ﺑﺎﺯﮔﺸﺖ ﺑﻪ ﺻﻔﺤﻪ ﻗﺒﻠﯽ
خرید پکیج
تعداد آیتم قابل مشاهده باقیمانده : 3 مورد
نسخه الکترونیک
medimedia.ir

Mechanisms of atrial fibrillation

Mechanisms of atrial fibrillation
Literature review current through: Jan 2024.
This topic last updated: Sep 05, 2023.

INTRODUCTION — There are many mechanisms responsible for and contributory to development of atrial fibrillation (AF). The chance of developing AF is tied closely to age; AF is rare before age 50 [1]. In addition to age, there are many types of cardiac and medical conditions that are also closely linked to AF. These include hypertension, coronary artery disease, heart failure, valvular heart disease, obesity [2], and sleep-apnea syndrome. It is well established that high levels of alcohol [3] can increase the probability of developing AF, and that hyperthyroidism can cause AF. Evidence for caffeine and energy drinks, while suspected to be contributory, is questionable [4]. Furthermore, while exercise can be protective against atrial fibrillation, endurance athletics may be a cause for atrial fibrillation [5]. It is also well established that AF is more common in individuals who have a first-degree relative who developed AF at a young age. There is also a variety of acute conditions that can initiate AF such as cardiac surgery, pulmonary embolus, and inflammatory lung conditions. (See "Epidemiology, risk factors, and prevention of atrial fibrillation".)

The precise mechanisms by which age and the other conditions listed above increase the propensity for development of AF are understood poorly (figure 1). However, these conditions may impact the triggers for AF, which commonly arise in the pulmonary veins or the substrate for maintenance of AF, which broadly relates to atrial size and the extent of fibrosis. Some of the factors that may play a role in the mechanisms of AF include autonomic tone, inflammation, atrial pressure and wall stress, and genetics. (See "Epidemiology, risk factors, and prevention of atrial fibrillation", section on 'Other factors'.)

This topic will provide a broad overview of the current understanding of the mechanisms of AF. This discussion will provide a relatively simplistic approach to a complex topic. The reader will be referred to a rapidly growing literature on this topic, including some comprehensive reviews [6].

DEFINITIONS — The following terms are defined to help the reader’s understanding of the material below:

Trigger – Rapid firing of the left atrium from sites arising in the pulmonary veins often initiates atrial fibrillation (AF).

Triggered activity – One of three mechanisms of cardiac arrhythmias (including automaticity and reentry). Triggered activity refers to additional depolarizations, which occur during or immediately following a cardiac tissue electrical depolarization and may initiate a sustained cardiac arrhythmia.

Substrate – Mechanical and anatomic structure of the atria in which AF can occur and perpetuate.

Substrate remodeling – Changes in the mechanical and anatomic macro, micro, and ultrastructure of the atrial substrate that result from the development of AF and increase the propensity for the development and maintenance of AF over time.

Electrical remodeling – Changes in the atrial electrical properties (refractoriness and conduction) that result from the development of AF and increase the propensity for the development and maintenance of AF over time.

Dispersion of refractoriness – A range of differences in the refractory period properties throughout the atrial tissue.

Spatial heterogeneity of refractoriness – Dispersion of refractoriness manifest as variability in refractoriness throughout the atrial anatomy.

Complex fractionated electrograms – Local measurement of electrical activity recorded as bipolar electrograms obtained from areas of the atrium that are low amplitude with multiple components.

Reentry/reentrant mechanism – One of three mechanisms of cardiac arrhythmias (including automaticity and triggered activity). Reentry is the most common mechanism of cardiac arrhythmias and refers to the presence of one or more electrical circuit(s) in which electrical activation proceeds in a circular fashion to complete a self-sustaining circuit.

Atrial anisotropy – Altered conduction properties related to directionality of conduction through atrial tissue.

BASIC ATRIAL ELECTROPHYSIOLOGY — The electrophysiologic properties of normal and fibrillating atria have been studied extensively [7]. A basic understanding of these properties is necessary to understand the pathologic processes that play a role in initiating and perpetuating atrial fibrillation (AF). In the aggregate, these electrophysiologic properties permit the development of very complex patterns of conduction and an extremely rapid atrial rate as seen in AF.

The atrial myocardium consists of so-called "fast-response" tissues that depend on the rapidly activating sodium current for phase 0 of the action potential. (See "Cardiac excitability, mechanisms of arrhythmia, and action of antiarrhythmic drugs".)

Normal atrial myocardium has the following properties [7-9]:

A short action-potential duration.

Cellular reactivation can occur rapidly due to the short refractory period (in contrast to Purkinje fibers and ventricular muscle).

Very rapid electrical conduction can occur.

The refractory period shortens with increasing rate.

In the aggregate, these electrophysiologic properties permit the development of very complex patterns of conduction and an extremely rapid atrial rate as seen in AF.

CLINICAL FACTORS ASSOCIATED WITH AF — The following are common clinical conditions associated with atrial fibrillation (AF) in developed countries, and the percent of AF cases in which they are found (see "Epidemiology, risk factors, and prevention of atrial fibrillation", section on 'Chronic disease associations') [6]:

Hypertension (60 to 80 percent).

Cardiovascular disease, including cardiomyopathy, valvular and coronary artery disease (25 to 30 percent).

New York Heart Association class II to IV heart failure (30 percent).

Diabetes (20 percent).

Age.

Each of the first three is often associated with left atrial dilatation, which is important in the development of a substrate for AF and also may increase the probability of electrical firing from the pulmonary veins. (See 'Mechanisms of atrial fibrillation: triggers and substrates' below.)

The following section will discuss the link between these conditions and AF.

MECHANISMS OF ATRIAL FIBRILLATION: TRIGGERS AND SUBSTRATES — Atrial fibrillation (AF) may present as a paroxysmal (self-terminating AF within seven days), a persistent (one that lasts greater than seven days), or a long-standing persistent AF (continuous AF for 12 months or greater). The term “permanent AF” should be used when both the patient and physicians agree to not pursue strategies to restore or maintain sinus rhythm and AF continues unabated. (See "Atrial fibrillation: Overview and management of new-onset atrial fibrillation", section on 'Classification and terminology'.)

This wide range of clinical presentations is likely due to an interaction between a trigger and the substrate (figure 1). AF is initiated by rapid firing (or triggers) from the pulmonary veins (PV). Early in the course of AF the atrium is relatively healthy and as a result sinus rhythm is spontaneously restored. As the substrate remodels further over time, AF no longer terminates spontaneously and becomes persistent. With more extensive remodeling of the atrium, it becomes increasingly difficult to maintain sinus rhythm and the patient and physician may agree no longer to attempt to maintain sinus rhythm, with the AF thereby becoming permanent [10].

Triggers of AF — It has been known for many years that a single focus firing rapidly in the atria can be a trigger for fibrillatory conduction throughout the atria [11]. The most common site of the rapid atrial firing that triggers AF is the PVs. Catheter ablation of AF depends in large part on the electrical isolation of the PVs from the remainder of the atrium. Electrophysiologic evaluation of the PVs has identified myocardial tissue that can lead to repetitive firing or even the presence of episodic reentrant activation in the veins [6].

Additionally, stretch can increase the propensity for rapid firing from the PVs as a result of stretch sensitive ion channels. [12]. It has been speculated that the mechanism of atrial stretch may help explain the association between AF and mitral regurgitation as well as various types of heart failure that increase left atrial pressure.

Role of premature atrial complex and other arrhythmia triggers — AF is initiated (triggered) predominantly by rapid firing from PVs. Much less commonly, AF can be triggered by non-PV sites of rapid firing (such as tissue near the PV including the Vein of Marshall, the superior vena cava, or coronary sinus) or by other types of supraventricular arrhythmias including atrioventricular nodal reentrant tachycardia (AVNRT), orthodromic AV reciprocating tachycardia, and atrial flutter [6,13-23]. In some patients, successful elimination of AF with catheter ablation requires both isolation of the PVs, as well as elimination of these non-PV triggers. (See "Atrial fibrillation: Catheter ablation" and "Atrial fibrillation: Surgical ablation".)

Role of atrial flutter and supraventricular tachycardias — Atrial tachycardia, atrial flutter, and other supraventricular tachycardias can initiate AF in predisposed patients. The interaction between these arrhythmias and AF is not well understood, but atrial flutter and AF commonly coexist.

In some instances, elimination of atrial flutter will diminish and/or eliminate episodes of AF. Nevertheless, elimination of the right atrial reentry circuit responsible for typical flutter frequently does not eliminate the predisposition to AF that is predominately a left-atrial problem in a large number of patients. Many studies have demonstrated that patients who undergo catheter ablation of typical atrial flutter have a very high probability of developing AF over the ensuing five years. This is true regardless of whether AF had been observed prior to development of typical atrial flutter. This has clinical implications when it comes to ablation, but also has implications for anticoagulation strategies and patient follow-up. Nevertheless, for most patients, it makes sense to try to eliminate the organized supraventricular tachycardia, especially if right-sided by ablation before considering PV isolation and/or other more extensive ablation procedures to eliminate AF, as the AF may be reduced or eliminated by eliminating the other tachycardia first.

Role of the autonomic nervous system — The autonomic nervous system plays an important role in the development and maintenance of AF [24-26]. Clinical studies using heart rate variability analysis in patients with AF suggest that fluctuation in autonomic tone may be a major determinant of AF in patients with focal ectopy originating from the PVs [27]. Studies have also demonstrated a change in heart rate variability after PV ablation [28], further suggesting that PV triggers may be at least partially modulated by autonomic activity. Another study showed that the occurrence of paroxysmal AF greatly depends on variations of the autonomic tone, with a primary increase in adrenergic tone followed by an abrupt shift toward vagal predominance [29].

Anatomic studies of the autonomic innervation of the atria also indicate that the PVs and posterior left atrium (PLA) have a unique autonomic profile with a rich innervation from sympathetic and parasympathetic nerves [30-35]. The autonomic nervous system may also be playing a role in the genesis of AF in diseased hearts [30,36,37]. Studies suggest that the parasympathetic and sympathetic nervous system may also be playing a role in creation of AF substrate in the setting of heart failure [36,37].

Both the sympathetic and parasympathetic nervous systems have been implicated in the genesis [30,38,39] and maintenance of AF:

Sympathetic effects

Early studies suggested that exercise-induced AF may be sympathetically driven [30,40].

