimate of
the total duration of atrial electrical activation. J Am Soc Echocar-
diogr 18: 940 –944, 2005.
377. Merckx KLTR, Folkeringa RJ, Pinto YM, Nieman FH, Cheriex
EC, Crijns HJ. Use of statins is associated with reduced incidence
of atrial fibrillation in patients with left vetricular hypertrophy and
left atrial dilatation (Abstract). Heart Rhythm 1: S104, 2004.
378. Michelucci A, Padeletti L, Fradella GA. Atrial refractoriness
and spontaneous or induced atrial fibrillation. Acta Cardiol 37:
333–344, 1982.
379. Mihm MJ, Yu F, Carnes CA, Reiser PJ, McCarthy PM, Van
Wagoner DR, Bauer JA. Impaired myofibrillar energetics and
oxidative injury during human atrial fibrillation. Circulation 104:
174 –180, 2001.
380. Milliez P, Deangelis N, Rucker-Martin C, Leenhardt A, Vicaut
E, Robidel E, Beaufils P, Delcayre C, Hatem SN. Spironolac-
tone reduces fibrosis of dilated atria during heart failure in rats
with myocardial infarction. Eur Heart J 26: 2193–2199, 2005.
381. Mines GR. On circulating excitation on heart muscles and their
possible relation to tachycardia and fibrillation. Trans R Soc Can 4:
43–53, 1914.
382. Ming Z, Nordin C, Aronson RS. Role of L-type calcium channel
window current in generating current-induced early afterdepolar-
izations. J Cardiovasc Electrophysiol 5: 323–334, 1994.
383. Misier AR, Opthof T, van Hemel NM, Defauw JJ, de Bakker
JM, Janse MJ, van Capelle FJ. Increased dispersion of “refrac-
toriness” in patients with idiopathic paroxysmal atrial fibrillation. J
Am Coll Cardiol 19: 1531–1535, 1992.
384. Miyasaka Y, Barnes ME, Gersh BJ, Cha SS, Seward JB, Bailey
KR, Iwasaka T, Tsang TS. Time trends of ischemic stroke inci-
dence and mortality in patients diagnosed with first atrial fibrilla-
tion in 1980 to 2000: report of a community-based study. Stroke 36:
2362–2366, 2005.
385. Modre R, Tilg B, Fischer G, Hanser F, Messnarz B, Seger M,
Schocke MF, Berger T, Hintringer F, Roithinger FX. Atrial
noninvasive activation mapping of paced rhythm data. J Cardio-
vasc Electrophysiol 14: 712–719, 2003.
386. Moe GK. A conceptual model of atrial fibrillation. J Electrocardiol
1: 145–146, 1968.
387. Moe GK, Abildskov JA. Atrial fibrillation as a self-sustaining
arrhythmia independent of focal discharge. Am Heart J 58: 59 –70,
1959.
388. Moe GK, Rheinboldt WC, Abildskov JA. A computer model of
atrial fibrillation. Am Heart J 67: 200 –220, 1964.
389. Molkentin JD. Calcineurin and Beyond. Cardiac hypertrophic
signalling. Circ Res 87: 731–738, 2000.
390. Moore EN, Spear JF. Electrophysiological studies on atrial fibril-
lation. Heart Vessels Suppl 2: 32–39, 1987.
391. Morel E, Meyronet D, Thivolet-Bejuy F, Chevalier P. Identifi-
cation and distribution of interstitial Cajal cells in human pulmo-
nary veins. Heart Rhythm 5: 1063–1067, 2008.
392. Morganroth J, Horowitz LN, Josephson ME, Kastor JA. Rela-
tionship of atrial fibrillatory wave amplitude to left atrial size and
etiology of heart disease. An old generalization re-examined. Am
Heart J 97: 184 –186, 1979.
393. Morillo CA, Klein G, Jones DL, Guiraudon C. Chronic rapid
atrial pacing. Structural, functional, and electrophysiological char-
acteristics of a new model of sustained atrial fibrillation. Circula-
tion 91: 1588 –1595, 1995.
394. Mozaffarian DPB, Rimm EB, Lemaitre RN, Burke GL, Lyles
MF, Lefkowitz D, Siscovick DS. Fish intake and risk of incident
atrial fibrillation. Circulation 110: 368 –373, 2004.
395. Muller FU, Lewin G, Baba HA, Boknik P, Fabritz L, Kirch-
hefer U, Kirchhof P, Loser K, Matus M, Neumann J, Riemann
B, Schmitz W. Heart-directed expression of a human cardiac
isoform of cAMP-response element modulator in transgenic mice. J
Biol Chem 280: 6906 – 6914, 2005.
396. Nabauer M, Gerth A, Limbourg T, Schneider S, Oeff M, Kirch-
hof P, Goette A, Lewalter T, Ravens U, Meinertz T, Breithardt
G, Steinbeck G. The Registry of the German Competence NET-
work on Atrial Fibrillation: patient characteristics and initial man-
agement. Europace 11: 423– 434, 2009.
397. Nademanee K, Lockwood E, Oketani N, Gidney B. Catheter
ablation of atrial fibrillation guided by complex fractionated atrial
electrogram mapping of atrial fibrillation substrate. J Cardiol 55:
1–12.
398. Nademanee K, McKenzie J, Kosar E, Schwab M, Sunsaneewi-
tayakul B, Vasavakul T, Khunnawat C, Ngarmukos T. A new
approach for catheter ablation of atrial fibrillation: mapping of the
electrophysiologic substrate. J Am Coll Cardiol 43: 2044–2053,
2004.
399. Nademanee K, Schwab M, Porath J, Abbo A. How to perform
electrogram-guided atrial fibrillation ablation. Heart Rhythm 3:
981–984, 2006.
400. Nademanee K, Schwab MC, Kosar EM, Karwecki M, Moran
MD, Visessook N, Michael AD, Ngarmukos T. Clinical outcomes
of catheter substrate ablation for high-risk patients with atrial
fibrillation. J Am Coll Cardiol 51: 843– 849, 2008.
401. Nagasawa H, Fujiki A, Fujikura N, Matsuda T, Yamashita T,
Inoue H. Effects of a novel class III antiarrhythmic agent, NIP-142,
on canine atrial fibrillation and flutter. Circ J 66: 185–191, 2002.
402. Nagy N, Szuts V, Horvath Z, Seprenyi G, Farkas AS, Acsai K,
Prorok J, Bitay M, Kun A, Pataricza J, Papp JG, Nanasi PP,
Varro A, Toth A. Does small-conductance calcium-activated po-
tassium channel contribute to cardiac repolarization? J Mol Cell
Cardiol 47: 656 – 663, 2009.
403. Nakao K, Seto S, Ueyama C, Matsuo K, Komiya N, Isomoto S,
Yano K. Extended distribution of prolonged and fractionated right
atrial electrograms predicts development of chronic atrial fibrilla-
318 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
tion in patients with idiopathic paroxysmal atrial fibrillation. J
Cardiovasc Electrophysiol 13: 996 –1002, 2002.
404. Nao T, Ohkusa T, Hisamatsu Y, Inoue N, Matsumoto T, Yamada J,
Shimizu A, Yoshiga Y, Yamagata T, Kobayashi S, Yano M, Hamano
K, Matsuzaki M. Comparison of expression of connexin in right atrial
myocardium in patients with chronic atrial fibrillation versus those in
sinus rhythm. Am J Cardiol 91: 678 – 683, 2003.
405. Narayan SM, Kazi D, Krummen DE, Rappel WJ. Repolarization
and activation restitution near human pulmonary veins and atrial
fibrillation initiation: a mechanism for the initiation of atrial fibril-
lation by premature beats. J Am Coll Cardiol 52: 1222–1230, 2008.
406. Narolska NA, Eiras S, van Loon RB, Boontje NM, Zaremba R,
Spiegelen Berg SR, Stooker W, Huybregts MA, Visser FC, van
der Velden J, Stienen GJ. Myosin heavy chain composition and
the economy of contraction in healthy and diseased human myo-
cardium. J Muscle Res Cell Motil 26: 39 – 48, 2005.
407. Nattel S. Defining “culprit mechanisms” in arrhythmogenic car-
diac remodeling. Circ Res 94: 1403–1405, 2004.
408. Nattel S, Maguy A, Le Bouter S, Yeh YH. Arrhythmogenic
ion-channel remodeling in the heart: heart failure, myocardial in-
farction, and atrial fibrillation. Physiol Rev 87: 425– 456, 2007.
409. Neef S, Dybkova N, Sossalla S, Ort KR, Fluschnik N, Neumann
K, Seipelt R, Schondube FA, Hasenfuss G, Maier LS. CaMKII-
dependent diastolic SR Ca
2
leak and elevated diastolic Ca
2
levels
in right atrial myocardium of patients with atrial fibrillation. Circ
Res 106: 1134 –1144, 2010.
410. Neuberger HR, Schotten U, Blaauw Y, Vollmann D, Eijsbouts
S, van Hunnik A, Allessie M. Chronic atrial dilation, electrical
remodeling, atrial fibrillation in the goat. J Am Coll Cardiol 47:
644 – 653, 2006.
411. Neuberger HR, Schotten U, Verheule S, Eijsbouts S, Blaauw
Y, van Hunnik A, Allessie M. Development of a substrate of atrial
fibrillation during chronic atrioventricular block in the goat. Cir-
culation 111: 30 –37, 2005.
412. Ng J, Goldberger JJ. Understanding and interpreting dominant
frequency analysis of AF electrograms. J Cardiovasc Electro-
physiol 18: 680 – 685, 2007.
413. Ng J, Kadish AH, Goldberger JJ. Effect of electrogram charac-
teristics on the relationship of dominant frequency to atrial activa-
tion rate in atrial fibrillation. Heart Rhythm 3: 1295–1305, 2006.
414. Ng J, Kadish AH, Goldberger JJ. Technical considerations for
dominant frequency analysis. J Cardiovasc Electrophysiol 18: 757–
764, 2007.
415. Nguyen BL, Fishbein MC, Chen LS, Chen PS, Masroor S.
Histopathological substrate for chronic atrial fibrillation in hu-
mans. Heart Rhythm 6: 454 – 460, 2009.
416. Nieuwlaat R, Capucci A, Camm AJ, Olsson SB, Andresen D,
Davies DW, Cobbe S, Breithardt G, Le Heuzey JY, Prins MH,
Levy S, Crijns HJ. Atrial fibrillation management: a prospective
survey in ESC member countries: the Euro Heart Survey on Atrial
Fibrillation. Eur Heart J 26: 2422–2434, 2005.
417. Nieuwlaat R, Prins MH, Le Heuzey JY, Vardas PE, Aliot E,
Santini M, Cobbe SM, Widdershoven JW, Baur LH, Levy S,
Crijns HJ. Prognosis, disease progression, and treatment of atrial
fibrillation patients during 1 year: follow-up of the Euro Heart
Survey on atrial fibrillation. Eur Heart J 29: 1181–1189, 2008.
418. Nishimaru K, Makuta R, Tanaka Y, Tanaka H, Shigenobu K.
Pharmacological properties of excitation-contraction mechanisms
in isolated mouse left atria. Pharmacology 62: 87–91, 2001.
419. Nitta T, Ishii Y, Miyagi Y, Ohmori H, Sakamoto S, Tanaka S.
Concurrent multiple left atrial focal activations with fibrillatory
conduction and right atrial focal or reentrant activation as the
mechanism in atrial fibrillation. J Thorac Cardiovasc Surg 127:
770 –778, 2004.
420. Ogawa M, Zhou S, Tan AY, Song J, Gholmieh G, Fishbein MC,
Luo H, Siegel RJ, Karagueuzian HS, Chen LS, Lin SF, Chen
PS. Left stellate ganglion and vagal nerve activity and cardiac
arrhythmias in ambulatory dogs with pacing-induced congestive
heart failure. J Am Coll Cardiol 50: 335–343, 2007.
421. Oh S, Zhang Y, Bibevski S, Marrouche NF, Natale A, Mazgalev
TN. Vagal denervation and atrial fibrillation inducibility: epicardial
fat pad ablation does not have long-term effects. Heart Rhythm 3:
701–708, 2006.
422. Ohtani K, Yutani C, Nagata S, Koretsune Y, Hori M, Kamada
T. High prevalence of atrial fibrosis in patients with dilated cardio-
myopathy. J Am Coll Cardiol 25: 1162–1169, 1995.
423. Okuyama Y, Miyauchi Y, Park AM, Hamabe A, Zhou S, Hayashi
H, Miyauchi M, Omichi C, Pak HN, Brodsky LA, Mandel WJ,
Fishbein MC, Karagueuzian HS, Chen PS. High resolution map-
ping of the pulmonary vein and the vein of Marshall during induced
atrial fibrillation and atrial tachycardia in a canine model of pacing-
induced congestive heart failure. J Am Coll Cardiol 42: 348–360,
2003.
