Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Aug 1.
Published in final edited form as: Biochim Biophys Acta. 2011 Aug 9;1818(8):1831–1843. doi: 10.1016/j.bbamem.2011.08.006

The Connexin43 Carboxyl Terminus and Cardiac Gap Junction Organization

Joseph A Palatinus 1, J Matthew Rhett 1, Robert G Gourdie 1,*
PMCID: PMC3244566  NIHMSID: NIHMS322426  PMID: 21856279

Abstract

The precise spatial order of gap junctions at intercalated disks in adult ventricular myocardium is thought vital for maintaining cardiac synchrony. Breakdown or remodeling of this order is a hallmark of arrhythmic disease of the heart. The principal component of gap junction channels between ventricular cardiomyocytes is connexin43 (Cx43). Protein-protein interactions and modifications of the carboxyl-terminus of Cx43 are key determinants of gap junction function, size, distribution and organization during normal development and in disease processes. Here, we review data on the role of proteins interacting with the Cx43 carboxyl-terminus in the regulation of cardiac gap junction organization, with particular emphasis on Zonula Occludens-1. The rapid progress in this area suggests that in coming years we are likely to develop a fuller understanding of the molecular mechanisms causing pathologic remodeling of gap junctions. With these advances come the promise of novel approach to the treatment of arrhythmia and the prevention of sudden cardiac death.

Keywords: Connexin 43, carboxyl terminus, gap junction, ZO-1, intercalated disk

1. Introduction

Gap junctions (GJ)s provide for cytoplasmic continuity between cells, permitting direct intercellular transmission of ions, second messengers, and other molecules less than 1kDa [1-3]. Most mammalian cells exhibit abundant GJ contacts with neighboring cells, providing the basis for integrated networks within tissues. The propagation of electrical activation in the heart is one of the more widely known examples of tissue integration by an intercellular network of GJs. That GJs have additional roles in tissue homeostasis and dynamics is suggested by the near ubiquitous presence of GJs between excitable and non-excitable cells in vertebrates.

The channels that are responsible for electrical coupling of cardiomyocytes are composed of proteins encoded by the connexin gene family [4]. Connexins comprise four transmembrane spanning motifs, with both amino-(NT) and carboxyl-(CT) termini located on the cytoplasmic side of the plasma membrane. Connexin subunits form hexameric connexon oligmers, which traffic to the membrane, and dock with connexons from adjacent cells to form functional intercellular channels [5] reviewed in [6]. Aggregation of docked connexons results in the formation of the membrane organelle responsible for cell-to-cell communication - the GJ plaque (Figure 1).

Figure 1. The Gap Junction and Perinexus.

Figure 1

Top) Schematic of gap junction plaque with adjacent perinexus regulating hemichannel addition at the plaque edge. Bottom) Immunoconfocal view of rat epicardial cells labeled for Cx43 in green and Cx43/ZO-1 interaction in red as labeled in situ by the Duolink protein-protein interaction assay. Note that the extent of red signal in the perinexus exceeds that of the green immunolabeled gap junction. For more information see [95].

The CT sequence is the site of greatest amino acid (aa) variability between connexin isoforms [7] and thought to be the main regulatory domain of most connexins. The connexin CT has roles in GJ assembly and channel function, via numerous post-translational modifications and protein-protein interactions (Figure 2). These assignments within the CT are also critical to governance of the trafficking, size, localization and turnover of GJs, as well as the level of intercellular coupling mediated by the plaque. Disruption in the distribution or extent of GJs is characteristic of various disease states, including slow healing wounds in the skin, epileptic seizures in the brain and arrhythmias in the heart. Mutations in Cx43 have been linked to the inherited human disease oculodentodigital dysplasia (ODDD), though interestingly relative to other domains of Cx43 there is a low incidence mutations in the CT[8].

Figure 2. The Interacting Proteins of the Cx43 Carboxyl-Terminus (CT).

Figure 2

Amino acid sequence of human Cx43 from CT residues 234 to 382 labeled with sites of protein-protein interaction. NMR structure of Cx43 was acquired by Sorgen and others[65] and obtained from the World Wide Protein Data Bank (PDB ID: 1R5S).