PV ectopic foci appear to be at least partially modulated by autonomic signaling, with sympathetic stimulation with isoproterenol frequently utilized to elicit these triggers in patients undergoing ablation for AF [41].

Parasympathetic (vagal) effects

The parasympathetic nervous system may contribute to AF in young patients with no structural heart disease [42].

Animal studies show that vagal stimulation contributes to the genesis of AF by nonuniform shortening of atrial effective refractory periods, thereby setting up substrate for reentry. Vagal stimulation can also lead to the emergence of focal triggers in the atrium [43-45].

Bezold-Jarisch-like "vagal" reflexes can be elicited during radiofrequency ablation and occur in and around the PVs. It has been suggested that elimination of these vagal reflexes during ablation may improve efficacy of AF ablation procedures [46].

Vagal responsiveness also appears to decrease following ablation in the left atrium [47]. In some series, adding ganglionated plexi (GP) ablation to PV isolation appears to increase ablation success for AF [12]. Data suggest that areas in the atrium demonstrating complex fractionated atrial electrograms (CFAE) may represent a suitable target site for ablation; although several studies have reported that ablation at these sites may increase the efficacy of PV isolation procedures [48,49], enthusiasm for this approach has fallen over time. One possible explanation for the improvement in ablation success reported in these trials is that several CFAE sites anatomically overlie fat pads containing GPs [18,50]. As indicated above, autonomic denervation performed by GP ablation is thought to improve efficacy of AF ablation. (See "Atrial fibrillation: Catheter ablation" and "Catheter ablation for the treatment of atrial fibrillation: Technical considerations for non-electrophysiologists", section on 'Ablation techniques and targets'.)

In a study of 40 patients with paroxysmal AF scheduled to undergo catheter ablation, individuals were randomly assigned to noninvasive transcutaneous low-level stimulation of the tragus (the anterior protuberance of the ear where the auricular branch of the vagus nerve is accessible) or to sham stimulation for one hour. Compared with control, low-level stimulation suppressed AF as measured by the decreased duration of atrial pacing-induced AF and an increased AF cycle length [51].

Maintenance of atrial fibrillation — In patients with persistent AF, the prevailing understanding of the mechanism is that, once triggered, the arrhythmia is maintained (sustained) by one or more abnormalities in the atrial tissue. This process may explain why the failure rate of PV isolation is as high as 40 to 60 percent at one year: The trigger(s) may have been treated but not the abnormalities that sustain AF once triggered (initiated).

The role of localized sources (electrical rotors and focal impulses) in the initiation and maintenance of AF was explored in the CONFIRM trial of 92 patients undergoing ablation procedures for paroxysmal or persistent (72 percent) AF [52]. Consecutive patients were prospectively treated (not randomly assigned) in a 1:2 case-cohort design with either conventional ablation at sources identified within the atria followed by conventional ablation or conventional ablation alone. Localized sources were identified in 97 percent of cases (70 percent rotors and 30 percent focal impulses) with sustained AF, each with an average of 2.1 sources. During a median of 273 days, patients treated with treatment of both sources and conventional ablation had a significantly higher freedom from AF (82.4 versus 44.9 percent).

Similar information was reported, indicating that “driver domains,” located in specific areas of the atria, act as unstable re-entry circuits that perpetuate atrial fibrillation in patients who have persistent AF [53,54].

Murine cell cultures show a differential ion channel gene expression associated with atrial tissue remodeling (ie, decreased SCN5A, CACN1C, KCND3, and GJA1; and increased KCNJ2) [55]. Fibrillatory complexity, increased in late compared with early stage cultures, was associated with a decrease in rotor tip meandering and increase in wavefront curvature.

Rotors are not the only explanation. In a study using high-density, simultaneous, biatrial, epicardial mapping of persistent and longstanding persistent AF in patients undergoing open heart surgery, several non-reentrant drivers were present in both atria in 11 or 12 patients with two to four foci per patient; foci were seen in both atria but generally in the lateral left atrial free wall, and likely acted as drivers. Reentry was not found to be the mechanism [56].

Likely, the substrate to maintain AF is a combination of reentrant activity and focal triggers. In a study of biatrial epicardial mapping of AF in sheep, wave propagation patterns were passing wave (69 percent occurrence, 68.6 percent of total time), point source (20.4, 13.1 percent), wave collision (4, 2.8 percent), reentrant wave (0.7, 6.3 percent), half-rotation (2.9, 4.4 percent), wave splitting (2.7, 4.3 percent), conduction block (0.05, 0.03 percent) and figure of eight reentry (0.05, 0.05 percent) [57]. Periods of repetitive activity were detected in the left and right atria.

The following sections describe factors that might contribute to the maintenance of AF.

Atrial remodeling — Atrial remodeling involves the concept that there are structural changes, such as fibrosis, or electrical changes, such as refractory-period dispersion or conduction delay, in the atria that can predispose to the development and maintenance of AF. In some instances, structural and electrical changes occur simultaneously. These processes can facilitate or create electrical reentrant circuits or triggers that can lead to AF [13,58]. It is also well established that the presence of AF results in remodeling of the atrium over time [7]. This explains the well-established concept that AF begets AF (figure 2). Thus, the longer a patient has been in continuous AF, the less likely it is to terminate spontaneously, and harder it is to restore and maintain sinus rhythm [59].

Electrical remodeling — Paroxysmal AF commonly precedes persistent AF. It has been suggested even after only a few minutes, AF induces transient changes in atrial electrophysiology that promote its perpetuation [14]. This might occur through a tachycardia-induced cardiomyopathy or through "electrical remodeling" of the atria by AF, leading to a progressive decrease in atrial refractoriness [14,15]. Electrical remodeling results from the high rate of electrical activation, which stimulates the AF-induced changes in refractoriness [60]. Tachycardia-induced changes in refractoriness are spatially nonuniform and there is increased variability both within and among various atrial regions [61]. It is possible that the change in atrial refractory period observed after an episode of AF predisposes to the spontaneous recurrence of AF in the days following cardioversion.

In addition to the shortening of the refractory period, chronic, rapid, atrial pacing-induced AF results in other changes within the atria, including an increase in the expression and distribution of connexin 43 and heterogeneity in the distribution of connexin 40, both of which are intercellular gap junction proteins ("gap junctional remodeling") [16,17]; cellular remodeling is due to apoptotic death of myocytes with myolysis, which may not be entirely reversible [18]; the induction of sinus node dysfunction, demonstrated by prolonged corrected sinus node recovery time, reduced maximal heart rate in response to isoproterenol, and lower intrinsic heart rate after administration of atropine and propranolol [19]; and an increase in P wave duration and intraatrial conduction time.

A clinical study evaluated the hypothesis of electrical remodeling by the use of atrial pacing-induced AF in patients with a history of supraventricular tachycardia [20]. AF significantly shortened the right-atrial effective refractory period after only a few minutes, and temporal recovery of the refractory period occurred over about eight minutes. Upon termination of AF, there was an increased propensity for the induction of another episode of AF that decreased with increasing time after the initial AF reversion. The second also tends to last longer than the first.

The time to recurrence was also evaluated in a review of 61 patients who had daily electrocardiogram (ECG) recordings using transtelephonic monitoring: 57 percent had recurrent AF during the first month after cardioversion, with a peak incidence during the first five days [21]. Among patients with recurrence, there was a positive correlation between the duration of the shortest coupling interval of premature atrial complex (PAC; also referred to a premature atrial beat, premature supraventricular complex, or premature supraventricular beat) after cardioversion, which correlates with the refractory period and the timing of recurrence (figure 3). (See "Atrial fibrillation: Cardioversion".)

In contrast to the normal situation in which the atrial refractory period shortens with an increase in rate (as in AF) and prolongs when the rate decreases, the refractory period fails to lengthen appropriately at slow rates (eg, with return to sinus rhythm) in patients with acute or persistent AF. The duration of AF has no significant impact upon the extent of these electrophysiologic changes [22].

Atrial electrical remodeling is reversed gradually after the restoration of sinus rhythm [23,62]. This may be one of the explanations for the early or immediate return of AF after cardioversion. In one study of 25 patients, the atrial refractory period increased and the adaptation of atrial refractoriness to rate was normal by four weeks after cardioversion [23]. In another report of 38 patients, the atrial refractory period increased by one week, with some variation in different regions of the atrium [62]. This observation has important clinical implications.

The mechanism for electrical remodeling and shortening of the atrial refractory period is not entirely clear; a possible explanation is ion-channel remodeling, with a decrease in the protein content of the L-type calcium channel [63]. Support for this comes from an animal study in which verapamil, an L-type calcium antagonist, prevented electric remodeling of short-duration AF (one day or less) and hastened complete recovery, without affecting inducibility of AF [64]. Similar findings have been noted in humans as verapamil, but not procainamide, prevented remodeling when given prior to the electrophysiologic induction of AF [65]. Oral diltiazem is also effective in some patients [66], while beta blockers had no effect on electrical remodeling in an animal model [60].

In comparison, cytosolic calcium overload, induced by hypercalcemia or digoxin, which increases the intracellular concentration of calcium by activating the sodium-calcium exchanger, enhances electrical remodeling [64,67,68]. The effect of digoxin, which is not due to its vagotonic activity, is associated with an increase in the inducibility and duration of AF [68].

Calcium leak from the sarcoplasmic reticulum may trigger and maintain AF. It is known that protein kinase A (PKA) hyperphosphorylation of the cardiac ryanodine receptor (RyR2), resulting in dissociation of the channel-stabilizing subunit calstabin2, causes sarcoplasmic reticulum (SR) calcium leak in failing hearts. This phenomenon seems to be involved in triggering ventricular arrhythmias.

Using similar logic, these proteins were investigated in atrial tissues from both dogs and humans with AF [69]. Atrial tissue in those with AF showed a significant increase in PKA phosphorylation of RyR2 and a decrease in calstabin2 binding to the channel. Channels isolated from dogs with AF had an increased open probability under conditions simulating diastole compared with channels from control hearts, suggesting that these AF channels could predispose to a diastolic SR calcium leak. The conclusion was that SR calcium leak due to RyR2 PKA hyperphosphorylation may play a role in the initiation and/or maintenance of AF. Other studies also suggest that RyR2 receptor-mediated calcium leak drives progressive development of an atrial fibrillation substrate in a transgenic mouse model [70,71].