424. Olgin JE, Sih HJ, Hanish S, Jayachandran JV, Wu J, Zheng
QH, Winkle W, Mulholland GK, Zipes DP, Hutchins G. Heter-
ogeneous atrial denervation creates substrate for sustained atrial
fibrillation. Circulation 98: 2608 –2614, 1998.
425. Olson TM, Michels VV, Ballew JD, Reyna SP, Karst ML, Her-
ron KJ, Horton SC, Rodeheffer RJ, Anderson JL. Sodium
channel mutations and susceptibility to heart failure and atrial
fibrillation. JAMA 293: 447– 454, 2005.
426. Olsson SB, Cotoi S, Varnauskas E. Monophasic action potential
and sinus rhythm stability after conversion of atrial fibrillation.
Acta Med Scand 190: 381–387, 1971.
427. Ono N, Hayashi H, Kawase A, Lin SF, Li H, Weiss JN, Chen PS,
Karagueuzian HS. Spontaneous atrial fibrillation initiated by trig-
gered activity near the pulmonary veins in aged rats subjected to
glycolytic inhibition. Am J Physiol Heart Circ Physiol 292: H639 –
H648, 2007.
428. Oral H, Chugh A, Yoshida K, Sarrazin JF, Kuhne M, Crawford
T, Chalfoun N, Wells D, Boonyapisit W, Veerareddy S, Billa-
kanty S, Wong WS, Good E, Jongnarangsin K, Pelosi F Jr,
Bogun F, Morady F. A randomized assessment of the incremental
role of ablation of complex fractionated atrial electrograms after
antral pulmonary vein isolation for long-lasting persistent atrial
fibrillation. J Am Coll Cardiol 53: 782–789, 2009.
429. Oral H, Ozaydin M, Tada H, Chugh A, Scharf C, Hassan S, Lai
S, Greenstein R, Pelosi F Jr, Knight BP, Strickberger SA,
Morady F. Mechanistic significance of intermittent pulmonary vein
tachycardia in patients with atrial fibrillation. J Cardiovasc Elec-
trophysiol 13: 645– 650, 2002.
430. Ortiz J, Niwano S, Abe H, Rudy Y, Johnson NJ, Waldo AL.
Mapping the conversion of atrial flutter to atrial fibrillation and
atrial fibrillation to atrial flutter. Insights into mechanisms. Circ Res
74: 882– 894, 1994.
431. Ozaydin MDA, Varol E, Kapan S, Tuzun N, Peker O, Aslan SM,
Altinbas A, Ocal A, Ibrisim E. Statin use before by-pass surgery
decreases the incidence and shortens the duration of postoperative
atrial fibrillation. Cardiology 107: 117–121, 2007.
432. Ozgen N, Dun W, Sosunov EA, Anyukhovsky EP, Hirose M,
Duffy HS, Boyden PA, Rosen MR. Early electrical remodeling in
rabbit pulmonary vein results from trafficking of intracellular SK2
channels to membrane sites. Cardiovasc Res 75: 758 –769, 2007.
433. Packer DL, Bardy GH, Worley SJ, Smith MS, Cobb FR,
Coleman RE, Gallagher JJ, German LD. Tachycardia-induced
cardiomyopathy: a reversible form of left ventricular dysfunction.
Am J Cardiol 57: 563–570, 1986.
434. Page PL, Dandan N, Savard P, Nadeau R, Armour JA, Cardinal
R. Regional distribution of atrial electrical changes induced by
stimulation of extracardiac and intracardiac neural elements. J
Thorac Cardiovasc Surg 109: 377–388, 1995.
435. Page PL, Plumb VJ, Okumura K, Waldo AL. A new animal model
of atrial flutter. J Am Coll Cardiol 8: 872– 879, 1986.
436. Pan NHTH, Chang NC, Chen YJ, Chen SA. Aging dilates atrium
and pulmonary veins: implications for the genesis of atrial fibrilla-
tion. Chest 133: 190 –196, 2008.
437. Pandozi C, Bianconi L, Villani M, Castro A, Altamura G,
Toscano S, Jesi AP, Gentilucci G, Ammirati F, Lo Bianco F,
Santini M. Local capture by atrial pacing in spontaneous chronic
atrial fibrillation. Circulation 95: 2416 –2422, 1997.
438. Panfilov AV, Pertsov AM. Mechanism of spiral waves initiation in
active media, connected with critical curvature phenomenon.
Biofizika 27: 886 – 889, 1982.
439. Parker TGSM. Peptide growth factors can provoke fetal contrac-
tile proteins gene expression in rat cardiac myocytes. J Clin Invest
85: 507–514, 1990.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 319
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
440. Patterson E, Jackman WM, Beckman KJ, Lazzara R, Lock-
wood D, Scherlag BJ, Wu R, Po S. Spontaneous pulmonary vein
firing in man: relationship to tachycardia-pause early afterdepolar-
izations and triggered arrhythmia in canine pulmonary veins in
vitro. J Cardiovasc Electrophysiol 18: 1067–1075, 2007.
441. Patterson E, Lazzara R, Szabo B, Liu H, Tang D, Li YH,
Scherlag BJ, Po SS. Sodium-calcium exchange initiated by the
Ca
2
transient: an arrhythmia trigger within pulmonary veins. JAm
Coll Cardiol 47: 1196 –1206, 2006.
442. Patterson E, Po SS, Scherlag BJ, Lazzara R. Triggered firing in
pulmonary veins initiated by in vitro autonomic nerve stimulation.
Heart Rhythm 2: 624 – 631, 2005.
443. Patti GCM, Candura D, Pasceri V, D’Ambrosio A, Covino E, Di
Sciascio G. Randomized trial of atorvastatin for reduction of post-
operative atrial fibrillation in patients undergoing cardiac surgery:
results of the ARMYDA-3 (Atorvastatin for Reduction of MYocar-
dial Dysrhythmia After cardiac surgery) study. Circulation 114:
1455–1461, 2006.
444. Pau D, Workman AJ, Kane KA, Rankin AC. Electrophysiologi-
cal and arrhythmogenic effects of 5-hydroxytryptamine on human
atrial cells are reduced in atrial fibrillation. J Mol Cell Cardiol 42:
54 – 62, 2007.
445. Pedersen OD, Bagger H, Kober L, Torp-Pedersen C. Trandola-
pril reduces the incidence of atrial fibrillation after acute myocar-
dial infarction in patients with left ventricular dysfunction. Circu-
lation 100: 376 –380, 1999.
446. Pehrson S, Holm M, Meurling C, Ingemansson M, Smideberg
B, Sornmo L, Olsson SB. Non-invasive assessment of magnitude
and dispersion of atrial cycle length during chronic atrial fibrilla-
tion in man. Eur Heart J 19: 1836 –1844, 1998.
447. Perez-Lugones A, McMahon JT, Ratliff NB, Saliba WI,
Schweikert RA, Marrouche NF, Saad EB, Navia JL, McCarthy
PM, Tchou P, Gillinov AM, Natale A. Evidence of specialized
conduction cells in human pulmonary veins of patients with atrial
fibrillation. J Cardiovasc Electrophysiol 14: 803– 809, 2003.
448. Peters NS, Severs NJ, Rothery S, Lincoln C, Yacoub MH,
Green CR. Spatiotemporal relation between gap junctions and
fascia adherens junctions during postnatal development of human
ventricular myocardium. Circulation 90: 713–725, 1994.
449. Petrov VVFRH, Lijnen PJ. Stimulation of collagen production by
transforming growth factor-beta1 during differentiation of cardiac
fibroblasts to myofibroblast. Hypertension 39: 258 –263, 2002.
450. Pham TD, Fenoglio JJ. Right atrial ultrastructural in chronic
rheumatic heart disease. Int J Cardiol 1: 289 –304, 1982.
451. Pizzetti FTF, Franzosi M. Incidence and prognostic significance
of atrial fibrillation in acute myocardial infarction The GISSI-3 data.
Heart 86: 527–532, 2001.
452. Platou ES, Refsum H. Class III antiarrhythmic action in experi-
mental atrial fibrillation and flutter in dogs. J Cardiovasc Pharma-
col 4: 839 – 846, 1982.
453. Po SS, Li Y, Tang D, Liu H, Geng N, Jackman WM, Scherlag B,
Lazzara R, Patterson E. Rapid and stable re-entry within the
pulmonary vein as a mechanism initiating paroxysmal atrial fibril-
lation. J Am Coll Cardiol 45: 1871–1877, 2005.
454. Polontchouk L, Haefliger JA, Ebelt B, Schaefer T, Stuhlmann
D, Mehlhorn U, Kuhn-Regnier F, De Vivie ER, Dhein S. Effects
of chronic atrial fibrillation on gap junction distribution in human
and rat atria. J Am Coll Cardiol 38: 883– 891, 2001.
455. Porter KETNA, O’Regan DJ, Balmforth AJ, Ball SG. Simvasta-
tin reduces human atrial myofibroblast proliferation independently
of cholesterol lowering via inhibition of RhoA. Cardiovasc Res 61:
745–755, 2004.
456. Postma AV, van de Meerakker JB, Mathijssen IB, Barnett P,
Christoffels VM, Ilgun A, Lam J, Wilde AA, Lekanne Deprez
RH, Moorman AF. A gain-of-function TBX5 mutation is associated
with atypical Holt-Oram syndrome and paroxysmal atrial fibrilla-
tion. Circ Res 102: 1433–1442, 2008.
457. Pratola C, Baldo E, Notarstefano P, Toselli T, Ferrari R.
Radiofrequency ablation of atrial fibrillation: is the persistence of
all intraprocedural targets necessary for long-term maintenance of
sinus rhythm? Circulation 117: 136 –143, 2008.
458. Pratola C, Baldo E, Notarstefano P, Toselli T, Ferrari R.
Radiofrequency atrial fibrillation ablation based on pathophysiol-
ogy: a diversified protocol with long-term follow-up. J Cardiovasc
Med 9: 68 –75, 2008.
459. Qi XYYY, Xiao L, Burstein B, Maguy A, Chartier D, Villeneuve
LR, Brundel BJ, Dobrev D, Nattel S. Cellular signaling underly-
ing atrial tachycardia remodeling of L-type calcium current. Circ
Res 103: 845– 854, 2008.
460. Raitt MH, Kusumoto W, Giraud G, McAnulty JH. Reversal of
electrical remodeling after cardioversion of persistent atrial fibril-
lation. J Cardiovasc Electrophysiol 15: 507–512, 2004.
461. Ramanathan C, Ghanem RN, Jia P, Ryu K, Rudy Y. Noninvasive
electrocardiographic imaging for cardiac electrophysiology and
arrhythmia. Nat Med 10: 422– 428, 2004.
462. Ramanathan C, Jia P, Ghanem R, Ryu K, Rudy Y. Activation and
repolarization of the normal human heart under complete physio-
logical conditions. Proc Natl Acad Sci USA 103: 6309 – 6314, 2006.
463. Ramani GZM, Good CB, Macioce A, Sonel AF. Comparison of
frequency of new-onset atrial fibrillation or flutter in patients on
statins versus not on statins presenting with suspected acute cor-
onary syndrome. Am J Cardiol 100: 404 – 405, 2007.
464. Ravelli F, Faes L, Sandrini L, Gaita F, Antolini R, Scaglione
M, Nollo G. Wave similarity mapping shows the spatiotemporal
distribution of fibrillatory wave complexity in the human right
atrium during paroxysmal and chronic atrial fibrillation. J Cardio-
vasc Electrophysiol 16: 1071–1076, 2005.
465. Reiffel JA. Drug choices in the treatment of atrial fibrillation. Am
J Cardiol 85: 12D–19D, 2000.
466. Rocken C, Peters B, Juenemann G, Saeger W, Klein H, Huth
C, Roessner A, Goette A. Atrial amyloidosis: an arrhythmogenic
substrate for persistent atrial fibrillation. Circulation 106: 2091–
2097, 2002.
467. Roithinger FX, Karch MR, Steiner PR, SippensGroenewegen
A, Lesh MD. Relationship between atrial fibrillation and typical
atrial flutter in humans: activation sequence changes during spon-
taneous conversion. Circulation 96: 3484 –3491, 1997.
468. Rosen MR, Danilo P Jr, Weiss RM. Actions of adenosine on
normal and abnormal impulse initiation in canine ventricle. Am J
Physiol Heart Circ Physiol 244: H715–H721, 1983.
469. Rostock T, Rotter M, Sanders P, Takahashi Y, Jais P, Hocini
M, Hsu LF, Sacher F, Clementy J, Haissaguerre M. High-
density activation mapping of fractionated electrograms in the atria
of patients with paroxysmal atrial fibrillation. Heart Rhythm 3:
27–34, 2006.