Connexin 43 (Cx43) is a GJ protein commonly found in excitable tissues in mammals, including humans [9-11]. The ventricular muscle of the heart is particularly enriched for Cx43, where it localizes preferentially to cardiomyocyte longitudinal termini -i.e. the intercalated disk. Mislocalization of GJs from polarized distributions at the intercalated disk and remodeling of Cx43 to the lateral surfaces of cardiomyocytes has been noted in nearly every cardiac disease state, and is correlated with the propensity of arrhythmia generation in the affected tissues [12-18].

There are numerous informative and detailed surveys of the literature on Cx43 remodeling and cardiac arrhythmic disease [9, 19-21]. Reviews on the structure, molecular biology and function of the intercalated disk in its role of providing a platform for electromechanical coupling between cardiomyocytes are also available for interested readers [22-26].

Here, a perspective on regulation of GJ plaque size and localization in the plasma membrane (sarcolemma) of ventricular working myocardial cells will be provided. Emphasis in the review will be placed on the role of the Cx43 CT, and in particular, the Cx43 CT-interacting protein Zonula Occludens-1 (ZO-1). Where relevant, other cell types, connexins, junctional molecules and interacting proteins will be broached. The mechanisms determining normal GJ order in the healthy myocardium and pathological disorder in the diseased heart are a dual practical focus of this review. A working hypothesis is that the establishment of orderly arrangements of GJ organization in the sarcolemma is key to stable propagation of electrical activation in myocardial tissues. Breakdown of this order can lead to cardiac conduction disturbance, arrhythmia and sudden death.

2. Historical Background

The initial evidence that GJs in the mammalian heart demonstrated high levels of spatial order came from the electron microscope. Thin section electron microscopy localized large GJs at intercalated disks, particularly at lateral segments of apposed sarcolemma within disks [23, 26-29]. Based on ultrastructural data, there were debates in the literature as to the size of the largest GJs at the intercalated disk [23, 30], the resolution of which would await the advent of other approaches, including immuno-confocal microscopy. Small, interplicate GJs were also noted dispersed within the fascia adherens, the extended adhesive junctional complexes responsible for stabilizing mechanical interactions between cardiomyocytes [31, 32].

Pioneering freeze fracture electron microscopy provided high spatial and temporal resolution views of GJ aggregates in beating myocardial tissues[33, 34]. Aggregates of particles within GJs thought to represent individual connexons, were revealed to occur in closely packed arrays that were separated by particle-free aisles. Freeze fracture studies of embryonic, neonatal and perinatal ventricle also indicated the occurrence of developmental increases in the size of GJ [35, 36], suggestive that the differentiation of GJ organization in the developing heart was in itself a progressive and regulated process.

The identification of connexin gene sequences in the late 1980s[37], including that encoding Cx43 [38] led to the advent of highly specific antibodies. These antibodies, coupled with powerful new light microscopic techniques, such as laser scanning confocal microscopy (LSCM), began to provide new views of the 3D order of GJs in the adult and developing heart[39-42].

Confirmation of the preferential localization of presumptive GJs at intercalated disks using immunohistochemistry and Cx43 specific antibodies was first provided by Beyer and co-workers in 1989[39]. Shortly thereafter it was demonstrated that Cx43 GJ substructure showed further organization within the disk. Three-dimensional reconstructions of en-face views of the termini of rat ventricular cardiomyocytes from LSCM optical sections revealed circumferential distributions of large immunolabeled GJs [40, 44]. This spatial configuration was subsequently confirmed to occur in the ventricles of mice, humans and other mammals [40-45]. Quantitative comparison of plaque size from immunofluorescent labeling and freeze fracture electron microscopy were among approaches used to validate that myocardial GJs were being correctly identified and measured by the LSCM [41, 46].

The distinct population of large GJs at the intercalated disk circumference did not occur in the embryonic ventricle, or indeed become fully manifest until quite late into postnatal development [40, 42, 43, 47]. Several immunofluorescence studies indicated that distributions of Cx43 GJs in the neonate were uniformly organized around cardiomyocytes. But with subsequent maturational growth of the heart there was a loss of these lateral contacts between cardiomyocytes and progressive increases in the relative abundance of Cx43 immunolabeling at intercalated disks.