The effects of calcium overload are quite complex. It is likely that triggers and substrates initiate short episodes of AF that then lead to calcium overload and over a period of minutes there is activation of the I CaL current that increases I K1, decreases I Na, increases IKACh, and decreases ITO. This can affect the action-potential duration and allow for more reentry to occur. As reentry occurs, the substrate changes and there is remodeling through calcium handling abnormalities as well as mRNA transcription [59], and ultimately perhaps with protein decrease, changes in connexons, including, Cx40, that can affect conduction. The calcium-handling abnormalities can also lead to hypocontractility and atrial dilatation, thereby affecting even more the possibility of developing AF [59].

Both animal and human studies suggest that angiotensin II is involved in electrical and atrial myocardial remodeling [72,73] (see "Pathophysiology of heart failure: Neurohumoral adaptations", section on 'Renin-angiotensin system'). In an animal model, inhibition of angiotensin II with captopril or candesartan prevented shortening of the atrial effective refractory period and atrial electrical remodeling during rapid atrial pacing [72], while atrial tissue obtained during open heart surgery from patients with AF revealed downregulation of AT1 receptor proteins and upregulation of AT2 receptor [73]. The potential clinical importance of these changes is illustrated by the observations that angiotensin converting enzyme (ACE) inhibitors reduce the incidence of AF in patients with left ventricular dysfunction after myocardial infarction [74] and in patients with chronic left ventricular dysfunction due to ischemic heart disease [75]. (See "Angiotensin converting enzyme inhibitors and receptor blockers in acute myocardial infarction: Clinical trials".)

Another possible contributor to electrical remodeling and shortening of the atrial refractory period is atrial ischemia, which activates the sodium/hydrogen exchanger. The intravenous administration of HOE 642, a selective inhibitor of this sodium proton pump, to dogs undergoing rapid atrial pacing resulted in the lengthening of atrial refractoriness after one hour, while control dogs showed effective refractory-period shortening greater than 10 percent [76].

Role of fibrosis — The development of AF invokes atrial remodeling processes that involve electrophysiological and structural alterations that serve to maintain, promote, and propagate AF. In addition to electrophysiological alterations, such as shortening of the atrial action potential, increased dispersion of refractoriness, and conduction velocity shortening, morphological changes consist of fibrosis, hypertrophy, necrotic and apoptotic cell loss, and dilation [77].

Of these, fibrosis is considered especially important in the creation of AF substrate, especially in the setting of chronic atrial dilatation caused by heart failure. A canine model of heart failure has demonstrated a progressive increase in AF inducibility with increasing fibrosis [78]. An increase in conduction heterogeneity noted in this model is thought to play a major role in the creation of reentrant circuits in the dilated atria. Patients with AF also display increased atrial fibrous tissue content, along with increased expression of collagen I and III [79], as well as up-regulation of MMP-2 protein, and down-regulation of the tissue inhibitor of metalloproteinase, TIMP-1 [79]. Expression of the active form of MMP-9 and of monocyte chemoattractant protein-1, an inflammatory mediator, is increased in AF patients [80]. The left atrial free wall around the PV area presents particularly strong interstitial fibrotic changes [81-83].

Although the underlying molecular mechanisms that lead to the development of atrial fibrosis are complex, work suggests that the TGF-beta pathway may be an important contributor to the development of fibrosis (especially in the setting of increasing atrial stretch/dilatation resulting from congestive heart failure) [84-86].

Role of inflammation and oxidative stress — Emerging evidence suggests a significant role of inflammation in the pathogenesis of AF [87]. Evidence includes elevated serum levels of inflammatory biomarkers in patients with AF, the expression of inflammatory markers in atrial tissue from AF patients, and beneficial effects of antiinflammatory drugs in the setting of experimental AF [88]. Inflammation is suggested to be linked to various pathological processes, such as oxidative stress, apoptosis, and fibrosis that promote the creation and perpetuation of AF substrate. Several of the downstream effects of inflammation in the heart are thought to be mediated by oxidative stress [89].

Indeed, studies in patients with AF demonstrate increased generation of reactive oxygen species (ROS) in the fibrillating atrium compared with normal atria [90,91]. Several major enzymatic sources of ROS have been implicated in AF. Of these, NAPDH oxidase (specifically its NOX2 isoform) has been shown to be elevated in humans with AF in a variety of studies [92,93]. Other sources of ROS implicated in AF include uncoupled nitric oxide synthase [94] and xanthine oxidase [95]. In addition to the increase in ROS noted in tissue from patients with AF, experimental evidence suggests that ROS may be implicated not only in promoting AF but also in maintaining atrial arrhythmia. The administration of antioxidants such as vitamin C or statins (which are known to have pleiotropic antioxidant effects) decreased AF inducibility in canine models of tachypacing-induced AF [96,97]. Antioxidants such as vitamin C and n-acetylcysteine have been administered to patients undergoing cardiac surgery and have been shown to decrease postoperative AF [98,99]. These early results are encouraging and warrant further investigation of inflammation and oxidative stress as viable therapeutic targets in patients with AF.

Recent data indicate an association of gut microbiome dysbiosis and AF [100,101]. In a recent meta-analysis of 14 studies on 2479 patients, biomarkers of gut dysbiosis (eg, short-chain fatty acids, trimethylamine N-oxide, lipopolysaccharides, and bile acids) were shown to be related to AF. Whether inflammation plays a role in this association is uncertain. The microbiome may also be related to postoperative AF [102]; proposed mediators of this association include increased postoperative inflammation and electrolyte abnormalities.

Reentrant mechanism — Maps of AF in animals and humans suggest that this arrhythmia is caused by multiple wandering wavelets (figure 4), and these may be due to heterogeneity of atrial refractoriness and conduction. In addition, the response of atrial activity to adenosine infusion suggests a reentrant rather than a focal mechanism [103]. Adenosine increases the inward potassium rectifier current, which shortens refractory periods and would accelerate reentrant circuits. In contrast, this effect would slow an automatic or triggered focus. In a series of 33 patients with AF undergoing electrophysiology study, adenosine increased the dominant frequencies, supporting reentrant rather than focal sources for the perpetuation of AF.

It has been suggested that at least four to six independent wavelets are required to maintain AF [104]. These wavelets rarely reenter themselves but can re-excite portions of the myocardium recently activated by another wavefront, a process called random reentry [7,105-107]. As a result, there are multiple wavefronts of activation that may collide with each other, extinguishing themselves or creating new wavelets and wavefronts, thereby perpetuating the arrhythmia (figure 5).

The reentrant circuits are therefore unstable; some disappear, while others reform. These circuits have variable but short cycle lengths, resulting in multiple circuits to which atrial tissue cannot respond in a 1:1 fashion. As a result, functional block, slow conduction, and multiple wave fronts develop [107].

Patients with AF may have increased dispersion of refractoriness. This correlates with enhanced inducibility of AF and spontaneous episodes [108] likely related to unstable reentry circuits. Some patients have site-specific dispersion of atrial refractoriness and intraatrial conduction delays resulting from nonuniform atrial anisotropy [109]. This appears to be a common property of normal atrial tissue, but there are further conduction delays to and within area surrounding the AV node in patients with induced AF, suggesting an important role for the low right atrium in the genesis of AF.

Abnormalities in restitution as well as the spatial distribution of such abnormalities can be related to the persistence of AF. In one study, monophasic action potential recordings were evaluated in patients with AF [110]. The action potential duration was plotted as a function of the preceding diastolic interval, and the slope of the action potential duration versus the diastolic interval (the restitution curve) was determined. If the slope was greater than one, oscillations occurred that may cause localized conduction delay or block resulting in a wave break giving rise to atrial fibrillation.

These different patterns of conduction are reflected in the morphology of electrograms recorded with mapping during induced AF. Single potentials were indicative of rapid uniform conduction, short double potentials indicated collision, long double potentials were indicative of conduction block, while fragmented potentials were markers for pivoting points or slow conduction (figure 6) [111,112].

Sites of fragmented potentials or complex fractionated atrial electrograms are potential targets for radiofrequency ablation to terminate AF as they may represent critical areas from which AF originates and perpetuates. (See "Atrial fibrillation: Catheter ablation".)

This phenomenon has been termed “microreentry” to distinguish it from classic reentry in which the same reentrant pathway is repetitively traversed. The impulse may circulate around a central line of functional block, so-called leading circle reentry; this type of reentry tends not to be stable but rather to drift through the atria until it is extinguished. The perpetuation of AF may also depend importantly upon macroreentry around natural orifices and structures in the atrium, which provides a rationale and anatomic landmarks for ablative treatment. The collision of wavefronts cancels many atrial depolarizations that might otherwise reach the AV node, resulting in a slower heart rate than might otherwise have occurred (figure 7A-B).

Although multiple wandering wavelets probably account for the majority of AF, one study reported nine patients in whom a single, rapidly firing focus was identified with electrophysiologic mapping [113]. Organized and rapid atrial activity with a centrifugal and consistent pattern of atrial activation resulted from this focus, but it fired irregularly with striking and abrupt changes in atrial cycle lengths. In most of the patients, the focus was near the ostia of great vessels and was amenable to radiofrequency ablation (figure 8 and figure 9).

Small reentrant sources, called rotors, may drive or maintain AF in some cases. These rotors result in a hierarchical distribution of frequencies throughout the atria that may be identified with spectral analysis of intracardiac recordings. Ablation of such sites has terminated paroxysmal AF, suggesting that they may play an important role [114], but it is not clear that the rotors are responsible for AF or are fixed in most instances. AF may be chaotic and have wavelets and rotors that are secondary rather than the predominant cause of AF [115]. However, antral pulmonary venous reentrant and focal drivers may be responsible for AF [54]. The complexity of such drivers increase with prolonged AF. These sites are often localized near the PV orifices in patients with paroxysmal AF, and are more often localized to the left or right atria in patients with chronic AF [103].

The fibrillating atrium cannot be captured by pacing when the atrial electrograms are disorganized. This observation supports the presence of microreentry, since there is no excitable gap (or it is very small) to permit capture. However, when type I (figure 9) AF (which has organized atrial electrograms) is induced by rapid atrial pacing, the fibrillating atrium can be captured with rapid atrial pacing, suggesting the presence of an excitable gap [116].

ROLE OF THE ATRIOVENTRICULAR NODE — The atrioventricular (AV) node regulates the number of atrial impulses that reach the ventricle. The ventricular rate in atrial fibrillation (AF) is typically irregularly irregular, with a ventricular rate that may be slow, moderate, or rapid depending on the capacity of the AV node to conduct impulses. The rate of AV nodal conduction is dependent upon multiple factors, including electrical properties of the node and the influence of the autonomic nervous system [117]. In addition, the use drugs such as digoxin, calcium channel blockers, or beta blockers may influence AV nodal function. There also may be a circadian rhythm for both AV nodal refractoriness and concealed conduction, accounting for the circadian variation in ventricular response rate [118].