470. Roux JF, Gojraty S, Bala R, Liu CF, Hutchinson MD, Dixit S,
Callans DJ, Marchlinski F, Gerstenfeld EP. Complex fraction-
ated electrogram distribution and temporal stability in patients
undergoing atrial fibrillation ablation. J Cardiovasc Electrophysiol
19: 815– 820, 2008.
471. Roy D, Talajic M, Nattel S, Wyse DG, Dorian P, Lee KL,
Bourassa MG, Arnold JM, Buxton AE, Camm AJ, Connolly SJ,
Dubuc M, Ducharme A, Guerra PG, Hohnloser SH, Lambert J,
Le Heuzey JY, O’Hara G, Pedersen OD, Rouleau JL, Singh BN,
Stevenson LW, Stevenson WG, Thibault B, Waldo AL. Rhythm
control versus rate control for atrial fibrillation and heart failure. N
Engl J Med 358: 2667–2677, 2008.
472. Ryu K, Li L, Khrestian CM, Matsumoto N, Sahadevan J, Ruehr
ML, Van Wagoner DR, Efimov IR, Waldo AL. Effects of sterile
pericarditis on connexins 40 and 43 in the atria: correlation with
abnormal conduction and atrial arrhythmias. Am J Physiol Heart
Circ Physiol 293: H1231–H1241, 2007.
473. Saba S, Janczewski AM, Baker LC, Shusterman V, Gursoy EC,
Feldman AM, Salama G, McTiernan CF, London B. Atrial con-
tractile dysfunction, fibrosis, and arrhythmias in a mouse model of
cardiomyopathy secondary to cardiac-specific overexpression of
tumor necrosis factor-
.Am J Physiol Heart Circ Physiol 289:
H1456 –H1467, 2005.
474. Sahadevan J, Ryu K, Peltz L, Khrestian CM, Stewart RW,
Markowitz AH, Waldo AL. Epicardial mapping of chronic atrial
fibrillation in patients: preliminary observations. Circulation 110:
3293–3299, 2004.
475. Saito T, Tamura K, Uchida D, Saito T, Togashi M, Nitta T,
Sugisaki Y. Histopathological features of the resected left atrial
appendage as predictors of recurrence after surgery for atrial fi-
brillation in valvular heart disease. Circ J 71: 70 –78, 2007.
320 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
476. Saito T, Waki K, Becker AE. Left atrial myocardial extension onto
pulmonary veins in humans: anatomic observations relevant for atrial
arrhythmias. J Cardiovasc Electrophysiol 11: 888 – 894, 2000.
477. Sakabe M, Fujiki A, Nishida K, Sugao M, Nagasawa H,
Tsuneda T, Mizumaki K, Inoue H. Enalapril prevents perpetua-
tion of atrial fibrillation by suppressing atrial fibrosis and over-
expression of connexin43 in a canine model of atrial pacing-in-
duced left ventricular dysfunction. J Cardiovasc Pharmacol 43:
851– 859, 2004.
478. Sakabe M, Shiroshita-Takeshita A, Maguy A, Dumesnil C,
Nigam A, Leung TK, Nattel S. Omega-3 polyunsaturated fatty
acids prevent atrial fibrillation associated with heart failure but not
atrial tachycardia remodeling. Circulation 116: 2101–2109, 2007.
479. Salata JJ, Jurkiewicz NK, Jow B, Folander K, Guinosso PJ Jr,
Raynor B, Swanson R, Fermini B. I
K
of rabbit ventricle is com-
posed of two currents: evidence for I
Ks
.Am J Physiol Heart Circ
Physiol 271: H2477–H2489, 1996.
480. Salehian OHJ, Stambler B, Alnemer K, Almerri K, Grover J,
Bata I, Mann J, Matthew J, Pogue J, Yusuf S, Dagenais G,
Lonn E, Investigators HOPE. Impact of ramipril on the incidence
of atrial fibrillation: results of the Heart Outcomes Prevention
Evaluation study. Am Heart J 154: 448 – 453, 2007.
481. Sanchez JE, Plumb VJ, Epstein AE, Kay GN. Evidence for
longitudinal and transverse fiber conduction in human pulmonary
veins: relevance for catheter ablation. Circulation 108: 590 –597,
2003.
482. Sanders P, Berenfeld O, Hocini M, Jais P, Vaidyanathan R,
Hsu LF, Garrigue S, Takahashi Y, Rotter M, Sacher F, Scavee
C, Ploutz-Snyder R, Jalife J, Haissaguerre M. Spectral analysis
identifies sites of high-frequency activity maintaining atrial fibrilla-
tion in humans. Circulation 112: 789 –797, 2005.
483. Sanders P, Morton JB, Davidson NC, Spence SJ, Vohra JK,
Sparks PB, Kalman JM. Electrical remodeling of the atria in
congestive heart failure: electrophysiological and electroanatomic
mapping in humans. Circulation 108: 1461–1468, 2003.
484. Sanders P, Nalliah CJ, Dubois R, Takahashi Y, Hocini M,
Rotter M, Rostock T, Sacher F, Hsu LF, Jonsson A, O’Neill
MD, Jais P, Haissaguerre M. Frequency mapping of the pulmo-
nary veins in paroxysmal versus permanent atrial fibrillation. J
Cardiovasc Electrophysiol 17: 965–972, 2006.
485. Sanfilippo AJ, Abascal VM, Sheehan M, Oertel LB, Harrigan
P, Hughes RA, Weyman AE. Atrial enlargement as a consequence
of atrial fibrillation. A prospective echocardiographic study. Cir-
culation 82: 792–797, 1990.
486. Sarfstein RGY, Mizrahi A, Berdichevsky Y, Molshanski-Mor S,
Weinbaum C, Hirshberg M, Dagher MC, Pick E. Dual role of
Rac in the assembly of NADPH oxidase, tethering to the membrane
and activation of p67
phox
: a study based on mutagenesis of p67
phox
-
Rac1 chimeras. J Biol Chem 279: 16007–16016, 2004.
487. Sarmast F, Kolli A, Zaitsev A, Parisian K, Dhamoon AS, Guha
PK, Warren M, Anumonwo JM, Taffet SM, Berenfeld O, Jalife
J. Cholinergic atrial fibrillation: I(K,ACh) gradients determine un-
equal left/right atrial frequencies and rotor dynamics. Cardiovasc
Res 59: 863– 873, 2003.
488. Savelieva I, Camm J. Anti-arrhythmic drug therapy for atrial
fibrillation: current anti-arrhythmic drugs, investigational agents,
and innovative approaches. Europace 10: 647– 665, 2008.
489. Schauerte P, Scherlag BJ, Patterson E, Scherlag MA, Matsu-
daria K, Nakagawa H, Lazzara R, Jackman WM. Focal atrial
fibrillation: experimental evidence for a pathophysiologic role of
the autonomic nervous system. J Cardiovasc Electrophysiol 12:
592–599, 2001.
490. Schauerte P, Scherlag BJ, Pitha J, Scherlag MA, Reynolds D,
Lazzara R, Jackman WM. Catheter ablation of cardiac autonomic
nerves for prevention of vagal atrial fibrillation. Circulation 102:
2774 –2780, 2000.
491. Scherf D, Blumenfeld S, Gurbuzer B, Jody A. Effect of hyper-
capnia on aconitine-induced atrial fibrillation and on ventricular
fibrillation caused by focal cooling of the rapidly beating ventricles
of the dog. Am Heart J 55: 733–738, 1958.
492. Scherf D, Blumenfeld S, Taner D, Yildiz M. The effect of diphe-
nylhydantoin (Dilantin) sodium on atrial flutter and fibrillation
provoked by focal application of aconitine or delphinine. Am Heart
J60: 936 –947, 1960.
493. Scherlag BJ, Po S. The intrinsic cardiac nervous system and atrial
fibrillation. Curr Opin Cardiol 21: 51–54, 2006.
494. Scherlag BJ, Yamanashi W, Patel U, Lazzara R, Jackman WM.
Autonomically induced conversion of pulmonary vein focal firing
into atrial fibrillation. J Am Coll Cardiol 45: 1878 –1886, 2005.
495. Scherr D, Dalal D, Cheema A, Nazarian S, Almasry I, Bilchick
K, Cheng A, Henrikson CA, Spragg D, Marine JE, Berger RD,
Calkins H, Dong J. Long- and short-term temporal stability of
complex fractionated atrial electrograms in human left atrium dur-
ing atrial fibrillation. J Cardiovasc Electrophysiol 20: 13–21, 2009.
496. Schild LBA, Gardemann A, Polczyk P, Keilhoff G, Täger M.
Rapid pacing of embryoid bodies impairs mitochondrial ATP syn-
thesis by a calcium-dependent mechanism: a model of in vitro
differentiated cardiomyocytes to study molecular effects of tachy-
cardia. Biochim Biophys Acta 1762: 608 – 615, 2006.
497. Schimpf R, Giustetto C, Eckardt L, Veltmann C, Wolpert C,
Gaita F, Breithardt G, Borggrefe M. Prevalence of supraventric-
ular tachyarrhythmias in a cohort of 115 patients with Brugada
syndrome. Ann Noninvasive Electrocardiol 13: 266 –269, 2008.
498. Schlondorff JBC. Metalloprotease-disintegrins: modular proteins
capable of promoting cell-cell interactions and triggering signals by
protein-ectodomain shedding. J Cell Sci 112: 3603–3617, 1999.
499. Schmieder REKS, Julius S, McInnes GT, Zanchetti A, Hua TA,
Trial Group VALUE, Reduced incidence of new-onset atrial fibril-
lation with angiotensin II receptor blockade: the VALUE trial. J
Hypertens 26: 403– 411, 2008.
500. Schoonderwoerd BA, Ausma J, Crijns H, Van Veldhuisen DJ,
Blaauw EH, Van Gelder IC. Atrial ultrastructural changes during
experimental atrial tachycardia depend on high ventricular rate. J
Cardiovasc Electrophysiol 15: 1167–1174, 2004.
501. Schotten U, Ausma J, Stellbrink C, Sabatschus I, Vogel M,
Frechen D, Schoendube F, Hanrath P, Allessie MA. Cellular
mechanisms of depressed atrial contractility in patients with
chronic atrial fibrillation. Circulation 103: 691– 698, 2001.
502. Schotten U, de Haan S, Neuberger HR, Eijsbouts S, Blaauw Y,
Tieleman R, Allessie M. Loss of atrial contractility is primary
cause of atrial dilatation during first days of atrial fibrillation. Am J
Physiol Heart Circ Physiol 287: H2324 –H2331, 2004.
503. Schotten U, de Haan S, Verheule S, Harks EG, Frechen D,
Bodewig E, Greiser M, Ram R, Maessen J, Kelm M, Allessie M,
Van Wagoner DR. Blockade of atrial-specific K
-currents in-
creases atrial but not ventricular contractility by enhancing reverse
mode Na
/Ca
2
-exchange. Cardiovasc Res 73: 37– 47, 2007.
504. Schotten U, Duytschaever M, Ausma J, Eijsbouts S, Neu-
berger HR, Allessie M. Electrical and contractile remodeling
during the first days of atrial fibrillation go hand in hand. Circula-
tion 107: 1433–1439, 2003.
505. Schotten U, Greiser M, Benke D, Buerkel K, Ehrenteidt B,
Stellbrink C, Vazquez-Jimenez JF, Schoendube F, Hanrath P,
Allessie M. Atrial fibrillation-induced atrial contractile dysfunc-
tion: a tachycardiomyopathy of a different sort. Cardiovasc Res 53:
192–201, 2002.
506. Schotten U, Haase H, Frechen D, Greiser M, Stellbrink C,
Vazquez-Jimenez JF, Morano I, Allessie M, Hanrath P. The
L-type Ca
2
-channel subunits alpha1C and beta2 are not downregu-
lated in atrial myocardium of patients with chronic atrial fibrilla-
tion. J Mol Cell Cardiol 35: 437– 443, 2003.
507. Schreieck J, Wang Y, Overbeck M, Schomig A, Schmitt C.
Altered transient outward current in human atrial myocytes of
patients with reduced left ventricular function. J Cardiovasc Elec-
trophysiol 11: 180 –192, 2000.
508. Schrickel JW, Bielik H, Yang A, Schimpf R, Shlevkov N,
Burkhardt D, Meyer R, Grohe C, Fink K, Tiemann K, Luderitz
B, Lewalter T. Induction of atrial fibrillation in mice by rapid
transesophageal atrial pacing. Basic Res Cardiol 97: 452– 460, 2002.