The culmination of this developmental process resulted in 60 % or greater of immunodetectable Cx43 at the disk in the ventricle of most mammals[32]. It is thought that this polarization of GJs to the ends of cardiomyocytes contributes to changes in conduction of the ventricle from isotropic to more directional propagation of electrical impulse along the longitudinal axis of myofibers [48, 49]. This being said, mathematical modeling indicates that the postnatal increase in cardiomyocyte size is probably a larger factor in the emergence of directional propagation of cardiac action potentials than GJ distribution [50].

In 1991, studies of infarcted ventricular myocardium in humans [51] and dogs[52] provided the first indication that the pattern of Cx43 GJs is profoundly altered in a narrow zone of surviving muscle bordering the necrotic injury or scar - the infarct border zone. Since publication of these results, a number of groups have confirmed the close linkage between GJ remodeling in the infarct border zone and the propensity of ventricles to develop reentrant arrhythmias [15, 31, 53-59]. Typical of infarct border zone remodeling is a loss of GJs from the intercalated disk and redistribution of Cx43 to lateral domains of sarcolemma: an arrangement reminiscent of the developing ventricle.

With the parallels between disease and development being raised, an important distinction should be made. Cx43 shows homogenously lateralized distributions between cells throughout the myocardial tissues of the immature ventricle. By contrast, the lateralization of Cx43 at the border zone of healed coronary infarcts in an adult heart is an isolated and striking non-uniformity. Just a few cell layers away from the border zone, Cx43 GJs can be seen to assume normal distributions at intercalated disks. The remodeling of GJs in the infarct border zone to lateralized distributions occurs relatively rapidly, becoming notable within 24 hours[60]. The extent to which Cx43 along myocytes sides exists as GJs or connexons, or indeed whether lateralized Cx43 has channel function remains unclear.

The GJ is not the only element of the intercalated disk to show progressive change over postnatal ventricular development. Adherens junctions and desmosomal mechanical junctions also change from lateral, side-by-side contacts between cardiomyocytes in the immature ventricle to preferential accumulation at disks in the adult [47, 61]. The accumulation of GJs to the intercalated disk notably lags behind mechanical junctions, which assume mature localization at cardiomyocyte termini significantly earlier during postnatal development[61-63]. The important concept that intercalated disk assembly is a sequential process will be discussed further in relation to differing hypotheses regarding how GJs may be targeted to the intercalated disk.

3. Structure and Function of the Cx43 Carboxyl-Terminus

The main regulatory locus of Cx43 is the CT. While the CT sequence distal to amino acid 258 appears dispensable for polarization to cardiomyocyte termini [64], the Cx43 CT is required for proper organization in the intercalated disk, as well as regulating the channel responses to various chemical stimuli. Sorgen and others reported that the CT of Cx43 is predominantly a random coil composed of 132 amino acids containing two helical domains and extending from the cytoplasmic face of the membrane-docked channel [65] and Figures 1 and 2. This group further demonstrated that binding partners of the Cx43 CT could significantly alter its secondary structure. Importantly, these structural alterations were observed at sites remote to the region of binding. This suggested that the Cx43 CT was not simply a site of protein-protein interaction, but a platform through which connexin-binding partners could interact and signal using this sequence as an intermediary.

Proteins interacting with the Cx43 CT have been previously reviewed by Giepmans in 2004 [66] and Herve et al. in 2007 [2]. An updated survey of proteins interacting with the Cx43 CT is presented in Figure 2 and Table 1. The Cx43 CT-protein interactions of particular note with respect to GJ channel assembly, regulation and trafficking are given particular attention. Figure 2 shows the Cx43 CT sequence from amino acid 234 to 382, with the known sites of interaction of several Cx43 binding partners labeled. The consequence of CT-protein interaction is additionally noted.

Table 1.

Proteins Interacting with the Cx43 Carboxyl-Terminus.