AV nodal tissue consists of so-called "slow response" fibers, which depend on a mixed calcium/sodium current. This current is often called the inward calcium current, since in a normal physiologic environment, the ions are almost exclusively calcium. The mixed current uses a kinetically slow channel and is responsible for phase 0 depolarization. (See "Cardiac excitability, mechanisms of arrhythmia, and action of antiarrhythmic drugs".)

These characteristics lead to properties that are quite different from "fast-response" tissue in the atria, which as noted above, depend on an inward sodium current that uses a kinetically fast channel for phase 0 depolarization [8,119]:

Partial and complete reactivation returns only 100 ms or more after return to the diastolic potential (versus 10 to 50 ms in the atria).

The refractory period changes little as a function of rate.

Conduction velocity is relatively slow, ranging from 0.01 to 0.1 m/s.

Unlike tissue generating a fast action potential that has an all-or-none response (ie, the velocity of impulse conduction is similar at all stimulation rates until block occurs), tissue that generates a slow action potential exhibits a graded or decremental response, in which the velocity of impulse conduction slows as the stimulation rate increases.

As noted above, the ventricular rate usually ranges 90 and 170 beats/min. Ventricular rates below 60 beats/min are seen with AV nodal disease, drugs that affect conduction, and high vagal tone as can occur in a well-conditioned athlete. Ventricular rates above 200 beats/min suggest catecholamine excess, parasympathetic withdrawal, or the existence of an accessory bypass tract as occurs in the preexcitation syndrome. The QRS complexes are widened in the last setting and must be distinguished from a rate-related or underlying bundle branch block.

In the classical view, the AV node is bombarded by impulses from the fibrillating atria. Some impulses traverse the AV node and reach the specialized infranodal conduction system and then the ventricles. However, most atrial impulses penetrate the AV node from varying distances and then are extinguished when they encounter the refractoriness of an earlier wavefront; this phenomenon of concealed conduction in turn creates a refractory wave that affects succeeding impulses. The failure of the refractory period to shorten with increasing rate (as occurs in the atria) further decreases the likelihood of an impulse traversing the AV node.

Anatomically distinct AV nodal inputs, called the slow and fast pathways, are involved in the ventricular response to AF. The importance of these pathways has been demonstrated in radiofrequency ablation studies in which ablation reduced the number of beats that successfully reached the infranodal conduction system and the ventricles [120-123]. (See "Atrioventricular nodal reentrant tachycardia".)

In addition to its intrinsic properties, the AV node is richly supplied and affected by both components of the autonomic nervous system. AV conduction is enhanced and refractoriness reduced by the sympathetic fibers, and conduction reduced and refractoriness prolonged by the parasympathetic fibers.

The net effect of the electrophysiologic properties of the AV node is that the rate of conduction into the specialized infranodal conduction system is (fortunately) much slower than the rate of the fibrillating atria. In some cases, the high degree of refractoriness in the AV node with AF results in high-grade or third-degree block. In this setting, the pacemaker that controls the ventricles is below the AV node. (See "The electrocardiogram in atrial fibrillation".)

In patients with the preexcitation syndrome, the AV node is bypassed by "fast-response" tracts, which activate and reactivate much faster than the AV node and are therefore capable of rapid conduction. The development of AF in such a patient can result in very rapid transmission of atrial impulses to the ventricles [123] and can rarely cause ventricular fibrillation [15]. (See "The electrocardiogram in atrial fibrillation".)

It is also important to recognize that the presence of an accessory pathway can increase the propensity for development of AF, particularly if the pathway conducts only antegrade. The mechanism(s) for how accessory pathways affect the presence of AF is uncertain. In patients with AF who have Wolff-Parkinson-White (WPW) syndrome, catheter ablation of the accessory pathway is indicated to lower the sudden death risk but also to decrease the probability of recurrent AF.

Unexpected ventricular rates — The ventricular response to AF characteristically is irregularly irregular although it may appear regular in the presence of complete AV block. The usual ventricular rate in AF is between 90 and 170 beats per minute in the absence of AV node disease, drugs that affect conduction, or enhanced vagal inputs. Ventricular rates that are clearly outside this range suggest some concurrent problem:

A ventricular rate below 60 beats per minute, in the absence of AV nodal blocking agents, suggests AV nodal disease that may be associated with the sinus node dysfunction. (See "Sinus node dysfunction: Epidemiology, etiology, and natural history".)

A ventricular rate above 170 beats per minute suggests thyrotoxicosis, catecholamine excess, parasympathetic withdrawal, or the existence of an accessory bypass tract in the preexcitation syndrome. (See "Epidemiology, risk factors, and prevention of atrial fibrillation".)

SPECIFIC CLINICAL SITUATIONS

Late recurrent AF after catheter ablation — The etiology of late recurrent atrial fibrillation (AF) following pulmonary vein isolation (PVI) has been debated. In some cases, triggering foci outside of the PVs may initiate AF [124-127]. Alternatively, persistence of the substrate for maintaining AF (abnormal electrical properties of the atria themselves) may be more important than the triggering foci, especially in chronic AF.

However, there is increasing evidence that when AF does recur late after PVI, it often represents incomplete electrical isolation of the PVs, either due to resumption of conduction across the ablation scar or to residual conduction in PVs that were not successfully ablated. Most [128-131], but not all [132], studies of the former mechanism support the hypothesis that resumption of PV-left atrial (LA) conduction is associated with an increased risk of recurrent AF. However, recurrent conduction across ablated lesions is more common than clinically evident recurrent AF [130,133].

Pre-existing LA scarring may predispose patients to late recurrence. In a series of 700 consecutive patients undergoing first-time PVI, scarring was detected in 6 percent [134]. These patients had a much higher rate of recurrence than those without scarring (57 versus 19 percent).

Possible causes of scarring include atrial remodeling and inflammation. The patients with scarring had significant elevations in serum C-reactive protein (CRP) compared to those without scarring (5.9 versus 0.31 mg/L). This is consistent with other studies showing a relationship between serum CRP and AF [135]. (See "Epidemiology, risk factors, and prevention of atrial fibrillation", section on 'Inflammation and infection'.)

After cardiac surgery — AF occurs frequently (approximately one of four patients) after cardiac surgery. Nonuniform atrial conduction is greatest on days two and three in this setting, and the longest atrial conduction time is greatest on day three after open heart surgery; these abnormalities coincide with the time of greatest risk for AF [136]. The degree of atrial inflammation after surgery in dogs was associated with a proportional increase in the inhomogeneity of atrial conduction and in the duration of AF; antiinflammatory therapy decreased the inhomogeneity [137]. Nevertheless, the mechanism of AF in the postoperative period is likely multifactorial. It is important to note that in most of the patients, especially those without a prior history of AF, that the AF is self-limited, and antiarrhythmic drug therapy can usually be stopped two to three months following surgery when the inflammation has subsided. (See "Atrial fibrillation and flutter after cardiac surgery".)

Hyperthyroidism — It is well established that hyperthyroidism can increase susceptibility to development of AF. As a consequence, all patients with new onset AF should have some measure of thyroid function tested. Successful treatment of the hyperthyroidism often results in elimination of the AF.

Obesity — Obesity has been associated with AF and it is possible that both are related mechanistically [138]. In a sheep model, weight gain was associated with increased left atrial volume, fibrosis, inflammatory infiltrates, and lipidosis. There was reduced conduction velocity in atrial tissue and increased inducible and spontaneous AF with obesity. Atrial endothelin-A and -B receptors, endothelin-1, atrial interstitial and cytoplasmic transforming growth factor beta1, and platelet-derived growth factor were higher with obesity. In a clinical study of 110 patients undergoing AF ablation versus 20 reference patients without AF, pericardial fat volumes were associated with AF, its chronicity, and its symptom burden. Pericardial fat predicted AF recurrence post-ablation [139]. Associations persisted after adjusting for body weight but body mass index was not associated with these outcomes in multivariate-adjusted models. In another report [140], weight management with subsequent weight loss was associated with improved AF symptom burden scores, symptom severity scores, number of episodes, and cumulative duration of AF. This preliminary information does not yet prove that obesity causes AF by any specific mechanism. In a study of atrial sheep myocytes, acute, short-term incubation in free fatty acids resulted in no differences in passive or active properties of isolated left atrial myocytes but stearic acid reduced membrane capacitance and abbreviated the action potential duration, likely due to a reduction of the L-type calcium and of the transient outward potassium currents [141].

Sleep apnea — Sleep apnea has been associated with AF [142]. It is uncertain if this association is mediated by other shared comorbidities (eg, obesity, alcohol). Observational data suggest that the treatment of sleep apnea is associated with reduction in AF. Potential mechanisms underlying the association between sleep apnea and AF include atrial distention, increased systemic inflammation and oxidative stress, autonomic perturbations, hypoxia, hypercapnia, effects on connexin dysregulation and increased fibrosis, metabolic dysregulation, and the possible role of epicardial fat secretome [142].

Alcohol — Alcohol is a known provocateur of AF. The mechanism may be a direct toxic effect. Alcohol can reduce L-type calcium and sodium current density in rats. Furthermore, there can be an increase in I KACH activity, shortening the action potential. Additionally, alcohol can shorten pulmonary vein action potential duration by increasing ITO and cause an upregulation in protein expression of the potassium channel Kir 3.1. Alcohol can affect sympathetic and parasympathetic modulation. Alcohol can damage cellular gap junctions, cause inflammation with oxidative stress, autonomic dysregulation, and affect atrial electrical channels, thereby leading to AF [3].

GENETICS OF AF — Over the last decade, a preponderance of evidence suggests a large genetic contribution to atrial fibrillation (AF) [143,144]. Having a family member with AF is associated with a 40 percent increased risk for the arrhythmia [145]. Initially, traditional genetic techniques such as linkage analysis led to the discovery of rare, monogenic causes of AF. The first such study identified a genetic locus for AF using a series of related families with early onset AF [146]. A later study identified the first gene for familial AF [147]. Using a large Chinese kindred with autosomal dominant AF, they found a gain-of-function mutation in KCNQ1 (the gene encoding the α subunit of the potassium channel current, IKs). Since then, several additional gain-of-function variants have been identified in KCNQ1 [148,149]. In addition to KCNQ1, mutations have been identified in other potassium channels genes, including KCNA5 [150], KCND3 [151], and KCNJ2 [152], and accessory subunits KCNE1 [153], KCNE2 [154], KCNE3 [155], and KCNE5 [156,157].