509. Schuessler RB, Grayson TM, Bromberg BI, Cox JL, Boineau
JP. Cholinergically mediated tachyarrhythmias induced by a single
extrastimulus in the isolated canine right atrium. Circ Res 71:
1254 –1267, 1992.
510. Schuessler RB, Kawamoto T, Hand DE, Mitsuno M, Bromberg
BI, Cox JL, Boineau JP. Simultaneous epicardial and endocardial
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 321
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
activation sequence mapping in the isolated canine right atrium.
Circulation 88: 250 –263, 1993.
511. Schuessler RB, Kay MW, Melby SJ, Branham BH, Boineau JP,
Damiano RJ Jr. Spatial and temporal stability of the dominant
frequency of activation in human atrial fibrillation. J Electrocardiol
39: S7–12, 2006.
512. Shah DC, Haissaguerre M, Jais P, Clementy J. High-resolution
mapping of tachycardia originating from the superior vena cava:
evidence of electrical heterogeneity, slow conduction, and possible
circus movement reentry. J Cardiovasc Electrophysiol 13: 388 –
392, 2002.
513. Shan Z, van der Voort P, Blaauw Y, Duytschaever M, Allessie
M. Fractionation of electrograms and linking of activation during
pharmacologic cardioversion of persistent atrial fibrillation in the
goat. J Cardiovasc Electrophysiol 15: 572–580, 2004.
514. Shaw RM, Rudy Y. Ionic mechanisms of propagation in cardiac
tissue. Roles of the sodium and L-type calcium currents during
reduced excitability and decreased gap junction coupling. Circ Res
81: 727–741, 1997.
515. Sheehan KA, Zima AV, Blatter LA. Regional differences in spon-
taneous Ca
2
spark activity and regulation in cat atrial myocytes. J
Physiol 572: 799 – 809, 2006.
516. Shi H, Wang H, Li D, Nattel S, Wang Z. Differential alterations of
receptor densities of three muscarinic acetylcholine receptor sub-
types and current densities of the corresponding K
channels in
canine atria with atrial fibrillation induced by experimental con-
gestive heart failure. Cell Physiol Biochem 14: 31– 40, 2004.
517. Shi Y, Ducharme A, Li D, Gaspo R, Nattel S, Tardif JC. Re-
modeling of atrial dimensions and emptying function in canine
models of atrial fibrillation. Cardiovasc Res 52: 217–225, 2001.
518. Shi Y, Li D, Tardif JC, Nattel S. Enalapril effects on atrial
remodeling and atrial fibrillation in experimental congestive heart
failure. Cardiovasc Res 54: 456 – 461, 2002.
519. Shimano MTY, Inden Y, Kitamura K, Uchikawa T, Harata S,
Nattel S, Murohara T. Pioglitazone, a peroxisome proliferator-
activated receptor-gamma activator, attenuates atrial fibrosis and
atrial fibrillation promotion in rabbits with congestive heart failure.
Heart Rhythm 5: 451– 459, 2008.
520. Shimizu A, Nozaki A, Rudy Y, Waldo AL. Onset of induced atrial
flutter in the canine pericarditis model. J Am Coll Cardiol 17:
1223–1234, 1991.
521. Shinagawa K, Li D, Leung TK, Nattel S. Consequences of atrial
tachycardia-induced remodeling depend on the preexisting atrial
substrate. Circulation 105: 251–257, 2002.
522. Shinagawa K, Shi Y, Tardif JC, Leung TK, Nattel S. Dynamic
nature of atrial fibrillation substrate during development and re-
versal of heart failure in dogs. Circulation 105: 2672–2678, 2002.
523. Shiroshita-Takeshita A, Brundel BJ, Burstein B, Leung TK,
Mitamura H, Ogawa S, Nattel S. Effects of simvastatin on the
development of the atrial fibrillation substrate in dogs with con-
gestive heart failure. Cardiovasc Res 74: 75– 84, 2007.
524. Sinner MF, Pfeufer A, Akyol M, Beckmann BM, Hinterseer M,
Wacker A, Perz S, Sauter W, Illig T, Nabauer M, Schmitt C,
Wichmann HE, Schomig A, Steinbeck G, Meitinger T, Kaab S.
The non-synonymous coding IKr-channel variant KCNH2-K897T is
associated with atrial fibrillation: results from a systematic candi-
date gene-based analysis of KCNH2 (HERG). Eur Heart J 29:
907–914, 2008.
525. Skanes A, Mandapati R, Berenfeld O, Davidenko JM, Jalife J.
Spatiotemporal periodicity during atrial fibrillation in the isolated
sheep heart. Circulation 98: 1236 –1248, 1998.
526. Skasa M, Jungling E, Picht E, Schondube F, Luckhoff A. L-type
calcium currents in atrial myocytes from patients with persistent and
non-persistent atrial fibrillation. Basic Res Cardiol 96: 151–159, 2001.
527. Sosunov EA, Anyukhovsky EP, Rosen MR. Adrenergic-cholin-
ergic interaction that modulates repolarization in the atrium is
altered with aging. J Cardiovasc Electrophysiol 13: 374 –379, 2002.
528. Spach MS, Barr RC, Jewett PH. Spread of excitation from the
atrium into thoracic veins in human beings and dogs. Am J Cardiol
30: 844 – 854, 1972.
529. Spach MS, Boineau JP. Microfibrosis produces electrical load
variations due to loss of side-to-side cell connections. Pacing Clin
Electrophysiol 20: 397– 413, 1997.
530. Spach MS, Dolber PC. Relating extracellular potentials and their
derivatives to anisotropic propagation at a microscopic level in
human cardiac muscle. Evidence for electrical uncoupling of side-
to-side fiber connections with increasing age. Circ Res 58: 356 –371,
1986.
531. Spach MS, Heidlage JF. The stochastic nature of cardiac propa-
gation at a microscopic level. Electrical description of myocardial
architecture and its application to conduction. Circ Res 76: 366–
380, 1995.
532. Spach MS, Heidlage JF, Barr RC, Dolber PC. Cell size and
communication: role in structural and electrical development and
remodeling of the heart. Heart Rhythm 1: 500 –515, 2004.
533. Spach MS, Heidlage JF, Dolber PC, Barr RC. Electrophysiolog-
ical effects of remodeling cardiac gap junctions and cell size:
experimental and model studies of normal cardiac growth. Circ
Res 86: 302–311, 2000.
534. Spach MS, Heidlage JF, Dolber PC, Barr RC. Mechanism of
origin of conduction disturbances in aging human atrial bundles:
experimental and model study. Heart Rhythm 4: 175–185, 2007.
535. Spach MS, Josephson M. Initiating reentry: the role of nonuni-
form anisotropy in small circuits. J Cardiovasc Electrophysiol 5:
182–209, 1994.
536. Spach MS, Miller WT, Dolber PC, Kootsey JM, Sommer JR,
Mosher CE. The functional role of structural complexities in the
propagation of depolarization in the atrium of the dog. Cardiac
conduction disturbances due to discontinuities of effective axial
resistivity. Circ Res 50: 175–191, 1982.
537. Sparks PB, Mond HG, Vohra JK, Yapanis AG, Grigg LE, Kal-
man JM. Mechanical remodeling of the left atrium after loss of
atrioventricular synchrony. A long-term study in humans. Circula-
tion 100: 1714 –1721, 1999.
538. Spitzer SG, Richter P, Knaut M, Schuler S. Treatment of atrial
fibrillation in open heart surgery: the potential role of microwave
energy. Thorac Cardiovasc Surg 47 Suppl 3: 374 –378, 1999.
539. Stambler BS, Fenelon G, Shepard RK, Clemo HF, Guiraudon
CM. Characterization of sustained atrial tachycardia in dogs with
rapid ventricular pacing-induced heart failure. J Cardiovasc Elec-
trophysiol 14: 499 –507, 2003.
540. Steinberg JS, Zelenkofske S, Wong SC, Gelernt M, Sciacca R,
Menchavez E. Value of the P-wave signal-averaged ECG for pre-
dicting atrial fibrillation after cardiac surgery. Circulation 88:
2618 –2622, 1993.
541. Steiner I, Hajkova P. Patterns of isolated atrial amyloid: a study
of 100 hearts on autopsy. Cardiovasc Pathol 15: 287–290, 2006.
542. Stewart S, Hart CL, Hole DJ, McMurray JJ. A population-based
study of the long-term risks associated with atrial fibrillation: 20-
year follow-up of the Renfrew/Paisley study. Am J Med 113: 359 –
364, 2002.
543. Stiles MK, Brooks AG, Kuklik P, John B, Dimitri H, Lau DH,
Wilson L, Dhar S, Roberts-Thomson RL, Mackenzie L, Young
GD, Sanders P. High-density mapping of atrial fibrillation in hu-
mans. J Cardiovasc Electrophysiol 19: 1245–1253, 2008.
544. Sueda T, Nagata H, Shikata H, Orihashi K, Morita S, Sueshiro
M, Okada K, Matsuura Y. Simple left atrial procedure for chronic
atrial fibrillation associated with mitral valve disease. Ann Thorac
Surg 62: 1796 –1800, 1996.
545. Sugden PHCA. Regulation of the ERK subgroup of MAP kinase
cascade through G protein-coupled receptors. Cell Signal 9: 337–
351, 1997.
546. Sugiura H, Joyner RW. Action potential conduction between
guinea pig ventricular cells can be modulated by calcium current.
Am J Physiol Heart Circ Physiol 263: H1591–H1604, 1992.
547. Sun H, Chartier D, Leblanc N, Nattel S. Intracellular calcium
changes and tachycardia-induced contractile dysfunction in canine
atrial myocytes. Cardiovasc Res 49: 751–761, 2001.
548. Tai CT, Chen SA, Tzeng JW, Kuo BI, Ding YA, Chang MS,
Shyu LY. Prolonged fractionation of paced right atrial electro-
grams in patients with atrial flutter and fibrillation. J Am Coll
Cardiol 37: 1651–1657, 2001.
549. Taigen TDWLJ, Lim HW, Molkentin JD. Targeted inhibition of
calcineurin prevents agonist-induced cardiomyocyte hypertrophy.
Proc Natl Acad Sci USA 97: 1196 –1201, 2000.
322 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
550. Takahashi Y, O’Neill MD, Hocini M, Dubois R, Matsuo S,
Knecht S, Mahapatra S, Lim KT, Jais P, Jonsson A, Sacher F,
Sanders P, Rostock T, Bordachar P, Clementy J, Klein GJ,
Haissaguerre M. Characterization of electrograms associated
with termination of chronic atrial fibrillation by catheter ablation. J
Am Coll Cardiol 51: 1003–1010, 2008.
551. Takahashi Y, Sanders P, Jais P, Hocini M, Dubois R, Rotter M,
Rostock T, Nalliah CJ, Sacher F, Clementy J, Haissaguerre M.
Organization of frequency spectra of atrial fibrillation: relevance to
radiofrequency catheter ablation. J Cardiovasc Electrophysiol 17:
382–388, 2006.
552. Takeuchi S, Akita T, Takagishi Y, Watanabe E, Sasano C,
Honjo H, Kodama I. Disorganization of gap junction distribution
in dilated atria of patients with chronic atrial fibrillation. Circ J 70:
575–582, 2006.
553. Tan AY, Li H, Wachsmann-Hogiu S, Chen LS, Chen PS, Fish-
bein MC. Autonomic innervation and segmental muscular discon-
nections at the human pulmonary vein-atrial junction: implications
for catheter ablation of atrial-pulmonary vein junction. J Am Coll
Cardiol 48: 132–143, 2006.
554. Tan AY, Zhou S, Ogawa M, Song J, Chu M, Li H, Fishbein MC,
Lin SF, Chen LS, Chen PS. Neural mechanisms of paroxysmal
atrial fibrillation and paroxysmal atrial tachycardia in ambulatory
canines. Circulation 118: 916 –925, 2008.
555. Tanaka H, Hashimoto N. A multiple ion channel blocker, NIP-
142, for the treatment of atrial fibrillation. Cardiovasc Drug Rev 25:
342–356, 2007.
556. Tanaka K, Zlochiver S, Vikstrom KL, Yamazaki M, Moreno J,
Klos M, Zaitsev AV, Vaidyanathan R, Auerbach DS, Landas S,
Guiraudon G, Jalife J, Berenfeld O, Kalifa J. Spatial distribu-
tion of fibrosis governs fibrillation wave dynamics in the posterior
left atrium during heart failure. Circ Res 101: 839 – 847, 2007.
557. Tasaki H. Electrophysiological study of the striated muscle cells
of the extrapulmonary vein of the guinea-pig. Jpn Circ J 33: 1087–
1098, 1969.
558. 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.
559. Tessier S, Karczewski P, Krause EG, Pansard Y, Acar C,
Lang-Lazdunski M, Mercadier JJ, Hatem SN. Regulation of the
transient outward K
current by Ca
2
/calmodulin-dependent pro-
tein kinases II in human atrial myocytes. Circ Res 85: 810 – 819,
1999.