Interacting
protein
Amino acids Consequence Reference
β -tubulin 234-262 Trafficking and TGF̄β
signaling
[140, 141, 144]
CaMKII S244 Phosphorylation, UKS [175]
Src SH2 Y247 Phosphorylation (disrupts
communication)
[96, 176]
MAPK/CaMKII S255 Phosphorylation, UKS [175, 177, 178]
CIP85 P253-P256 (*rat Cx43) Induce GJ turnover [179]
CaMKII S257 Phosphorylation, UKS [175]
PKC S262 Phosphorylation, (reduces
DNA synthesis)
[180], [170]
Src SH2 Y265 Phosphorylation (disrupts
communication)
[181, 182]
V-Src SH3 274-287 SRC-Cx43 putative binding [183]
MAPK Y274 S279 S282 Phosphorylation (disrupts
communication)
[178]
Nedd4 282-286 ubiquitination [184]
CaMKII S296, S297, S306, S314, Phosphorylation, UKS [175]
CK1, CaMKII S325 S328, S330 Phosphorylation
Enhanced GJ assembly
[175, 185, 186]
MMP-7 A357-I358 Potentially generate free
carboxy-terminal peptides?
[173]
PKA, CaMKII S364 Enhances channel assembly [165, 175]
PKA CaMKII S365 Phosphorylation
Gatekeeper:must be
dephosphorylated for PKC to
phosphorylate S368
[175, 187]
PKC PKA S368 Phosphorylation (disrupts
communication);
Stabilizes GJs at Intercalated
disks;
Decreased assembly during
S-phase
[163, 169, 188, 189]
PKA, CaMKII S369 Increased Channel
formation/activity
[189]
PKC S372 Phosphorylation, UKS [190]
PKA, 14-3-3 S373 Phosphorylation required for
14-3-3 binding
[191]
MMP-7 D379-L380 Potentially cleave PDZ
binding domain?
[173]
ZO-1-PDZ2 379-382 Reduced GJ size [72, 100]
Proteins with
uncharacterized
binding sites
NOV/CCN3
between 257-374 Reduced cell proliferation [192]
Caveolin-1/-2/-3 binds whole C-terminus but
not D244 mutant
Indirect regulation of channel
activity
[193, 194]
Drebrin Unknown Cytoskeletal anchor [195]

Note: While some studies have clearly identified certain serines and tyrosines as targets of specific kinases in vivo, it is apparent that multiple kinases can phosphorylate the same serine residue under different conditions. In vivo it is difficult to determine if a specific kinase is mediating the phosphorylation event directly or via activation of another kinase. UKS = unknown significance.

Table 1 lists the Cx43 CT binding partners shown in figure 2, with additional information on site and functional consequences of interaction. The table is organized in order of interaction site from juxta-membrane to the distal terminus. Figure 2 and table 1 will serve as a reference for the discussion of the CT binding partners that follows. In the following sections we will address the role of these interactions in regulation of Cx43 GJ organization. Extra attention will be given to ZO-1, one of the most widely studied binding partners on the Cx43 CT.

In addition to participating in intermolecular interactions, the Cx43 CT has been reported to dimerize under acidic conditions in vitro [67]. This ability to dimerize suggests potential for intramolecular interactions in which a Cx43 CT may be capable of interacting with itself or another domain of Cx43. The CT was proposed to regulate GJ channels via a “ball and chain” mechanism [68] and intramolecular interactions of this type appear to be involved in pH sensitive gating[69] [70]. Mimetic peptides based on the CT have also been shown to interact with the cytoplasmic loop (CL) domain of Cx43 and effect hemichannel function in a Xenopus oocyte model [71]. Some of the ODDD mutations may inhibit the Cx43 CT-CL intramolecular interaction[8]. Presently, it is unclear whether Cx43 CT intramolecular interactions contribute to GJ size control or cellular distribution.

4. The Role of ZO-1 Interaction with Cx43 in Regulation of GJ Organization

A key step in understanding control of GJ organization in the heart came with the identification of the interaction of Cx43 with ZO-1 [72, 73]. While there are numerous Cx43 CT binding partners, many of which may be involved in regulating GJ size and cellular distribution, few other interacting proteins have been studied as extensively as ZO-1[11, 74-79].