The majority of these functionally validated, AF-associated potassium channel variants have a gain-of-function channel, with an expected shortening of the atrial action potential duration and atrial refractory period. Variation in sodium channel subunits has also been identified as an important factor in the development of familial AF, with AF-causing variants observed in both the major cardiac sodium channel alpha subunit SCN5A [158] and its associated beta subunits [159,160]. Several variants have also been identified in genes that do not directly alter the atrial action potential, but instead would be expected to cause AF through alternative mechanisms, eg, somatic mutations in GJA5, which encodes the gap junctional protein; connexin 40, a frameshift mutation that resulted in early truncation of NPPA [161], which encodes for the precursor for atrial natriuretic peptide; and genetic variation in several developmentally related cardiac transcription factors, ie, NKX2.5, PITX2, GATA4, GATA5, and GATA6 [160,162,163].

Genome-wide association studies (GWAS) have been used to identify genetic loci associated with AF. GWAS rely on the unbiased comparison of common single-nucleotide polymorphisms (SNPs) throughout the genome, with SNPs occurring with different frequency in individuals with a disease versus controls being used to localize disease-related genetic loci. The first GWAS performed for AF identified a region on chromosome 4q25, which was associated with AF in those of European and Asian descent [164]. Subsequently, these findings were broadly replicated in individuals of European, Asian, and African descent [165,166]. Genetic variants on chromosome 4q25 that are most significantly associated with AF reside about 150 kilobases upstream of the nearest gene PITX2. PITX2 encodes the paired-like homeodomain transcription factor 2, which helps determine cardiac laterality, suppresses the default expression of a sinoatrial nodal gene programme in the left atrium, and encodes the pulmonary venous myocardium [167]. In addition, PITX2 is associated with formation of the pulmonary veins. These findings are particularly interesting in light of the fact that AF triggers frequently arise in the pulmonary veins. In addition to the role of PITX2 in development, studies demonstrate a role for the PitX2c transcript in expression of gene-encoding ion channels, calcium cycling proteins, and gap junctions; these direct electrophysiological influences likely lead to formation of substrate for triggered activity as well as reentry [168].

The AF-associated ZFHX3 locus was shown to cause atrial dysregulation in knockout mice. Specifically, ZFHX3 knockout lead to significant increase in incidence and burden of AF, with alterations in conduction velocity, atrial action potential duration, calcium handling and the development of atrial enlargement and thrombus, and dilated cardiomyopathy [169].

Related analysis identified the same genomic region as being associated with an increased risk of cardioembolic stroke [160,170] and a prolonged PR interval [171]. To date, GWAS have identified 14 genomic regions of susceptibility for AF, with 17 independent signals at these loci [172]. These include the ZFHX3 gene that encodes a zinc finger homeobox transcription factor [173], the KCNN3 gene that encodes the SK3 potassium channel [174], and the PRRX1 gene that encodes a member of the paired-related homeobox gene family [173]. Whole exome and genome sequencing has been increasingly used to identify rare variants associated with AF [160]. For example, Oleson et al reported a much higher prevalence of rare variants in genes associated with AF (KCNQ1, KCNH2, SCN5A, KCNA5, KCND3, KCNE1, 2, 5, KCNJ2, SCN1-3B, NPPA, and GJA5) in early onset, lone AF patients than in the background population [175]. This approach is beginning to identify rare candidate variants in genes not previously linked to other types of Mendelian disease and thus may offer new insights into AF pathogenesis and disease pathways that could ultimately provide novel therapeutic targets for this common condition.

A 2018 meta-analysis of genome-wide association studies (GWAS) for AF to date, consisting of more than 500,000 individuals, sought to identify AF-associated genes at the GWAS loci by performing RNA sequencing and expression quantitative trait locus analyses in 101 left atrial samples (which is the most relevant tissue for AF) [176]. A transcriptome-wide analysis was also performed; this analysis identified 57 AF-associated genes, 42 of which overlap with GWAS loci. The identified loci-implicated genes enriched within cardiac developmental, electrophysiological, contractile, and structural pathways.

There have been several attempts to create polygenic risk scores in order to incorporate the influence of multiple disease-associated genetic risk variants on the risk of development of AF [177].

SUMMARY

Introduction – The precise mechanisms by which age and other risk factors such as hypertension, coronary artery or valvular heart disease, or heart failure increase the propensity for development of atrial fibrillation (AF) are poorly understood (figure 1). These conditions may affect the triggers of or the substrate for the maintenance of AF. (See 'Introduction' above.)

Triggers and substrate – These mechanisms are complex and involve a dynamic interplay between the triggers and substrate abnormalities. It is likely that short-lived episodes are due to specific triggers, including autonomic perturbations, focal discharges, specific reentry circuits in the pulmonary veins (PVs), and effects of stretch, whereas inflammation, dilatation, fibrosis, repolarization abnormalities, and conduction disturbances allow for perpetuation of episodes of AF. (See 'Mechanisms of atrial fibrillation: triggers and substrates' above.)

Triggers of AF – AF is most often initiated (triggered) by rapid firing from the PV. (See 'Triggers of AF' above.)

Electrical remodeling – Paroxysmal AF commonly precedes chronic AF. This suggests that, in addition to other predisposing factors, AF may play a role in its own natural history. (See 'Electrical remodeling' above.)

Autonomic nervous system – This likely influences the initiation and perpetuation of AF. (See 'Role of the autonomic nervous system' above.)