560. Tidball JG, Cederdahl JE, Bers DM. Quantitative analysis of
regional variability in the distribution of transverse tubules in
rabbit myocardium. Cell Tissue Res 264: 293–298, 1991.
561. Tieleman R, Van Gelder IC, Crijns H, De Kam PJ, Van Den
Berg MP, Haaksma J, Van Der Woude HJ, Allessie M. Early
recurrences of atrial fibrillation after electrical cardioversion: a
result of fibrillation-induced electrical remodeling of the atria? J
Am Coll Cardiol 31: 167–173, 1998.
562. Todd DM, Fynn SP, Walden AP, Hobbs WJ, Arya S, Garratt
CJ. Repetitive 4-week periods of atrial electrical remodeling pro-
mote stability of atrial fibrillation: time course of a second factor
involved in the self-perpetuation of atrial fibrillation. Circulation
109: 1434 –1439, 2004.
563. Tomita T, Takei M, Saikawa Y, Hanaoka T, Uchikawa S, Tsut-
sui H, Aruga M, Miyashita T, Yazaki Y, Imamura H, Kinoshita
O, Owa M, Kubo K. Role of autonomic tone in the initiation and
termination of paroxysmal atrial fibrillation in patients without
structural heart disease. J Cardiovasc Electrophysiol 14: 559 –564,
2003.
564. Tsai CTLL, Kuo KT, Hwang JJ, Hsieh CS, Hsu KL, Tseng CD,
Tseng YZ, Chiang FT, Lin JL. Angiotensin II activates signal
transducer and activators of transcription 3 via Rac1 in atrial
myocytes and fibroblasts: implication for the therapeutic effect of
statin in atrial structural remodeling. Circulation 117: 344 –355,
2008.
565. Tseng GN, Wit AL. Characteristics of a transient inward current
that causes delayed afterdepolarizations in atrial cells of the canine
coronary sinus. J Mol Cell Cardiol 19: 1105–1119, 1987.
566. Tseng GN, Wit AL. Effects of reducing [Na
]
o
on catecholamine-
induced delayed afterdepolarizations in atrial cells. Am J Physiol
Heart Circ Physiol 253: H115–H125, 1987.
567. Tuteja D, Xu D, Timofeyev V, Lu L, Sharma D, Zhang Z, Xu Y,
Nie L, Vazquez AE, Young JN, Glatter KA, Chiamvimonvat N.
Differential expression of small-conductance Ca
2
-activated K
channels SK1, SK2, and SK3 in mouse atrial and ventricular myo-
cytes. Am J Physiol Heart Circ Physiol 289: H2714 –H2723, 2005.
568. Ueng KC, Tsai TP, Yu WC, Tsai CF, Lin MC, Chan KC, Chen
CY, Wu DJ, Lin CS, Chen SA. Use of enalapril to facilitate sinus
rhythm maintenance after external cardioversion of long-standing
persistent atrial fibrillation. Results of a prospective and controlled
study. Eur Heart J 24: 2090 –2098, 2003.
569. Uno K, Kumagai K, Khrestian CM, Waldo AL. New insights
regarding the atrial flutter reentrant circuit: studies in the canine
sterile pericarditis model. Circulation 100: 1354 –1360, 1999.
570. Vaidya D, Morley GE, Samie FH, Jalife J. Reentry and fibrilla-
tion in the mouse heart. A challenge to the critical mass hypothesis.
Circ Res 85: 174 –181, 1999.
571. Van Capelle FJ, Durrer D. Computer simulation of arrhythmias
in a network of coupled excitable elements. Circ Res 47: 454 – 466,
1980.
572. Van der Velden HM, van Kempen MJ, Wijffels M, van Zijver-
den M, Groenewegen WA, Allessie M, Jongsma HJ. Altered
pattern of connexin40 distribution in persistent atrial fibrillation in
the goat. J Cardiovasc Electrophysiol 9: 596 – 607, 1998.
573. Van der Velden HMW, van der Zee L, Wijffels M, van Leuven
C, Dorland R, Vos MA, Jongsma HJ, Allessie M. Atrial fibrilla-
tion in the goat induces changes in monophasic action potential
and mRNA expression of ion channels involved in repolarization. J
Cardiovasc Electrophysiol 11: 1262–1269, 2000.
574. Van Gelder IC, Crijns HJ, Van Gilst WH, Hamer HP, Lie KI.
Decrease of right and left atrial sizes after direct-current electrical
cardioversion in chronic atrial fibrillation. Am J Cardiol 67: 93–95,
1991.
575. Van Veen TA, van Rijen HV, Jongsma HJ. Physiology of cardio-
vascular gap junctions. Adv Cardiol 42: 18 – 40, 2006.
576. Van Wagoner DR, Pond AL, Lamorgese M, Rossie SS, Mc-
Carthy PM, Nerbonne JM. Atrial L-type Ca
2
currents and human
atrial fibrillation. Circ Res 85: 428 – 436, 1999.
577. Van Wagoner DR, Pond AL, McCarthy PM, Trimmer JS, Ner-
bonne JM. Outward K
current densities and Kv1.5 expression are
reduced in chronic human atrial fibrillation. Circ Res 80: 772–781,
1997.
578. Vaquero M, Calvo D, Jalife J. Cardiac fibrillation: from ion
channels to rotors in the human heart. Heart Rhythm 5: 872– 879,
2008.
579. Vassalle M, Hoffman BF. The spread of sinus activation during
potassium administration. Circ Res 17: 285–295, 1965.
580. Venetucci LA, Trafford AW, Eisner DA. Increasing ryanodine
receptor open probability alone does not produce arrhythmogenic
calcium waves: threshold sarcoplasmic reticulum calcium content
is required. Circ Res 100: 105–111, 2007.
581. Verdecchia P, Reboldi G, Gattobigio R, Bentivoglio M, Bor-
gioni C, Angeli F, Carluccio E, Sardone MG, Porcellati C.
Atrial fibrillation in hypertension: predictors and outcome. Hyper-
tension 41: 218 –223, 2003.
582. Verheule S, Sato T, Everett Tt Engle SK, Otten D, Rubart-von
der Lohe M, Nakajima H, Nakajima H, Field LJ, Olgin J.
Increased vulnerability to atrial fibrillation in transgenic mice with
selective atrial fibrosis caused by overexpression of TGF-beta1.
Circ Res 94: 1458 –1465, 2004.
583. Verheule S, van Batenburg CA, Coenjaerts FE, Kirchhoff S,
Willecke K, Jongsma HJ. Cardiac conduction abnormalities in
mice lacking the gap junction protein connexin40. J Cardiovasc
Electrophysiol 10: 1380 –1389, 1999.
584. Verheule S, van Kempen MJ, te Welscher PH, Kwak BR,
Jongsma HJ. Characterization of gap junction channels in adult
rabbit atrial and ventricular myocardium. Circ Res 80: 673– 681,
1997.
585. Verheule S, Wilson E, Arora R, Engle SK, Scott L, Olgin J.
Tissue structure and connexin expression of canine pulmonary
veins. Cardiovasc Res 55: 727–738, 2002.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 323
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
586. Verheule S, Wilson E, Banthia S, Everett T, Shanbhag S, Sih
HJ, Olgin J. Direction-dependent conduction abnormalities in a
canine model of atrial fibrillation due to chronic atrial dilatation.
Am J Physiol Heart Circ Physiol 287: H634 –H644, 2004.
587. Verheule S, Wilson E, Everett T, Shanbhag S, Golden C, Olgin
J. Alterations in atrial electrophysiology and tissue structure in a
canine model of chronic atrial dilatation due to mitral regurgita-
tion. Circulation 107: 2615–2622, 2003.
587a.Verheule S, Tuyls E, van Hunnik A, Kuiper M, Schotten U,
Allessie MA. Fibrillatory conduction in the atrial free walls of
goats in persistent and permanent atrial fibrillation. Circ Arrhythm
Electrophysiol 2010 Oct 11. [Epub ahead of print]
588. Vermes ETJ, Bourassa MG, Racine N, Levesque S, White M,
Guerra PG, Ducharme A. Enalapril decreases the incidence of
atrial fibrillation in patients with left ventricular dysfunction: in-
sight from the Studies Of Left Ventricular Dysfunction (SOLVD)
trials. Circulation 107: 2926 –2931, 2003.
589. Vest JA, Wehrens XH, Reiken SR, Lehnart SE, Dobrev D,
Chandra P, Danilo P, Ravens U, Rosen MR, Marks AR. Defec-
tive cardiac ryanodine receptor regulation during atrial fibrillation.
Circulation 111: 2025–2032, 2005.
590. Virani SSNV, Razavi M, Lee VV, Elayda M, Wilson JM, Ballan-
tyne CM. Preoperative statin therapy is not associated with a
decrease in the incidence of postoperative atrial fibrillation in
patients undergoing cardiac surgery. Am Heart J 155: 541–546,
2008.
591. Vitebskiy SA, Khrestian CM, Waldo AL. Termination of a tachy-
arrhythmia by flunarizine is not a specific marker for a triggered
mechanism. Heart Rhythm 4: 1544 –1552, 2007.
592. Voigt N, Friedrich A, Bock M, Wettwer E, Christ T, Knaut M,
Strasser RH, Ravens U, Dobrev D. Differential phosphorylation-
dependent regulation of constitutively active and muscarinic recep-
tor-activated I
K,ACh
channels in patients with chronic atrial fibril-
lation. Cardiovasc Res 74: 426 – 437, 2007.
593. Volders PG, Kulcsar A, Vos MA, Sipido KR, Wellens HJ, Laz-
zara R, Szabo B. Similarities between early and delayed afterde-
polarizations induced by isoproterenol in canine ventricular myo-
cytes. Cardiovasc Res 34: 348 –359, 1997.
594. Wachtell KLM, Gerdts E, Olsen MH, Hornestam B, Dahlöf B,
Ibsen H, Julius S, Kjeldsen SE, Lindholm LH, Nieminen MS,
Devereux RB. Angiotensin II receptor blockade reduces new-
onset atrial fibrillation and subsequent stroke compared with
atenolol: the Losartan Intervention For End Point Reduction in
Hypertension (LIFE) study. J Am Coll Cardiol 45: 712–719, 2005.
595. Waggoner AD, Adyanthaya AV, Quinones MA, Alexander JK.
Left atrial enlargement. Echocardiographic assessment of electro-
cardiographic criteria. Circulation 54: 553–557, 1976.
596. Wagner ML, Lazzara R, Weiss RM, Hoffman BF. Specialized
conducting fibers in the interatrial band. Circ Res 18: 502–518, 1966.
596a.Wakili R, Yeh YH, Yan Qi X, Greiser M, Chartier D, Nishida K,
Maguy A, Villeneuve LR, Boknik P, Voigt N, Krysiak J, Kääb S,
Ravens U, Linke WA, Stienen GJ, Shi Y, Tardif JC, Schotten
U, Dobrev D, Nattel S. Multiple potential molecular contributors
to atrial hypocontractility caused by atrial tachycrdia remodeling in
dogs. Circ Arrhythm Electrophysiol 3: 530 –541, 2010.
597. Wakimoto H, Maguire CT, Kovoor P, Hammer PE, Gehrmann
J, Triedman JK, Berul CI. Induction of atrial tachycardia and
fibrillation in the mouse heart. Cardiovasc Res 50: 463– 473, 2001.
598. Waldo AI, Cooper TB. Spontaneous onset of type I atrial flutter in
patients. J Am Coll Cardiol 28: 707–712, 1996.
599. Waldo AL. Anticoagulation: stroke prevention in patients with
atrial fibrillation. Med Clin N Am 92: 143–159, 2008.
600. Waldo AL. From bedside to bench: entrainment and other stories.
Heart Rhythm 1: 94 –106, 2004.
601. Waldo AL, Feld GK. Inter-relationships of atrial fibrillation and
atrial flutter mechanisms and clinical implications. J Am Coll Car-
diol 51: 779 –786, 2008.
602. Wang TM, Chiang CE, Sheu JR, Tsou CH, Chang HM, Luk HN.
Homogenous distribution of fast response action potentials in ca-
nine pulmonary vein sleeves: a contradictory report. Int J Cardiol
89: 187–195, 2003.
603. Wang Y, Cuculich PS, Woodard PK, Lindsay BD, Rudy Y. Focal
atrial tachycardia after pulmonary vein isolation: noninvasive map-
ping with electrocardiographic imaging (ECGI). Heart Rhythm 4:
1081–1084, 2007.
604. Wang Y, Schuessler RB, Damiano RJ, Woodard PK, Rudy Y.
Noninvasive electrocardiographic imaging (ECGI) of scar-related
atypical atrial flutter. Heart Rhythm 4: 1565–1567, 2007.