The second Postsynaptic density/Disks large/ZO-1 (PDZ) domain of ZO-1 interacts directly with the distal extremity of the Cx43 CT [72, 73]. Originally discovered in association with tight junctions [80], ZO-1 is a member of the membrane-associated guanylate kinase (MAGUK) family of proteins that function in protein targeting, signal transduction, and determination of cell polarity[81] [82]. MAGUKs, such as ZO-1, synaptic protein PSD95 and the Drosophila tumor suppressor discs large (dlg) are characterized by amino-terminal protein-protein binding motifs, including one or more PDZ domains.

Initial immunoprecipitation and yeast 2-hybrid studies showed that a 4 amino acid PDZ-binding ligand at the extreme CT of Cx43 was necessary for interaction with the second of three PDZ domains on ZO-1 [72, 73]. Ongoing work has indicated that the ZO-1-binding domain comprises up to the CT-most 19 amino acids of Cx43 and this sequence likely interacts with a ZO-1 dimer [83, 84]. Structural studies have demonstrated roles for ZO-1 binding in regulating Cx43 CT structural confirmation and signaling [65]. It is also now known that numerous connexins interact with ZO-1 via a CT PDZ-binding sequence similar to that of Cx43 [83, 85-89]. A second MAGUK, ZO-2, has been reported to interact with Cx43 in a cell-cycle dependent manner [90].

4.1 ZO-1 Interaction with Cx43 and GJ Internalization

When the interaction between Cx43 and ZO-1 was first identified in the late 1990s, it was proposed that its function was to stabilize GJs in the plasma membrane. However, subsequent findings by a number of groups, suggested that the view of ZO-1 as a passive scaffold at the GJ was inadequate. Notably, ZO-1 has an actin-binding domain at its CT, enabling linkage of GJs to cytoskeletal dynamics[91]. Consistent with this ability to mediate linkage to actin, ongoing reports indicate that the interaction between Cx43 and ZO-1 encompasses active functions, including in GJ remodeling [74, 92-103].

In the early 2000s, we published results showing that Cx43/ZO-1 interactions may be involved in the internalization of GJs [77, 92]. The approach was to demonstrate that ZO-1 and Cx43 association showed an acute and conspicuous increase following collagenase dispersal of cardiomyocytes from intact ventricles. Severs and others demonstrated in an earlier study that cardiomyocyte dissociation prompted mass endocytosis of GJs, forming a stable population of vesicular annular junctions within the cytoplasm [104]. We hypothesized in the 2002 study that a function of ZO-1 may have been to link Cx43 to actin-based motors, and that this contributed to the rapid internalization of GJs when cardiomyocytes were disaggregated from intact ventricular muscle by collagenase.

Integral to this discussion is the parallel story of Cx43/Src interaction. Src was identified as a potent inhibitor of Cx43 GJ communication in 1990[105, 106]. Tyrosine phosphorylation was found to be a hallmark of Src-mediated Cx43 inhibition, and it was demonstrated that Src phosphorylates the Cx43 CT on residues Y247 and Y265[107]. Saliently, Toyofuku and co-workers later determined that phosphorylation mediated by Src played a role in GJ communication deficits in the cardiomyopathic hamster heart[108]. The same group subsequently demonstrated that Src binding and phosphorylation are mutually exclusive to ZO-1 interaction with Cx43 [97]. In a final twist, recent data has indicated that Src decreased interaction between Cx43 and ZO-1 by competitively binding to ZO-1 in the infarct border zone of a canine coronary ligation model [109], with an associated loss of Cx43 from the intercalated disk.

The above findings on Cx43, Src and ZO-1 indicated the occurrence of complex interplay between the three proteins. Studies by Pointis, Segretain and co-workers have provided further insight into this set of interactions. [74].The aforementioned laboratories undertook their work on association between Cx43 and ZO-1 during GJ internalization in Sertoli and other cell lines[74]. In 2008, this group made the observation that during a chemically induced GJ endocytosis, ZO-1 dissociated from one side of the GJ, but remained associated with the other side of the plaque[98]. This loss of ZO-1 from a cytoplasmic face of the GJ appeared to occur in response to competition for binding to the Cx43 CT by Src. Asymmetric maintenance of ZO-1 on the other plaque face culminated in focal concentration of the actin-binding protein within the interior of endocytozed GJ vesicles. Interestingly, the co-localization pattern reported by this group, wherein ZO-1 signals appeared to be contained with larger Cx43 particles, was similar to that located at annular junctions in enzymatically-dispersed cardiomyocytes[92]. More recent data from the French group shows that internalization of annular junctions is dependent on the protein dynamin2, which they demonstrate to interact with Cx43 during plaque endocytosis[110].