  1. Wasmer K, Eckardt L, Breithardt G. Predisposing factors for atrial fibrillation in the elderly. J Geriatr Cardiol 2017; 14:179.
  2. Goudis CA, Korantzopoulos P, Ntalas IV, et al. Obesity and atrial fibrillation: A comprehensive review of the pathophysiological mechanisms and links. J Cardiol 2015; 66:361.
  3. Voskoboinik A, Prabhu S, Ling LH, et al. Alcohol and Atrial Fibrillation: A Sobering Review. J Am Coll Cardiol 2016; 68:2567.
  4. Turagam MK, Velagapudi P, Kocheril AG, Alpert MA. Commonly consumed beverages in daily life: do they cause atrial fibrillation? Clin Cardiol 2015; 38:317.
  5. Sanchis-Gomar F, Perez-Quilis C, Lippi G, et al. Atrial fibrillation in highly trained endurance athletes - Description of a syndrome. Int J Cardiol 2017; 226:11.
  6. Schotten U, Verheule S, Kirchhof P, Goette A. Pathophysiological mechanisms of atrial fibrillation: a translational appraisal. Physiol Rev 2011; 91:265.
  7. Allessie MA, Konings K, Kirchhof CJ, Wijffels M. Electrophysiologic mechanisms of perpetuation of atrial fibrillation. Am J Cardiol 1996; 77:10A.
  8. Arnsdorf, MF. Electrophysiology and biophysics of cardiac excitation. In: Current Concepts of Cardiovascular Physiology, 1st ed, Garfein, O (Eds), Academic Press, Orlando 1990. p.465.
  9. Fozzard, HA, Arnsdorf, MF. Cardiac electrophysiology. In: The Heart and Cardiovascular System, Fozzard, HA, Haber, E, Jennings, RB, Katz, AM, Morgan, HE (Eds), Raven Press, New York 1991. p.63.
  10. Allessie MA, Boyden PA, Camm AJ, et al. Pathophysiology and prevention of atrial fibrillation. Circulation 2001; 103:769.
  11. SCHERF D. Studies on auricular tachycardia caused by aconitine administration. Proc Soc Exp Biol Med 1947; 64:233.
  12. Kalifa J, Jalife J, Zaitsev AV, et al. Intra-atrial pressure increases rate and organization of waves emanating from the superior pulmonary veins during atrial fibrillation. Circulation 2003; 108:668.
  13. Thijssen VL, Ausma J, Liu GS, et al. Structural changes of atrial myocardium during chronic atrial fibrillation. Cardiovasc Pathol 2000; 9:17.
  14. Wijffels MC, Kirchhof CJ, Dorland R, Allessie MA. Atrial fibrillation begets atrial fibrillation. A study in awake chronically instrumented goats. Circulation 1995; 92:1954.
  15. Klein GJ, Bashore TM, Sellers TD, et al. Ventricular fibrillation in the Wolff-Parkinson-White syndrome. N Engl J Med 1979; 301:1080.
  16. Elvan A, Huang XD, Pressler ML, Zipes DP. Radiofrequency catheter ablation of the atria eliminates pacing-induced sustained atrial fibrillation and reduces connexin 43 in dogs. Circulation 1997; 96:1675.
  17. van der Velden HM, Ausma J, Rook MB, et al. Gap junctional remodeling in relation to stabilization of atrial fibrillation in the goat. Cardiovasc Res 2000; 46:476.
  18. Aimé-Sempé C, Folliguet T, Rücker-Martin C, et al. Myocardial cell death in fibrillating and dilated human right atria. J Am Coll Cardiol 1999; 34:1577.
  19. Elvan A, Wylie K, Zipes DP. Pacing-induced chronic atrial fibrillation impairs sinus node function in dogs. Electrophysiological remodeling. Circulation 1996; 94:2953.
  20. Daoud EG, Bogun F, Goyal R, et al. Effect of atrial fibrillation on atrial refractoriness in humans. Circulation 1996; 94:1600.
  21. Tieleman RG, Van Gelder IC, Crijns HJ, et al. Early recurrences of atrial fibrillation after electrical cardioversion: a result of fibrillation-induced electrical remodeling of the atria? J Am Coll Cardiol 1998; 31:167.
  22. Franz MR, Karasik PL, Li C, et al. Electrical remodeling of the human atrium: similar effects in patients with chronic atrial fibrillation and atrial flutter. J Am Coll Cardiol 1997; 30:1785.
  23. Pandozi C, Bianconi L, Villani M, et al. Electrophysiological characteristics of the human atria after cardioversion of persistent atrial fibrillation. Circulation 1998; 98:2860.
  24. Sheng X, Scherlag BJ, Yu L, et al. Prevention and reversal of atrial fibrillation inducibility and autonomic remodeling by low-level vagosympathetic nerve stimulation. J Am Coll Cardiol 2011; 57:563.
  25. Scherlag BJ, Nakagawa H, Jackman WM, et al. Non-pharmacological, non-ablative approaches for the treatment of atrial fibrillation: experimental evidence and potential clinical implications. J Cardiovasc Transl Res 2011; 4:35.
  26. Lu Z, Scherlag BJ, Lin J, et al. Autonomic mechanism for initiation of rapid firing from atria and pulmonary veins: evidence by ablation of ganglionated plexi. Cardiovasc Res 2009; 84:245.
  27. Zimmermann M, Kalusche D. Fluctuation in autonomic tone is a major determinant of sustained atrial arrhythmias in patients with focal ectopy originating from the pulmonary veins. J Cardiovasc Electrophysiol 2001; 12:285.
  28. Hsieh MH, Chiou CW, Wen ZC, et al. Alterations of heart rate variability after radiofrequency catheter ablation of focal atrial fibrillation originating from pulmonary veins. Circulation 1999; 100:2237.
  29. Bettoni M, Zimmermann M. Autonomic tone variations before the onset of paroxysmal atrial fibrillation. Circulation 2002; 105:2753.
  30. Shen MJ, Choi EK, Tan AY, et al. Neural mechanisms of atrial arrhythmias. Nat Rev Cardiol 2011; 9:30.
  31. Hou Y, Scherlag BJ, Lin J, et al. Ganglionated plexi modulate extrinsic cardiac autonomic nerve input: effects on sinus rate, atrioventricular conduction, refractoriness, and inducibility of atrial fibrillation. J Am Coll Cardiol 2007; 50:61.
  32. Hou Y, Scherlag BJ, Lin J, et al. Interactive atrial neural network: Determining the connections between ganglionated plexi. Heart Rhythm 2007; 4:56.
  33. Arora R, Ulphani JS, Villuendas R, et al. Neural substrate for atrial fibrillation: implications for targeted parasympathetic blockade in the posterior left atrium. Am J Physiol Heart Circ Physiol 2008; 294:H134.
  34. Ulphani JS, Arora R, Cain JH, et al. The ligament of Marshall as a parasympathetic conduit. Am J Physiol Heart Circ Physiol 2007; 293:H1629.
  35. Arora R, Ng J, Ulphani J, et al. Unique autonomic profile of the pulmonary veins and posterior left atrium. J Am Coll Cardiol 2007; 49:1340.
  36. Ogawa M, Zhou S, Tan AY, et al. Left stellate ganglion and vagal nerve activity and cardiac arrhythmias in ambulatory dogs with pacing-induced congestive heart failure. J Am Coll Cardiol 2007; 50:335.
  37. Ng J, Villuendas R, Cokic I, et al. Autonomic remodeling in the left atrium and pulmonary veins in heart failure: creation of a dynamic substrate for atrial fibrillation. Circ Arrhythm Electrophysiol 2011; 4:388.
  38. Chou CC, Chen PS. New concepts in atrial fibrillation: neural mechanisms and calcium dynamics. Cardiol Clin 2009; 27:35.
  39. Yamaguchi Y, Kumagai K, Nakashima H, Saku K. Long-term effects of box isolation on sympathovagal balance in atrial fibrillation. Circ J 2010; 74:1096.
  40. Coumel P. Autonomic influences in atrial tachyarrhythmias. J Cardiovasc Electrophysiol 1996; 7:999.
  41. Crawford T, Chugh A, Good E, et al. Clinical value of noninducibility by high-dose isoproterenol versus rapid atrial pacing after catheter ablation of paroxysmal atrial fibrillation. J Cardiovasc Electrophysiol 2010; 21:13.
  42. Chen PS, Tan AY. Autonomic nerve activity and atrial fibrillation. Heart Rhythm 2007; 4:S61.
  43. Nemirovsky D, Hutter R, Gomes JA. The electrical substrate of vagal atrial fibrillation as assessed by the signal-averaged electrocardiogram of the P wave. Pacing Clin Electrophysiol 2008; 31:308.
  44. Sharifov OF, Fedorov VV, Beloshapko GG, et al. Roles of adrenergic and cholinergic stimulation in spontaneous atrial fibrillation in dogs. J Am Coll Cardiol 2004; 43:483.
  45. Hirose M, Carlson MD, Laurita KR. Cellular mechanisms of vagally mediated atrial tachyarrhythmia in isolated arterially perfused canine right atria. J Cardiovasc Electrophysiol 2002; 13:918.
  46. Pappone C, Santinelli V, Manguso F, et al. Pulmonary vein denervation enhances long-term benefit after circumferential ablation for paroxysmal atrial fibrillation. Circulation 2004; 109:327.
  47. Verma A, Saliba WI, Lakkireddy D, et al. Vagal responses induced by endocardial left atrial autonomic ganglion stimulation before and after pulmonary vein antrum isolation for atrial fibrillation. Heart Rhythm 2007; 4:1177.
  48. Arruda M, Natale A. Ablation of permanent AF: adjunctive strategies to pulmonary veins isolation: targeting AF NEST in sinus rhythm and CFAE in AF. J Interv Card Electrophysiol 2008; 23:51.
  49. Porter M, Spear W, Akar JG, et al. Prospective study of atrial fibrillation termination during ablation guided by automated detection of fractionated electrograms. J Cardiovasc Electrophysiol 2008; 19:613.
  50. Katritsis D, Giazitzoglou E, Sougiannis D, et al. Complex fractionated atrial electrograms at anatomic sites of ganglionated plexi in atrial fibrillation. Europace 2009; 11:308.
  51. Stavrakis S. Low-level transcutaneous electrical vagus nerve stimulation. J Am Coll Cardiol 2015; :867.
  52. Narayan SM, Krummen DE, Shivkumar K, et al. Treatment of atrial fibrillation by the ablation of localized sources: CONFIRM (Conventional Ablation for Atrial Fibrillation With or Without Focal Impulse and Rotor Modulation) trial. J Am Coll Cardiol 2012; 60:628.
  53. Haissaguerre M, Hocini M, Denis A, et al. Driver domains in persistent atrial fibrillation. Circulation 2014; 130:530.
  54. Lim HS, Hocini M, Dubois R, et al. Complexity and Distribution of Drivers in Relation to Duration of Persistent Atrial Fibrillation. J Am Coll Cardiol 2017; 69:1257.
  55. Climent AM, Guillem MS, Fuentes L, et al. Role of atrial tissue remodeling on rotor dynamics: an in vitro study. Am J Physiol Heart Circ Physiol 2015; 309:H1964.
  56. Lee S, Sahadevan J, Khrestian CM, et al. Simultaneous Biatrial High-Density (510-512 Electrodes) Epicardial Mapping of Persistent and Long-Standing Persistent Atrial Fibrillation in Patients: New Insights Into the Mechanism of Its Maintenance. Circulation 2015; 132:2108.
  57. Kuklik P, Lau DH, Ganesan AN, et al. High-density mapping of atrial fibrillation in a chronic substrate: evidence for distinct modes of repetitive wavefront propagation. Int J Cardiol 2015; 199:407.
  58. Allessie MA. Atrial fibrillation-induced electrical remodeling in humans: what is the next step? Cardiovasc Res 1999; 44:10.
  59. Nattel S, Burstein B, Dobrev D. Atrial remodeling and atrial fibrillation: mechanisms and implications. Circ Arrhythm Electrophysiol 2008; 1:62.
  60. Wijffels MC, Kirchhof CJ, Dorland R, et al. Electrical remodeling due to atrial fibrillation in chronically instrumented conscious goats: roles of neurohumoral changes, ischemia, atrial stretch, and high rate of electrical activation. Circulation 1997; 96:3710.