605. Wang YG, Wagner MB, Kumar R, Goolsby WN, Joyner RW.
Fast pacing facilitates discontinuous action potential propagation
between rabbit atrial cells. Am J Physiol Heart Circ Physiol 279:
H2095–H2103, 2000.
606. Wang Z, Fermini B, Nattel S. Sustained depolarization-induced
outward current in human atrial myocytes. Evidence for a novel
delayed rectifier K
current similar to Kv1.5 cloned channel cur-
rents. Circ Res 73: 1061–1076, 1993.
607. Welikovitch L, Lafreniere G, Burggraf GW, Sanfilippo AJ.
Change in atrial volume following restoration of sinus rhythm in
patients with atrial fibrillation: a prospective echocardiographic
study. Can J Cardiol 10: 993–996, 1994.
608. Wettwer E, Hala O, Christ T, Heubach JF, Dobrev D, Knaut
M, Varro A, Ravens U. Role of I
Kur
in controlling action potential
shape and contractility in the human atrium: influence of chronic
atrial fibrillation. Circulation 110: 2299 –2306, 2004.
609. White CM, Giri S, Tsikouris JP, Dunn A, Felton K, Reddy P,
Kluger J. A comparison of two individual amiodarone regimens to
placebo in open heart surgery patients. Ann Thorac Surg 74: 69 –74,
2002.
610. White HD, Antman EM, Glynn MA, Collins JJ, Cohn LH,
Shemin RJ, Friedman PL. Efficacy and safety of timolol for
prevention of supraventricular tachyarrhythmias after coronary
artery bypass surgery. Circulation 70: 479 – 484, 1984.
611. Whitlock RP, Chan S, Devereaux PJ, Sun J, Rubens FD, Thor-
lund K, Teoh KH. Clinical benefit of steroid use in patients un-
dergoing cardiopulmonary bypass: a meta-analysis of randomized
trials. Eur Heart J 29: 2592–2600, 2008.
612. Wiegerinck RF, Verkerk AO, Belterman CN, van Veen TA,
Baartscheer A, Opthof T, Wilders R, de Bakker JM, Coronel
R. Larger cell size in rabbits with heart failure increases myocardial
conduction velocity and QRS duration. Circulation 113: 806 – 813,
2006.
613. Wiener NRA. The mathematical formation of the problem of conduc-
tion of impulses in a network of connected excitable elements, spe-
cifically in cardiac mucle.. Arch Inst Cardiol Mex 16: 205–265, 1946.
614. Wijffels M, Kirchhof C, Dorland R, Allessie M. Atrial fibrillation
begets atrial fibrillation. A study in awake chronically instrumented
goats. Circulation 92: 1954 –1968, 1995.
615. Wijffels M, Kirchhof C, Dorland R, Power J, Allessie M. Elec-
trical remodeling due to atrial fibrillation in chronically instru-
mented conscious goats: roles of neurohumoral changes, ischemia,
atrial stretch, and high rate of electrical activation. Circulation 96:
3710 –3720, 1997.
616. Wilders R, Jongsma HJ. Limitations of the dual voltage clamp
method in assaying conductance and kinetics of gap junction chan-
nels. Biophys J 63: 942–953, 1992.
617. Wilhelm M, Kirste W, Kuly S, Amann K, Neuhuber W, Weyand
M, Daniel WG, Garlichs C. Atrial distribution of connexin 40 and
43 in patients with intermittent, persistent, and postoperative atrial
fibrillation. Heart Lung Circulation 15: 30 –37, 2006.
618. Winfree AT. Electrical turbulence in three-dimensional heart mus-
cle. Science 266: 1003–1006, 1994.
619. Winfree AT. Spiral waves of chemical activity. Science 175: 634 –
636, 1972.
620. Wit AL, Boyden PA. Triggered activity and atrial fibrillation.
Heart Rhythm 4: S17–23, 2007.
621. Wit AL, Cranefield PF. Triggered and automatic activity in the
canine coronary sinus. Circ Res 41: 434 – 445, 1977.
622. Wit AL, Fenoglio JJ Jr, Hordof AJ, Reemtsma K. Ultrastruc-
ture and transmembrane potentials of cardiac muscle in the human
anterior mitral valve leaflet. Circulation 59: 1284 –1292, 1979.
623. Wit AL, Fenoglio JJ Jr, Wagner BM, Bassett AL. Electrophys-
iological properties of cardiac muscle in the anterior mitral valve
leaflet and the adjacent atrium in the dog Possible implications for
the genesis of atrial dysrhythmias. Circ Res 32: 731–745, 1973.
624. Wongcharoen W, Chen YC, Chen YJ, Chen SY, Yeh HI, Lin CI,
Chen SA. Aging increases pulmonary veins arrhythmogenesis and
324 SCHOTTEN ET AL.
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
susceptibility to calcium regulation agents. Heart Rhythm 4: 1338 –
1349, 2007.
625. Wongcharoen W, Chen YC, Chen YJ, Lin CI, Chen SA. Effects
of aging and ouabain on left atrial arrhythmogenicity. J Cardiovasc
Electrophysiol 18: 526 –531, 2007.
626. Workman AJ, Kane KA, Rankin AC. Cellular bases for human
atrial fibrillation. Heart Rhythm 5: S1– 6, 2008.
627. Workman AJ, Kane KA, Rankin AC. Characterisation of the Na,
K pump current in atrial cells from patients with and without
chronic atrial fibrillation. Cardiovasc Res 59: 593– 602, 2003.
628. Workman AJ, Kane KA, Rankin AC. The contribution of ionic
currents to changes in refractoriness of human atrial myocytes
associated with chronic atrial fibrillation. Cardiovasc Res 52: 226 –
235, 2001.
629. Workman AJ, Pau D, Redpath CJ, Marshall GE, Russell JA,
Norrie J, Kane KA, Rankin AC. Atrial cellular electrophysiolog-
ical changes in patients with ventricular dysfunction may predis-
pose to AF. Heart Rhythm 6: 445– 451, 2009.
630. Wright M, Haissaguerre M, Knecht S, Matsuo S, O’Neill MD,
Nault I, Lellouche N, Hocini M, Sacher F, Jais P. State of the
art: catheter ablation of atrial fibrillation. J Cardiovasc Electro-
physiol 19: 583–592, 2008.
631. Wu G, Huang CX, Tang YH, Jiang H, Wan J, Chen H, Xie Q,
Huang ZR. Changes of I
K,ATP
current density and allosteric mod-
ulation during chronic atrial fibrillation. Chin Med J 118: 1161–
1166, 2005.
632. Wu J, Estner H, Luik A, Ucer E, Reents T, Pflaumer A, Zren-
ner B, Hessling G, Deisenhofer I. Automatic 3D mapping of
complex fractionated atrial electrograms (CFAE) in patients with
paroxysmal and persistent atrial fibrillation. J Cardiovasc Electro-
physiol 19: 897–903, 2008.
633. Wu TJ, Doshi RN, Huang HL, Blanche C, Kass RM, Trento A,
Cheng W, Karagueuzian HS, Peter CT, Chen PS. Simultaneous
biatrial computerized mapping during permanent atrial fibrillation
in patients with organic heart disease. J Cardiovasc Electrophysiol
13: 571–577, 2002.
634. Xiao HD, Fuchs S, Campbell DJ, Lewis W, Dudley SC, Kasi VS,
Hoit BD, Keshelava G, Zhao H, Capecchi MR, Bernstein KE.
Mice with cardiac-restricted angiotensin-converting enzyme (ACE)
have atrial enlargement, cardiac arrhythmia, and sudden death. Am
J Pathol 165: 1019 –1032, 2004.
635. Xu J, Cui G, Esmailian F, Plunkett M, Marelli D, Ardehali A,
Odim J, Laks H, Sen L. Atrial extracellular matrix remodeling and
the maintenance of atrial fibrillation. Circulation 109: 363–368,
2004.
636. Xu Y, Tuteja D, Zhang Z, Xu D, Zhang Y, Rodriguez J, Nie L,
Tuxson HR, Young JN, Glatter KA, Vazquez AE, Yamoah EN,
Chiamvimonvat N. Molecular identification and functional roles
ofaCa
2
-activated K
channel in human and mouse hearts. J Biol
Chem 278: 49085– 49094, 2003.
637. Yagi T, Pu J, Chandra P, Hara M, Danilo P, Rosen MR, Boyden
PA. Density and function of inward currents in right atrial cells
from chronically fibrillating canine atria. Cardiovasc Res 54: 405–
415, 2002.
638. Yamamoto S, Suwa M, Ito T, Murakami S, Umeda T, Tokaji Y,
Sakai Y, Kitaura Y. Comparison of frequency of thromboembolic
events and echocardiographic findings in patients with chronic
nonvalvular atrial fibrillation and coarse versus fine electrocardio-
graphic fibrillatory waves. Am J Cardiol 96: 408 – 411, 2005.
639. Yamasaki Y, Furuya Y, Araki K, Matsuura K, Kobayashi M,
Ogata T. Ultra-high-resolution scanning electron microscopy of
the sarcoplasmic reticulum of the rat atrial myocardial cells. Anat
Rec 248: 70 –75, 1997.
639a.Yamazaki M, Vaquero LM, Hou L, Campbell K, Zlochiver S,
Klos M, Mironov S, Berenfeld O, Honjo H, Kodama I, Jalife J,
Kalifa J. Mechanisms of stretch-induced atrial fibrillation in the
presence and the absence of adrenocholinergic stimulation: inter-
play between rotors and focal discharges. Heart Rhythm 6: 1009 –
1017, 2009.
640. Yang SS, Han W, Zhou HY, Dong G, Wang BC, Huo H, Wei N,
Cao Y, Zhou G, Xiu CH, Li WM. Effects of spironolactone on
electrical and structural remodeling of atrium in congestive heart
failure dogs. Chin Med J 121: 38 – 42, 2008.
641. Yeh YH, Ehrlich JR, Qi X, Hebert TE, Chartier D, Nattel S.
Adrenergic control of a constitutively active acetylcholine-regu-
lated potassium current in canine atrial cardiomyocytes. Cardio-
vasc Res 74: 406 – 415, 2007.
642. Yeh YH, Qi X, Shiroshita-Takeshita A, Liu J, Maguy A,
Chartier D, Hebert T, Wang Z, Nattel S. Atrial tachycardia
induces remodelling of muscarinic receptors and their coupled
potassium currents in canine left atrial and pulmonary vein cardi-
omyocytes. Br J Pharmacol 152: 1021–1032, 2007.
643. Yeh YH, Wakili R, Qi XY, Chartier D, Boknik P, Kaab S,
Ravens U, Coutu P, Dobrev D, Nattel S. Calcium-handling ab-
normalities underlying atrial arrhythmogenesis and contractile dys-
function in dogs with congestive heart failure. Circ Arrhythm
Electrophysiol 1: 93–102, 2008.
644. Yoshihara FNT, Sasako Y, Hino J, Kobayashi J, Minatoya K,
Bando K, Kosakai Y, Horio T, Suga S, Kawano Y, Matsuoka H,
Yutani C, Matsuo H, Kitamura S, Ohe T, Kangawa K. Plasma
atrial natriuretic peptide concentration inversely correlates with
left atrial collagen volume fraction in patients with atrial fibrilla-
tion: plasma ANP as a possible biochemical marker to predict the
outcome of the maze procedure. J Am Coll Cardiol 39: 288 –294, 2002.
645. Young-Xu YJS, Goldberg R, Blatt CM, Graboys T, Bilchik B,
Ravid S. Usefulness of statin drugs in protecting against atrial
fibrillation in patients with coronary artery disease. Am J Cardiol
92: 1379 –1383, 2003.
646. Yu WC, Lee SH, Tai CT, Tsai CF, Hsieh MH, Chen CC, Ding
YA, Chang MS, Chen SA. Reversal of atrial electrical remodeling
following cardioversion of long-standing atrial fibrillation in man.
Cardiovasc Res 42: 470 – 476, 1999.
647. Yue L, Feng J, Gaspo R, Li GR, Wang Z, Nattel S. Ionic remod-
eling underlying action potential changes in a canine model of
atrial fibrillation. Circ Res 81: 512–525, 1997.
648. Yue L, Feng J, Wang Z, Nattel S. Adrenergic control of the
ultrarapid delayed rectifier current in canine atrial myocytes. J
Physiol 516: 385–398, 1999.
649. Yue L, Melnyk P, Gaspo R, Wang Z, Nattel S. Molecular mech-
anisms underlying ionic remodeling in a dog model of atrial fibril-
lation. Circ Res 84: 776 –784, 1999.