Evidence that ZO-1 interaction with Cx43 is a mechanistically necessary participant in GJ endocytosis came from papers examining the response of Cx43 to inflammatory mediators. In 2007, van Zeijl, et al. reported that ZO-1 was required for the down-regulation of GJ intercellular communication observed in response to endothelin-1 and thrombin-like agonists in Cx43-expressing Rat-1 cells[111]. In a subsequent paper from the Falk group, using endothelial and HeLa cells, it was demonstrated that thrombin and endothelin-1 inhibited intercellular coupling via inducing rapid and acute internalization of GJs[103]. Importantly, these authors showed that in the absence of competent ZO-1 interaction with Cx43, internalization of GJs and annular junction formation in response to treatment with the inflammatory agonists did not occur.

Taken together the evidence from these studies suggested that ZO-1 interaction with Cx43 is necessary for certain types of GJ internalization processes. However, it is widely recognized that certain GJ endocytotic events may not require ZO-1 [112, 113]. In one instance, classical clathrin-based mechanisms efficiently operate to mediate internalization of GJs from the plasma membrane [112]. Furthermore, HeLa cells expressing Cx43 rendered incompetent to interact with ZO-1 by a CT GFP tag form large junctions without a significant change in Cx43 half-life, suggesting that in this context loss of the PDZ2 interaction does not increase GJ turnover [100, 101]. In ongoing work, our group finds that ZO-1 binding to Cx43 CT can also direct actin-based force in the plane of the plasma membrane, mediating the fusion and pulling apart of GJ plaques [114].

Significant questions remain regarding the molecular motors that remodel Cx43 GJs. For example, in internalization events involving ZO-1, it remains unknown how actin forces are imparted on the GJ. Does it involve pulling or pushing of double-membrane GJs from the plasma membranes of adjoined cells? What are the other linkage and force generating constituents of these motors? Does the molecular machinery that operates constitutively during GJ turnover differ from that which operates during acute events such as in response to ischemia, mechanical destabilization of cell-cell contacts or inflammatory cytokines?

4.2 ZO-1 Interaction with Cx43 and Regulation of GJ Size: In Vitro Studies

Insight into a second role for ZO-1 at the GJ in the control of plaque size was initiated from studies of Cx43-ZO-1 co-localization. In a 2002 paper we noted that the level of ZO-1 association with Cx43 in intact ventricular myocardium was modest, with some GJ plaques at intercalated disks demonstrating no detectable ZO-1[92]. Immuno-confocal microscopy of cultured neonatal rat ventricular cardiomyocytes provided greater clarity of the association between the two proteins than could be achieved in tissue sections of cardiac muscle [115]. Confocal optical sectioning of monolayers of cardiomyocytes revealed that ZO-1 preferentially localized at the edge of GJ plaques.

Having observed this peripheral accumulation of ZO-1 at the GJ edge, the question arose as to what would happen if this association was inhibited. We determined that treatment of neonatal rat cardiomyocytes with a peptide mimetic of the Cx43 ZO-1-binding domain decreased co-localization of ZO-1 at the edge of GJs and increased plaque size[101]. This result was interpreted as implying that ZO-1 interaction with Cx43 at the GJ edge may be acting to constrain the extent of GJ channel aggregation.

Consistent with a role in inhibiting plaque size growth, it was also noted that abnormally large GJs formed between HeLa cells heterologously expressing a Cx43-GFP construct incompetent to interact with ZO-1[79] These large junctions were reduced to more normal sizes when the cells were transduced by adenovirus expressing wildtype Cx43[100, 101]. Subsequent work by Andrew Hunter in the lab established that the Cx43-GFP GJs also had deficits of ZO-1 at the plaque edge, and that the reduction in plaque size caused by co-expression of wildtype Cx43 was accompanied by recruitment of ZO-1 to the edge of GJs [101].