  61. Fareh S, Villemaire C, Nattel S. Importance of refractoriness heterogeneity in the enhanced vulnerability to atrial fibrillation induction caused by tachycardia-induced atrial electrical remodeling. Circulation 1998; 98:2202.
  62. Raitt MH, Kusumoto W, Giraud G, McAnulty JH. Reversal of electrical remodeling after cardioversion of persistent atrial fibrillation. J Cardiovasc Electrophysiol 2004; 15:507.
  63. Brundel BJ, Van Gelder IC, Henning RH, et al. Ion channel remodeling is related to intraoperative atrial effective refractory periods in patients with paroxysmal and persistent atrial fibrillation. Circulation 2001; 103:684.
  64. Tieleman RG, De Langen C, Van Gelder IC, et al. Verapamil reduces tachycardia-induced electrical remodeling of the atria. Circulation 1997; 95:1945.
  65. Daoud EG, Knight BP, Weiss R, et al. Effect of verapamil and procainamide on atrial fibrillation-induced electrical remodeling in humans. Circulation 1997; 96:1542.
  66. Tse HF, Lau CP, Wang Q, et al. Effect of diltiazem on the recurrence rate of paroxysmal atrial fibrillation. Am J Cardiol 2001; 88:568.
  67. Tieleman RG, Blaauw Y, Van Gelder IC, et al. Digoxin delays recovery from tachycardia-induced electrical remodeling of the atria. Circulation 1999; 100:1836.
  68. Sticherling C, Oral H, Horrocks J, et al. Effects of digoxin on acute, atrial fibrillation-induced changes in atrial refractoriness. Circulation 2000; 102:2503.
  69. Vest JA, Wehrens XH, Reiken SR, et al. Defective cardiac ryanodine receptor regulation during atrial fibrillation. Circulation 2005; 111:2025.
  70. Li N, Chiang DY, Wang S, et al. Ryanodine receptor-mediated calcium leak drives progressive development of an atrial fibrillation substrate in a transgenic mouse model. Circulation 2014; 129:1276.
  71. Dobrev D, Wehrens XHT. Calcium-mediated cellular triggered activity in atrial fibrillation. J Physiol 2017; 595:4001.
  72. Nakashima H, Kumagai K, Urata H, et al. Angiotensin II antagonist prevents electrical remodeling in atrial fibrillation. Circulation 2000; 101:2612.
  73. Goette A, Arndt M, Röcken C, et al. Regulation of angiotensin II receptor subtypes during atrial fibrillation in humans. Circulation 2000; 101:2678.
  74. Pedersen OD, Bagger H, Kober L, Torp-Pedersen C. Trandolapril reduces the incidence of atrial fibrillation after acute myocardial infarction in patients with left ventricular dysfunction. Circulation 1999; 100:376.
  75. Vermes E, Tardif JC, Bourassa MG, et al. Enalapril decreases the incidence of atrial fibrillation in patients with left ventricular dysfunction: insight from the Studies Of Left Ventricular Dysfunction (SOLVD) trials. Circulation 2003; 107:2926.
  76. Jayachandran JV, Zipes DP, Weksler J, Olgin JE. Role of the Na(+)/H(+) exchanger in short-term atrial electrophysiological remodeling. Circulation 2000; 101:1861.
  77. Dzeshka MS, Lip GY, Snezhitskiy V, Shantsila E. Cardiac Fibrosis in Patients With Atrial Fibrillation: Mechanisms and Clinical Implications. J Am Coll Cardiol 2015; 66:943.
  78. Li D, Fareh S, Leung TK, Nattel S. Promotion of atrial fibrillation by heart failure in dogs: atrial remodeling of a different sort. Circulation 1999; 100:87.
  79. Polyakova V, Miyagawa S, Szalay Z, et al. Atrial extracellular matrix remodelling in patients with atrial fibrillation. J Cell Mol Med 2008; 12:189.
  80. Pellman J, Lyon RC, Sheikh F. Extracellular matrix remodeling in atrial fibrosis: mechanisms and implications in atrial fibrillation. J Mol Cell Cardiol 2010; 48:461.
  81. Corradi D, Callegari S, Maestri R, et al. Heme oxygenase-1 expression in the left atrial myocardium of patients with chronic atrial fibrillation related to mitral valve disease: its regional relationship with structural remodeling. Hum Pathol 2008; 39:1162.
  82. Corradi D, Callegari S, Benussi S, et al. Myocyte changes and their left atrial distribution in patients with chronic atrial fibrillation related to mitral valve disease. Hum Pathol 2005; 36:1080.
  83. Corradi D, Callegari S, Benussi S, et al. Regional left atrial interstitial remodeling in patients with chronic atrial fibrillation undergoing mitral-valve surgery. Virchows Arch 2004; 445:498.
  84. Nakajima H, Nakajima HO, Salcher O, et al. Atrial but not ventricular fibrosis in mice expressing a mutant transforming growth factor-beta(1) transgene in the heart. Circ Res 2000; 86:571.
  85. Lamirault G, Gaborit N, Le Meur N, et al. Gene expression profile associated with chronic atrial fibrillation and underlying valvular heart disease in man. J Mol Cell Cardiol 2006; 40:173.
  86. Lee KW, Everett TH 4th, Rahmutula D, et al. Pirfenidone prevents the development of a vulnerable substrate for atrial fibrillation in a canine model of heart failure. Circulation 2006; 114:1703.
  87. Gutierrez A, Van Wagoner DR. Oxidant and Inflammatory Mechanisms and Targeted Therapy in Atrial Fibrillation: An Update. J Cardiovasc Pharmacol 2015; 66:523.
  88. Harada M, Van Wagoner DR, Nattel S. Role of inflammation in atrial fibrillation pathophysiology and management. Circ J 2015; 79:495.
  89. Sovari AA. Cellular and Molecular Mechanisms of Arrhythmia by Oxidative Stress. Cardiol Res Pract 2016; 2016:9656078.
  90. Youn JY, Zhang J, Zhang Y, et al. Oxidative stress in atrial fibrillation: an emerging role of NADPH oxidase. J Mol Cell Cardiol 2013; 62:72.
  91. Negi S, Sovari AA, Dudley SC Jr. Atrial fibrillation: the emerging role of inflammation and oxidative stress. Cardiovasc Hematol Disord Drug Targets 2010; 10:262.
  92. Kim YM, Guzik TJ, Zhang YH, et al. A myocardial Nox2 containing NAD(P)H oxidase contributes to oxidative stress in human atrial fibrillation. Circ Res 2005; 97:629.
  93. Kim YM, Kattach H, Ratnatunga C, et al. Association of atrial nicotinamide adenine dinucleotide phosphate oxidase activity with the development of atrial fibrillation after cardiac surgery. J Am Coll Cardiol 2008; 51:68.
  94. Bonilla IM, Sridhar A, Györke S, et al. Nitric oxide synthases and atrial fibrillation. Front Physiol 2012; 3:105.
  95. Dudley SC Jr, Hoch NE, McCann LA, et al. Atrial fibrillation increases production of superoxide by the left atrium and left atrial appendage: role of the NADPH and xanthine oxidases. Circulation 2005; 112:1266.
  96. Carnes CA, Chung MK, Nakayama T, et al. Ascorbate attenuates atrial pacing-induced peroxynitrite formation and electrical remodeling and decreases the incidence of postoperative atrial fibrillation. Circ Res 2001; 89:E32.
  97. Shiroshita-Takeshita A, Schram G, Lavoie J, Nattel S. Effect of simvastatin and antioxidant vitamins on atrial fibrillation promotion by atrial-tachycardia remodeling in dogs. Circulation 2004; 110:2313.
  98. Rodrigo R, Korantzopoulos P, Cereceda M, et al. A randomized controlled trial to prevent post-operative atrial fibrillation by antioxidant reinforcement. J Am Coll Cardiol 2013; 62:1457.
  99. Ozaydin M, Icli A, Yucel H, et al. Metoprolol vs. carvedilol or carvedilol plus N-acetyl cysteine on post-operative atrial fibrillation: a randomized, double-blind, placebo-controlled study. Eur Heart J 2013; 34:597.
  100. Rashid S, Noor TA, Saeed H, et al. Association of gut microbiome dysbiosis with the progression of atrial fibrillation: A systematic review. Ann Noninvasive Electrocardiol 2023; 28:e13059.
  101. Lu D, Zou X, Zhang H. The Relationship Between Atrial Fibrillation and Intestinal Flora With Its Metabolites. Front Cardiovasc Med 2022; 9:948755.
  102. Wang Y, He Y, Li R, et al. Gut Microbiota in Patients with Postoperative Atrial Fibrillation Undergoing Off-Pump Coronary Bypass Graft Surgery. J Clin Med 2023; 12.
  103. Atienza F, Almendral J, Moreno J, et al. Activation of inward rectifier potassium channels accelerates atrial fibrillation in humans: evidence for a reentrant mechanism. Circulation 2006; 114:2434.
  104. Allessie, MA, Lammers, et al. Experimental evaluation of Moe's multiple wavelet hypothesis of atrial fibrillation. In: Cardiac Arrhythmias, Zipes, DP, Jalife, J (Eds), Grune and Stratton, Orlando 1985. p.265.
  105. Moe GK. On the multiple wavelet hypothesis of atrial fibrillation. Arch Int Pharmacodyn Ther 1962; 140:183.
  106. Konings KT, Kirchhof CJ, Smeets JR, et al. High-density mapping of electrically induced atrial fibrillation in humans. Circulation 1994; 89:1665.
  107. 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 1997; 95:511.
  108. Ramanna H, Hauer RN, Wittkampf FH, et al. Identification of the substrate of atrial vulnerability in patients with idiopathic atrial fibrillation. Circulation 2000; 101:995.
  109. Papageorgiou P, Monahan K, Boyle NG, et al. Site-dependent intra-atrial conduction delay. Relationship to initiation of atrial fibrillation. Circulation 1996; 94:384.
  110. Kim BS, Kim YH, Hwang GS, et al. Action potential duration restitution kinetics in human atrial fibrillation. J Am Coll Cardiol 2002; 39:1329.
  111. Konings KT, Smeets JL, Penn OC, et al. Configuration of unipolar atrial electrograms during electrically induced atrial fibrillation in humans. Circulation 1997; 95:1231.
  112. Holm M, Johansson R, Brandt J, et al. Epicardial right atrial free wall mapping in chronic atrial fibrillation. Documentation of repetitive activation with a focal spread--a hitherto unrecognised phenomenon in man. Eur Heart J 1997; 18:290.
  113. Jaïs P, Haïssaguerre M, Shah DC, et al. A focal source of atrial fibrillation treated by discrete radiofrequency ablation. Circulation 1997; 95:572.
  114. Sanders P, Berenfeld O, Hocini M, et al. Spectral analysis identifies sites of high-frequency activity maintaining atrial fibrillation in humans. Circulation 2005; 112:789.
  115. Marcotte CD, Grigoriev RO. Dynamical mechanism of atrial fibrillation: A topological approach. Chaos 2017; 27:093936.
  116. Daoud EG, Pariseau B, Niebauer M, et al. Response of type I atrial fibrillation to atrial pacing in humans. Circulation 1996; 94:1036.
  117. Janse, MJ. Propagation of atrial impulses through the atrioventricular node. In: Atrial Arrhythmias: Current Concepts and Management, Touboul, P, Waldo, AL (Eds), Mosby, St. Louis 1990. p.141.
  118. Hayano J, Sakata S, Okada A, et al. Circadian rhythms of atrioventricular conduction properties in chronic atrial fibrillation with and without heart failure. J Am Coll Cardiol 1998; 31:158.
  119. Fozzard HA, Arnsdorf MF. Cardiac electrophysiology. In: The Heart and Cardiovascular System, Fozzard HA, Haber E, Jennings RB, et al (Eds), Raven Press, New York 1991. p.63.