650. Zado E, Callans DJ, Riley M, Hutchinson M, Garcia F, Bala R,
Lin D, Cooper J, Verdino R, Russo AM, Dixit S, Gerstenfeld E,
Marchlinski FE. Long-term clinical efficacy and risk of catheter
ablation for atrial fibrillation in the elderly. J Cardiovasc Electro-
physiol 19: 621– 626, 2008.
651. Zellerhoff S, Pistulli R, Moennig G, Hinterseer M, Beckmann
BM, Koebe J, Steinbeck G, Kaab S, Haverkamp W, Fabritz L,
Gradaus R, Breithardt G, Schulze-Bahr E, Boecker D, Kirch-
hof P. Atrial arrhythmias in long QT syndrome: a nested case
control study. J Cardiovasc Electrophysiol 20: 401– 407, 2009.
652. Zhang H, Garratt CJ, Zhu J, Holden AV. Role of up-regulation of
I
K1
in action potential shortening associated with atrial fibrillation
in humans. Cardiovasc Res 66: 493–502, 2005.
653. Zhou S, Chang CM, Wu TJ, Miyauchi Y, Okuyama Y, Park AM,
Hamabe A, Omichi C, Hayashi H, Brodsky LA, Mandel WJ,
Ting CT, Fishbein MC, Karagueuzian HS, Chen PS. Nonreen-
trant focal activations in pulmonary veins in canine model of
sustained atrial fibrillation. Am J Physiol Heart Circ Physiol 283:
H1244 –H1252, 2002.
654. Zipes DP, Mihalick MJ, Robbins GT. Effects of selective vagal
and stellate ganglion stimulation of atrial refractoriness. Cardio-
vasc Res 8: 647– 655, 1974.
655. Zou R, Kneller J, Leon LJ, Nattel S. Substrate size as a deter-
minant of fibrillatory activity maintenance in a mathematical model
of canine atrium. Am J Physiol Heart Circ Physiol 289: H1002–
H1012, 2005.
656. Zou YKI, Yamazaki T. Cell type-specific angiotensin II-evoked
signal transduction pathways: critical roles of Gbetagamma sub-
unit, Src family, and Ras in cardiac fibroblasts. Circ Res 82: 337–
345, 1998.
PATHOPHYSIOLOGICAL MECHANISMS OF ATRIAL FIBRILLATION 325
Physiol Rev VOL 91 JANUARY 2011 www.prv.org
on January 30, 2011physrev.physiology.orgDownloaded from
doi:10.1152/physrev.00031.2009 91:265-325, 2011.Physiol Rev
Ulrich Schotten, Sander Verheule, Paulus Kirchhof and Andreas Goette
A Translational Appraisal
Pathophysiological Mechanisms of Atrial Fibrillation:
You might find this additional info useful...
654 articles, 388 of which can be accessed free at:This article cites http://physrev.physiology.org/content/91/1/265.full.html#ref-list-1
including high resolution figures, can be found at:Updated information and services http://physrev.physiology.org/content/91/1/265.full.html
can be found at:Physiological Reviewsabout Additional material and information
http://www.the-aps.org/publications/prv
This infomation is current as of January 30, 2011.
website at http://www.the-aps.org/.
MD 20814-3991. Copyright © 2011 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 January 30, 2011physrev.physiology.orgDownloaded from
... 41 The pathophysiology, cellular and molecular basis of AF has been extensively reviewed. 169,170 AF is thought to develop as a result of atrial, electrical and electromechanical coupling and atrial remodeling, tissue fibrosis, and changes in ionic homeostasis, particularly with calcium handling in cardiac tissue as well as gap-junction proteins. [169][170][171] Pre-existing hypertension is seen in 60-80% and diabetes in about 20% of patients with AF. 172 In addition to age, other risk factors correlate with the development of AF in patients with diabetes include female sex, Caucasian race, obesity, hyperuricemia, non-alcoholic fatty liver disease, heart rate recovery and heart failure. ...
... 169,170 AF is thought to develop as a result of atrial, electrical and electromechanical coupling and atrial remodeling, tissue fibrosis, and changes in ionic homeostasis, particularly with calcium handling in cardiac tissue as well as gap-junction proteins. [169][170][171] Pre-existing hypertension is seen in 60-80% and diabetes in about 20% of patients with AF. 172 In addition to age, other risk factors correlate with the development of AF in patients with diabetes include female sex, Caucasian race, obesity, hyperuricemia, non-alcoholic fatty liver disease, heart rate recovery and heart failure. 173 The current therapy for AF includes rhythm control with antiarrhythmic drugs, catheter ablation, cryoballoon ablation, left atrial appendage closure, the Maze surgical procedure and anticoagulation therapy. ...
... 235 Ozcan et al further demonstrated that treatment of cardiac specific LKB1-knockout mice with metformin (10 mg/ kg/day) activated AMPK, and significantly reduced the incidence of AF; as well as the expression of connexin proteins. 236 Interestingly, a report based on studies with diabetic rats has linked elevated AGEs to reduced phosphorylation of AMPK and dysregulation of Cx43 thereby promoting AF. 170 Yang et al 170 also reported that the AMPK agonist, 5-aminoimidazole-4-carboxamide ribonucleoside (AICAR), reversed the effects of AGEs to down-regulate Cx43 expression. ...
Article
Full-text available
Metformin is an orally effective anti-hyperglycemic drug that despite being introduced over 60 years ago is still utilized by an estimated 120 to 150 million people worldwide for the treatment of type 2 diabetes (T2D). Metformin is used off-label for the treatment of polycystic ovary syndrome (PCOS) and for pre-diabetes and weight loss. Metformin is a safe, inexpensive drug with side effects mostly limited to gastrointestinal issues. Prospective clinical data from the United Kingdom Prospective Diabetes Study (UKPDS), completed in 1998, demonstrated that metformin not only has excellent therapeutic efficacy as an anti-diabetes drug but also that good glycemic control reduced the risk of micro- and macro-vascular complications, especially in obese patients and thereby reduced the risk of diabetes-associated cardiovascular disease (CVD). Based on a long history of clinical use and an excellent safety record metformin has been investigated to be repurposed for numerous other diseases including as an anti-aging agent, Alzheimer’s disease and other dementias, cancer, COVID-19 and also atrial fibrillation (AF). AF is the most frequently diagnosed cardiac arrythmia and its prevalence is increasing globally as the population ages. The argument for repurposing metformin for AF is based on a combination of retrospective clinical data and in vivo and in vitro pre-clinical laboratory studies. In this review, we critically evaluate the evidence that metformin has cardioprotective actions and assess whether the clinical and pre-clinical evidence support the use of metformin to reduce the risk and treat AF.
... Atrial fibrillation (AF) is the most common type of cardiac arrhythmia in which electrical stimulation does not follow a specific path in the heart [1]. AF occurs when the electrical wave of stimulation in the atria does not have a particular direction; that is, the muscle cells of the atrium are irregularly stimulated and thus contract [2]. As a result, the atria cannot fully pump blood to the ventricles because there is no regular atrium contraction [3]. ...
Article
Full-text available
Introduction This study evaluates the cost‐effectiveness of Apixaban and Rivaroxaban, compared to Warfarin, for stroke prevention in patients with non‐valvular atrial fibrillation in Iran. Method A Markov model with a 30‐year time horizon was employed to simulate and assess different treatment strategies' cost‐effectiveness. The study population comprised Iranian adults with NVAF, identified through specialist consultations, hospital visits, and archival record reviews. Direct medical costs, direct nonmedical, and indirect costs were included. Quality‐adjusted life years (QALY) were assessed using an EQ‐5D questionnaire. This study utilized a cost‐effectiveness threshold of $11 134 per QALY. Results Apixaban demonstrated superior cost‐effectiveness compared to Rivaroxaban and Warfarin. Over 30 years, total costs were lower in the Apixaban and Rivaroxaban groups compared to the Warfarin group ($126.18 and $109.99 vs. $150.49). However, Apixaban showed higher total QALYs gained compared to others (0.134 vs. 0.133 and 0.116). The incremental cost‐effectiveness ratio for comparing Apixaban to Warfarin was calculated at −1332.83 cost per QALY, below the threshold of $11 134, indicating Apixaban's cost‐effectiveness. Sensitivity analyses confirmed the robustness of the findings, with ICER consistently remaining below the threshold. Over 5 years (2024−2028) of Apixaban usage, the incremental cost starts at USD 70 250 296 in the first year and gradually rises to USD 71 770 662 in the fifth year. DSA and PSA were assessed to prove the robustness of the results. Conclusion This study shows that Apixaban is a cost‐effective option for stroke prevention in non‐valvular atrial fibrillation patients in Iran compared to Warfarin.
... Moreover, SGLT2 inhibitors can inhibit sympathetic overdrive, which plays an important role in the development of AF (Herat et al., 2020). In addition, recent studies suggested that SGLT2 inhibitors had more favorable pleiotropic effects on HbA1c level, body weight, and systolic blood pressures when compared with DPP4 inhibitors, which may result in further arterial dilation and improved atrial remodeling and, thus, reduce the occurrence of AF (Schotten et al., 2011;Kim et al., 2019). ...
Article
Full-text available
Objectives To investigate the risk of atrial fibrillation (AF) with sodium-glucose cotransporter-2 inhibitors (SGLT2is) compared to dipeptidyl peptidase-4 inhibitor (DPP4i) use in older US adults and across diverse subgroups. Methods We conducted a retrospective cohort analysis using claims data from 15% random samples of Medicare fee-for-service beneficiaries. Patients were adults with type 2 diabetes (T2D), no preexisting AF, and were newly initiated on SGLT2i or DPP4i. The outcome was the first incident AF. Inverse probability treatment weighting (IPTW) was used to balance the baseline covariates between the treatment groups including sociodemographics, comorbidities, and co-medications. Cox regression models were used to assess the effect of SGLT2i compared to DPP4i on incident AF. Results Of the 97,436 eligible individuals (mean age 71.2 ± 9.8 years, 54.6% women), 1.01% (n = 983) had incident AF over a median follow-up of 361 days. The adjusted incidence rate was 8.39 (95% CI: 6.67–9.99) and 11.70 (95% CI: 10.9–12.55) per 1,000 person-years in the SGLT2i and DPP4i groups, respectively. SGLT2is were associated with a significantly lower risk of incident AF (HR 0.73; 95% CI, 0.57 to 0.91; p = 0.01) than DPP4is. The risk reduction of incident AF was significant in non-Hispanic White individuals and subgroups with existing atherosclerotic cardiovascular diseases and chronic kidney disease. Conclusion Compared to the use of DPP4i, that of SGLT2i was associated with a lower risk of AF in patients with T2D. Our findings contribute to the real-world evidence regarding the effectiveness of SGLT2i in preventing AF and support a tailored therapeutic approach to optimize treatment selection based on individual characteristics.
... Finally, enhancement of repolarizing atrial K + channels observed here is consistent with shortening of APD reported in porcine atrial myocytes upon exposure to dapagliflozin [11]. Thus, in contrast to the "classical" mechanism of atrial arrhythmogenesis in AF involving reduced electrical conduction velocity and shortening of atrial APD and effective refractory periods (AERP) that perpetuate AF through promotion of electrical re-entry [24], we suggest that in HF patients treated with dapagliflozin, pharmacologic activation of K 2P 2.1 and K 2P 17.1 current levels may specifically counteract atrial remodeling and suppress atrial arrhythmias. ...
Article
Full-text available
The sodium-glucose co-transporter-2 (SGLT2) inhibitor dapagliflozin is increasingly used in the treatment of diabetes and heart failure. Dapagliflozin has been associated with reduced incidence of atrial fibrillation (AF) in clinical trials. We hypothesized that the favorable antiarrhythmic outcome of dapagliflozin use may be caused in part by previously unrecognized effects on atrial repolarizing potassium (K+) channels. This study was designed to assess direct pharmacological effects of dapagliflozin on cloned ion channels Kv11.1, Kv1.5, Kv4.3, Kir2.1, K2P2.1, K2P3.1, and K2P17.1, contributing to IKur, Ito, IKr, IK1, and IK2P K+ currents. Human channels coded by KCNH2, KCNA5, KCND3, KCNJ2, KCNK2, KCNK3, and KCNK17 were heterologously expressed in Xenopus laevis oocytes, and currents were recorded using the voltage clamp technique. Dapagliflozin (100 µM) reduced Kv11.1 and Kv1.5 currents, whereas Kir2.1, K2P2.1, and K2P17.1 currents were enhanced. The drug did not significantly affect peak current amplitudes of Kv4.3 or K2P3.1 K+ channels. Biophysical characterization did not reveal significant effects of dapagliflozin on current–voltage relationships of study channels. In conclusion, dapagliflozin exhibits direct functional interactions with human atrial K+ channels underlying IKur, IKr, IK1, and IK2P currents. Substantial activation of K2P2.1 and K2P17.1 currents could contribute to the beneficial antiarrhythmic outcome associated with the drug. Indirect or chronic effects remain to be investigated in vivo.