  120. Fleck RP, Chen PS, Boyce K, et al. Radiofrequency modification of atrioventricular conduction by selective ablation of the low posterior septal right atrium in a patient with atrial fibrillation and a rapid ventricular response. Pacing Clin Electrophysiol 1993; 16:377.
  121. Williamson BD, Man KC, Daoud E, et al. Radiofrequency catheter modification of atrioventricular conduction to control the ventricular rate during atrial fibrillation. N Engl J Med 1994; 331:910.
  122. Della Bella P, Carbucicchio C, Tondo C, Riva S. Modulation of atrioventricular conduction by ablation of the "slow" atrioventricular node pathway in patients with drug-refractory atrial fibrillation or flutter. J Am Coll Cardiol 1995; 25:39.
  123. Kreiner G, Heinz G, Siostrzonek P, Gössinger HD. Effect of slow pathway ablation on ventricular rate during atrial fibrillation. Dependence on electrophysiological properties of the fast pathway. Circulation 1996; 93:277.
  124. Lin WS, Tai CT, Hsieh MH, et al. Catheter ablation of paroxysmal atrial fibrillation initiated by non-pulmonary vein ectopy. Circulation 2003; 107:3176.
  125. Chen SA, Tai CT. Catheter ablation of atrial fibrillation originating from the non-pulmonary vein foci. J Cardiovasc Electrophysiol 2005; 16:229.
  126. Lee SH, Tai CT, Hsieh MH, et al. Predictors of non-pulmonary vein ectopic beats initiating paroxysmal atrial fibrillation: implication for catheter ablation. J Am Coll Cardiol 2005; 46:1054.
  127. Haïssaguerre M, Hocini M, Sanders P, et al. Localized sources maintaining atrial fibrillation organized by prior ablation. Circulation 2006; 113:616.
  128. Ouyang F, Ernst S, Chun J, et al. Electrophysiological findings during ablation of persistent atrial fibrillation with electroanatomic mapping and double Lasso catheter technique. Circulation 2005; 112:3038.
  129. Gerstenfeld EP, Callans DJ, Dixit S, et al. Incidence and location of focal atrial fibrillation triggers in patients undergoing repeat pulmonary vein isolation: implications for ablation strategies. J Cardiovasc Electrophysiol 2003; 14:685.
  130. Nanthakumar K, Plumb VJ, Epstein AE, et al. Resumption of electrical conduction in previously isolated pulmonary veins: rationale for a different strategy? Circulation 2004; 109:1226.
  131. Verma A, Kilicaslan F, Pisano E, et al. Response of atrial fibrillation to pulmonary vein antrum isolation is directly related to resumption and delay of pulmonary vein conduction. Circulation 2005; 112:627.
  132. Pratola C, Baldo E, Notarstefano P, et al. Radiofrequency ablation of atrial fibrillation: is the persistence of all intraprocedural targets necessary for long-term maintenance of sinus rhythm? Circulation 2008; 117:136.
  133. Cappato R, Negroni S, Pecora D, et al. Prospective assessment of late conduction recurrence across radiofrequency lesions producing electrical disconnection at the pulmonary vein ostium in patients with atrial fibrillation. Circulation 2003; 108:1599.
  134. Verma A, Wazni OM, Marrouche NF, et al. Pre-existent left atrial scarring in patients undergoing pulmonary vein antrum isolation: an independent predictor of procedural failure. J Am Coll Cardiol 2005; 45:285.
  135. Aviles RJ, Martin DO, Apperson-Hansen C, et al. Inflammation as a risk factor for atrial fibrillation. Circulation 2003; 108:3006.
  136. Tsikouris JP, Kluger J, Song J, White CM. Changes in P-wave dispersion and P-wave duration after open heart surgery are associated with the peak incidence of atrial fibrillation. Heart Lung 2001; 30:466.
  137. Ishii Y, Schuessler RB, Gaynor SL, et al. Inflammation of atrium after cardiac surgery is associated with inhomogeneity of atrial conduction and atrial fibrillation. Circulation 2005; 111:2881.
  138. Abed HS, Samuel CS, Lau DH, et al. Obesity results in progressive atrial structural and electrical remodeling: implications for atrial fibrillation. Heart Rhythm 2013; 10:90.
  139. Wong CX, Abed HS, Molaee P, et al. Pericardial fat is associated with atrial fibrillation severity and ablation outcome. J Am Coll Cardiol 2011; 57:1745.
  140. Abed HS, Wittert GA, Leong DP, et al. Effect of weight reduction and cardiometabolic risk factor management on symptom burden and severity in patients with atrial fibrillation: a randomized clinical trial. JAMA 2013; 310:2050.
  141. O'Connell RP, Musa H, Gomez MS, et al. Free Fatty Acid Effects on the Atrial Myocardium: Membrane Ionic Currents Are Remodeled by the Disruption of T-Tubular Architecture. PLoS One 2015; 10:e0133052.
  142. Mehra R, Chung MK, Olshansky B, et al. Sleep-Disordered Breathing and Cardiac Arrhythmias in Adults: Mechanistic Insights and Clinical Implications: A Scientific Statement From the American Heart Association. Circulation 2022; 146:e119.
  143. Ellinor PT, Yoerger DM, Ruskin JN, MacRae CA. Familial aggregation in lone atrial fibrillation. Hum Genet 2005; 118:179.
  144. Darbar D, Herron KJ, Ballew JD, et al. Familial atrial fibrillation is a genetically heterogeneous disorder. J Am Coll Cardiol 2003; 41:2185.
  145. Lubitz SA, Yin X, Fontes JD, et al. Association between familial atrial fibrillation and risk of new-onset atrial fibrillation. JAMA 2010; 304:2263.
  146. Brugada R, Tapscott T, Czernuszewicz GZ, et al. Identification of a genetic locus for familial atrial fibrillation. N Engl J Med 1997; 336:905.
  147. Chen YH, Xu SJ, Bendahhou S, et al. KCNQ1 gain-of-function mutation in familial atrial fibrillation. Science 2003; 299:251.
  148. Hasegawa K, Ohno S, Ashihara T, et al. A novel KCNQ1 missense mutation identified in a patient with juvenile-onset atrial fibrillation causes constitutively open IKs channels. Heart Rhythm 2014; 11:67.
  149. Bartos DC, Duchatelet S, Burgess DE, et al. R231C mutation in KCNQ1 causes long QT syndrome type 1 and familial atrial fibrillation. Heart Rhythm 2011; 8:48.
  150. Christophersen IE, Olesen MS, Liang B, et al. Genetic variation in KCNA5: impact on the atrial-specific potassium current IKur in patients with lone atrial fibrillation. Eur Heart J 2013; 34:1517.
  151. Olesen MS, Refsgaard L, Holst AG, et al. A novel KCND3 gain-of-function mutation associated with early-onset of persistent lone atrial fibrillation. Cardiovasc Res 2013; 98:488.
  152. Deo M, Ruan Y, Pandit SV, et al. KCNJ2 mutation in short QT syndrome 3 results in atrial fibrillation and ventricular proarrhythmia. Proc Natl Acad Sci U S A 2013; 110:4291.
  153. Olesen MS, Bentzen BH, Nielsen JB, et al. Mutations in the potassium channel subunit KCNE1 are associated with early-onset familial atrial fibrillation. BMC Med Genet 2012; 13:24.
  154. Yang Y, Xia M, Jin Q, et al. Identification of a KCNE2 gain-of-function mutation in patients with familial atrial fibrillation. Am J Hum Genet 2004; 75:899.
  155. Lundby A, Ravn LS, Svendsen JH, et al. KCNE3 mutation V17M identified in a patient with lone atrial fibrillation. Cell Physiol Biochem 2008; 21:47.
  156. Ravn LS, Aizawa Y, Pollevick GD, et al. Gain of function in IKs secondary to a mutation in KCNE5 associated with atrial fibrillation. Heart Rhythm 2008; 5:427.
  157. Enriquez A, Antzelevitch C, Bismah V, Baranchuk A. Atrial fibrillation in inherited cardiac channelopathies: From mechanisms to management. Heart Rhythm 2016; 13:1878.
  158. Darbar D, Kannankeril PJ, Donahue BS, et al. Cardiac sodium channel (SCN5A) variants associated with atrial fibrillation. Circulation 2008; 117:1927.
  159. Olesen MS, Holst AG, Svendsen JH, et al. SCN1Bb R214Q found in 3 patients: 1 with Brugada syndrome and 2 with lone atrial fibrillation. Heart Rhythm 2012; 9:770.
  160. Tucker NR, Ellinor PT. Emerging directions in the genetics of atrial fibrillation. Circ Res 2014; 114:1469.
  161. Hodgson-Zingman DM, Karst ML, Zingman LV, et al. Atrial natriuretic peptide frameshift mutation in familial atrial fibrillation. N Engl J Med 2008; 359:158.
  162. Huang RT, Xue S, Xu YJ, et al. A novel NKX2.5 loss-of-function mutation responsible for familial atrial fibrillation. Int J Mol Med 2013; 31:1119.
  163. Zhou YM, Zheng PX, Yang YQ, et al. A novel PITX2c loss‑of‑function mutation underlies lone atrial fibrillation. Int J Mol Med 2013; 32:827.
  164. Gudbjartsson DF, Arnar DO, Helgadottir A, et al. Variants conferring risk of atrial fibrillation on chromosome 4q25. Nature 2007; 448:353.
  165. Kääb S, Darbar D, van Noord C, et al. Large scale replication and meta-analysis of variants on chromosome 4q25 associated with atrial fibrillation. Eur Heart J 2009; 30:813.
  166. Delaney JT, Jeff JM, Brown NJ, et al. Characterization of genome-wide association-identified variants for atrial fibrillation in African Americans. PLoS One 2012; 7:e32338.
  167. Bapat A, Anderson CD, Ellinor PT, Lubitz SA. Genomic basis of atrial fibrillation. Heart 2018; 104:201.
  168. Gutierrez A, Chung MK. Genomics of Atrial Fibrillation. Curr Cardiol Rep 2016; 18:55.
  169. Jameson HS, Hanley A, Hill MC, et al. Loss of the Atrial Fibrillation-Related Gene, Zfhx3, Results in Atrial Dilation and Arrhythmias. Circ Res 2023; 133:313.
  170. Shi L, Li C, Wang C, et al. Assessment of association of rs2200733 on chromosome 4q25 with atrial fibrillation and ischemic stroke in a Chinese Han population. Hum Genet 2009; 126:843.
  171. Kolek MJ, Parvez B, Muhammad R, et al. A common variant on chromosome 4q25 is associated with prolonged PR interval in subjects with and without atrial fibrillation. Am J Cardiol 2014; 113:309.
  172. Tucker NR, Clauss S, Ellinor PT. Common variation in atrial fibrillation: navigating the path from genetic association to mechanism. Cardiovasc Res 2016; 109:493.
  173. Ellinor PT, Lunetta KL, Albert CM, et al. Meta-analysis identifies six new susceptibility loci for atrial fibrillation. Nat Genet 2012; 44:670.
  174. Ellinor PT, Lunetta KL, Glazer NL, et al. Common variants in KCNN3 are associated with lone atrial fibrillation. Nat Genet 2010; 42:240.
  175. Olesen MS, Andreasen L, Jabbari J, et al. Very early-onset lone atrial fibrillation patients have a high prevalence of rare variants in genes previously associated with atrial fibrillation. Heart Rhythm 2014; 11:246.
  176. Roselli C, Chaffin MD, Weng LC, et al. Multi-ethnic genome-wide association study for atrial fibrillation. Nat Genet 2018; 50:1225.
  177. Kavousi M, Ellinor PT. Polygenic risk scores for prediction of atrial fibrillation. Neth Heart J 2023; 31:1.
Topic 16402 Version 30.0

References

آیا می خواهید مدیلیب را به صفحه اصلی خود اضافه کنید؟