... AF is a multifactorial disease that normally occurs in response to underlying cardiac abnormalities and is supported by changes in the electrophysiological, anatomical, and structural properties, generally referred to as atrial remodeling. 28 Recently, miRNAs have been confirmed to play an essential role in the pathophysiology of AF by regulating the 29 This study demonstrated that the low expression of miR-29b-3p in AF patients has certain predictive values in AF recurrence after RFCA, and is an independent risk factor for the recurrence of AF after RFCA. Mounting evidence indicates that miRNA expression level is correlated with the occurrence and severity of AF. 12 A previous study has shown that miR-29b is down-regulated in the plasma of AF patients. ...
Article
Full-text available
Objective Atrial fibrillation (AF) is a common arrhythmia. This study explored serum miR-29b-3p expression in AF patients and its value in predicting AF recurrence after radiofrequency catheter ablation (RFCA). Methods Totally 100 AF patients who underwent RFCA were enrolled, with 100 individuals without AF as controls. Serum miR-29b-3p expression in participants was determined using RT-qPCR. The correlation between miR-29b-3p and atrial fibrosis markers (FGF-21/FGF-23) was assessed by Pearson analysis. The diagnostic efficacy of serum miR-29b-3p and FGF-21/FGF-23 in predicting AF recurrence after RFCA was analyzed by the receiver operating characteristic (ROC) curves. The Kaplan-Meier method was adopted to evaluate the effect of miR-29b-3p expression on the incidence of AF recurrence after RFCA. The independent risk factors for AF recurrence after RFCA were analyzed by logistic regression analysis. Results Serum miR-29b-3p was poorly expressed in AF patients. After RFCA, AF patients showed elevated serum miR-29b-3p expression. Serum miR-29b-3p expression in AF patients negatively correlated with serum FGF-21 and FGF-23 concentrations. The cut-off values of serum miR-29b-3p, FGF-21, and FGF-23 in identifying AF recurrence were 0.860 (sensitivity: 100.00%, specificity: 39.71%), 222.2 pg/mL (sensitivity: 96.88%, specificity: 32.35%) and 216.3 ng/mL (sensitivity: 53.13%, specificity: 70.59%), respectively. Patients with low miR-29b-3p expression had a significantly higher incidence of AF recurrence than patients with high miR-29b-3p expression. Serum miR-29b-3p expression was one of the independent risk factors for AF recurrence after RFCA. Conclusion Low miR-29b-3p expression in AF patients has certain predictive values and is one of the independent risk factors for AF recurrence after RFCA.
Article
Full-text available
Despite the key role of fibrosis in atrial fibrillation (AF), the effects of different spatial distributions and textures of fibrosis on wave propagation mechanisms in AF are not fully understood. To clarify these aspects, we performed a systematic computational study to assess fibrosis effects on the characteristics and stability of re-entrant waves in electrically-remodelled atrial tissues. A stochastic algorithm, which generated fibrotic distributions with controlled overall amount, average size, and orientation of fibrosis elements, was implemented on a monolayer spheric atrial model. 245 simulations were run at changing fibrosis parameters. The emerging propagation patterns were quantified in terms of rate, regularity, and coupling by frequency-domain analysis of correspondent synthetic bipolar electrograms. At the increase of fibrosis amount, the rate of reentrant waves significantly decreased and higher levels of regularity and coupling were observed (p < 0.0001). Higher spatial variability and pattern stochasticity over repetitions was observed for larger amount of fibrosis, especially in the presence of patchy and compact fibrosis. Overall, propagation slowing and organization led to higher stability of re-entrant waves. These results strengthen the evidence that the amount and spatial distribution of fibrosis concur in dictating re-entry dynamics in remodeled tissue and represent key factors in AF maintenance.
Article
Atrial fibrillation (AF), the most common cardiac arrhythmia, is strongly associated with several comorbidities including heart failure (HF). AF in general, and specifically in the context of HF, is progressive in nature and associated with poor clinical outcomes. Current therapies for AF are limited in number and efficacy and do not target the underlying causes of atrial remodeling such as inflammation or fibrosis. We previously identified the calcium-activated SK4 K+ channels, which are preferentially expressed in the atria relative to the ventricles in both rat and human hearts, as novel druggable target for AF treatment. Here we examined the ability of BA6b9, a novel allosteric inhibitor of SK4 channels that targets the specific calmodulin-PIP2 binding domain, to alter AF susceptibility and atrial remodeling in a systolic HF rat post-myocardial infarction (post-MI) model. Daily BA6b9 injection (20 mg/kg/day) for 3 weeks starting 1-week post-MI prolonged the atrial effective refractory period, reduced AF induction and duration, and dramatically prevented atrial structural remodeling. In the post-MI left atrium (LA), pronounced upregulation of the SK4 K+ channel was observed, with corresponding increases in collagen deposition, α-SMA levels, and NLRP3 inflammasome expression. Strikingly, BA6b9 treatment reversed these changes while also significantly reducing the lateralization of the atrial connexin Cx43 in the LA of post-MI rats. Our findings indicate that the blockade of SK4 K+ channels using BA6b9 not only favors rhythm control but also remarkably reduces atrial structural remodeling, a property that is highly desirable for novel AF therapies, particularly in patients with comorbid HF.
Article
Full-text available
Atrial fibrillation is the commonest cardiac dysrhythmia. It is associated with significant morbidity and mortality. There are two approaches to the management of atrial fibrillation: controlling the ventricular rate or converting to sinus rhythm in the expectation that this would abolish its adverse effects. The objective of this review was to assess the effects of pharmacological cardioversion of atrial fibrillation in adults on the annual risk of stroke, peripheral embolism, and mortality. We made a thorough search for existing evidence in the following databases: the Cochrane Controlled Trials Register (Issue 3, 2002), MEDLINE (2000 to 2002), EMBASE (1998 to 2002), CINAHL (1982 to 2002), Web of Science (1981 to 2002). We also handsearched the following journals: Circulation (1997 to 2002), Heart (1997 to 2002), European Heart Journal (1997-2002), Journal of the American College of Cardiology (1997-2002) and selected abstracts published on the web site of the North American Society of Pacing and Electrophysiology (2001, 2002). We selected trials based on the following criteria: randomised controlled trials or controlled clinical trials of pharmacological cardioversion versus rate control in adults (>18 years) with acute, paroxysmal or sustained atrial fibrillation or atrial flutter, of any duration and of any aetiology. We identified two completed studies AFFIRM (n=4060) and PIAF (n=252). We found no difference in mortality between rhythm control and rate control - relative risk 1.14 (95% confidence interval 1.00 to 1.31). Both studies show significantly higher rates of hospitalisation and adverse events in the rhythm control group and no difference in quality of life between the two treatment groups. In AFFIRM there was a similar incidence of ischaemic stroke, bleeding and systemic embolism in the two groups. Certain malignant dysrhythmias were significantly more likely to occur in the rhythm control group. There were similar scores of cognitive assessment in both groups. In PIAF, cardioverted patients enjoyed an improved exercise tolerance but there was no overall benefit in terms of symptom control or quality of life. There is no evidence that pharmacological cardioversion of atrial fibrillation to sinus rhythm is superior to rate control. Rhythm control is associated with more adverse effects and increased hospitalisation. It does not reduce the risk of stroke. These conclusions cannot be generalised to all people with atrial fibrillation as most of the patients included in these studies were relatively older (>60 years) with significant cardiovascular risk factors.
Article
OBJECTIVES The purpose of the study was to characterize the ionic and molecular mechanisms in the very early phases of electrical remodeling in a rabbit model of rapid atrial pacing (RAP). BACKGROUND Long-term atrial fibrillation reduces L-type Ca2+ (I-Ca,I-L) and transient outward K+ (I-to) currents by transcriptional downregulation of the underlying ionic channels. However, electrical remodeling starts early after the onset of rapid atrial rates. The time course of ion current and channel modulation in these early phases of remodeling is currently unknown. METHODS Rapid (600 beats/min) right atrial pacing was performed in rabbits. Animals were divided into five groups with pacing durations between 0 and 96 h. Ionic currents were measured by patch clamp techniques; messenger ribonucleic acid (mRNA) and protein expression were measured by reverse transcription-polymerase chain reaction and Western blot, respectively. RESULTS L-type calcium current started to be reduced (by 47%) after 12 h of RAP and continued to decline as pacing continued. Current changes were preceded or paralleled by decreased mRNA expression of the Ca2+ channel beta subunits CaB2a, CaB2b, and CaB3, whereas significant reductions in the alpha(1) subunit mRNA and protein expression began 24 h after pacing onset. Transient outward potassium current densities were not altered within the first 12 h, but after 24 h, currents were reduced by 48%. Longer pacing periods did not further decrease I-to. Current changes were paralleled by reduced Kv4.3 mRNA expression. Kv4.2, Kv1.4, and the auxiliary subunit KChIP2 were not affected. CONCLUSIONS L-type calcium current and I-to are reduced in early phases of electrical remodeling. A major mechanism appears to be transcriptional downregulation of underlying ion channels which partially preceded ion current changes. (C) 2003 by the American College of Cardiology Foundation.
Chapter
Atrial fibrillation (AF) is such a common arrhythmia that it is often wrongly regarded as an acceptable alternative to normal sinus rhythm. Its first onset may present with rapid and uncomfortable palpitations, breathlessness, dyspnoea, chest pain, and anxiety. Often it is entirely asymptomatic and discovered quite by chance. Paroxysmal and persistent recurrences may eventually lapse into permanent AF....
Article
Introduction: Cell uncoupling due to fibrosis or increased extracellular collagen matrix (ECM) affects the formation of ectopic focal activity. Whether or not the increase of ECM also exists in the pulmonary veins (PVs) with rapid atrial pacing is unknown. We sought to test the hypothesis that in chronic atrial pacing dogs with sustained atrial fibrillation (AF), the amount of ECM was increased in both atria and the PVs. Methods and Results: We induced sustained AF in dogs by rapid atrial pacing. Computerized mapping techniques were used to map both atria and the PVs. We also used histological assessment to quantify the amount of ECM. After 118 +/- 24 days of rapid atrial pacing, sustained AF was induced in 7 dogs. Repetitive rapid activities (RRAs) either continuously or intermittently arose from the PVs during sustained AF. Histological study shows that there was no fibrosis in both atrial free walls and the PVs. However, the amount of ECM was increased in these regions. The mean ECM surface area fraction at each region in the dogs with sustained AF was all significantly higher compared to the corresponding region in normal dogs. Similarly, the heterogeneity of the ECM surface area fraction at each region in the dogs with sustained AF was also all significantly higher compared to normal dogs. Conclusions: In chronic atrial pacing-induced sustained AF, structural remodeling (i.e., inhomogeneous increase of ECM) also involves the PVs. Reduced coupling of the myocytes in the PV due to histological changes may provide an additional mechanism facilitating RRAs.
Article
BACKGROUND—Atrial fibrillation is the most common supraventricular arrhythmia in patients with acute myocardial infarction. Recent advances in pharmacological treatment of myocardial infarction may have changed the impact of this arrhythmia. OBJECTIVE—To assess the incidence and prognosis of atrial fibrillation complicating myocardial infarction in a large population of patients receiving optimal treatment, including angiotensin converting enzyme (ACE) inhibitors. METHODS—Data were derived from the GISSI-3 trial, which included 17 944 patients within the first 24 hours after acute myocardial infarction. Atrial fibrillation was recorded during the hospital stay, and follow up visits were planned at six weeks and six months. Survival of the patients at four years was assessed through census offices. RESULTS—The incidence of in-hospital atrial fibrillation or flutter was 7.8%. Atrial fibrillation was associated with indicators of a worse prognosis (age > 70 years, female sex, higher Killip class, previous myocardial infarction, treated hypertension, high systolic blood pressure at entry, insulin dependent diabetes, signs or symptoms of heart failure) and with some adverse clinical events (reinfarction, sustained ventricular tachycardia, ventricular fibrillation). After adjustment for other prognostic factors, atrial fibrillation remained an independent predictor of increased in-hospital mortality: 12.6% v 5%, adjusted relative risk (RR) 1.98, 95% confidence interval (CI) 1.67 to 2.34. Data on long term mortality (four years after acute myocardial infarction) confirmed the persistent negative influence of atrial fibrillation (RR 1.78, 95% CI 1.60 to 1.99). CONCLUSIONS—Atrial fibrillation is an indicator of worse prognosis after acute myocardial infarction, both in the short term and in the long term, even in an unselected population. Keywords: atrial fibrillation; acute myocardial infarction; prognosis