It was also determined that the GJ size increase prompted by the Cx43 CT peptide, did so via inhibiting the interaction of the second PDZ domain of ZO-1 with Cx43[101]. The reduction in interaction mediated by the PDZ2-binding peptide in turn resulted in a shift within the total pool of Cx43 protein in cells from detergent-soluble to GJ-associated detergent-insoluble pools. Taken together, the results indicated that ZO-1 controlled the rate at which Cx43 transitioned from non-junctional to junctional pools of the protein [101].

In ongoing studies the question of how and where ZO-1 may perform this function has been addressed[95]. The approach taken was to use in situ protein-protein interaction assays to further define the relationship between Cx43 and ZO-1 in a HeLa cell model[95]. Using the Duo-link assay to fluorescently amplify Cx43 ZO-1 association, a previously unknown zone of interaction proximal to the GJ edge was identified. This sector, which we termed the Perinexus, appears to be the primary site at which control of the transition of non-junctional connexons into the junctional plaque occurs.

Based on the data reported, it was proposed that the mechanism of ZO-1-mediated control of GJ size is via governing the connexon hemichannel to GJ channel transition from the perinexus [95]. The perinexal scaffold is envisaged as a sieve-like network containing ZO-1 that extends from the GJ edge. This network is part of a mechanism that regulates, but does not prevent, the flow of connexons from surrounding membrane available to dock with connexons from the apposed cell (Figure 1). In this manner, it was hypothesized that ZO-1 regulates a balance between connexon-mediated membrane permeance and GJ intercellular communication. Indeed, it was demonstrated that the level of Cx43/ZO-1 interaction concomitantly and inversely affected hemichannel-mediated dye uptake and GJ-mediated cell-to-cell dye transfer [95].

The model of connexon flow from the perinexus is consistent with the widely held view that GJs are assembled from the outside edge of the plaque inwards [116, 117]. This was most notably illustrated in the elegant study of Gaietta et al., in which pulse-chase labeling of tetracysteine-tagged Cx43 with biarsenical compounds showed that Cx43 accreted to the edge of GJs [117]. The new insight provided by the Duolink assay is that the recruitment process draws from a pool of non-junctional connexons located proximal to the GJ edge in the perinexus [95]. .

The study also provided unexpected new data on the association between actin, ZO-1 and Cx43 at GJs. Co-immunoprecipitation demonstrated that ZO-1 interacted with both detergent soluble and insoluble Cx43-containing connexons, but only those connexons docked in the detergent insoluble fraction were also complexed with actin [95]. These data correlated well with immunofluorescenent data showing preferential overlap of actin signal with Cx43/ZO-1 interaction within the GJ as compared to the perinexus. At present, how this partitioning of Cx43 interaction with the actin cytoskeleton is regulated is unclear. Nonetheless, the linkage of actin to connexons in the GJs, but less avid association with connexons in the perinexus, may provide one explanation as to how ZO-1 appears to have dual functions: One in the transmission of force to the GJ via tethering to the actin cytoskeleton and the other in controlling the flow of connexons at the plaque perimeter.

One key question for ongoing work is the specifics of how perinexal ZO-1 acts to regulate connexon recruitment. Another outstanding issue is whether other connexins that interact with ZO-1, such as Cx40 [89] and Cx45 [86, 118], exhibit mechanisms of GJ size control similar to those of Cx43. There are also interesting related questions on whether heteromeric and heterotypic partnering of connexins affect control of GJ aggregation. In an example from the literature, coexpression of Cx43 with Cx45 in HeLa cells has been reported to decrease GJ size[119]. Characterization of CT interactions in cell models expressing and/or co-expressing different connexin isoforms may provide further insight into the role of perinexus in regulating GJ organization.

4.3 ZO-1 Interaction with Cx43 and Regulation of GJ Size: In Vivo Studies

There is growing evidence that ZO-1 contributes to regulation of GJ size in vivo. Delmar and co-workers determined that expression of a CT-truncated Cx43 incompetent to interact directly with ZO-1 resulted in formation of abnormally large GJs in lateral sarcolemma adjacent to the intercalated disk [64]. In patients with certain types of cardiomyopathy, increased ZO-1 interaction with Cx43 was reported to prompt attendant decreases in the abundance and size of GJs [11].

Recent studies from our lab have also elucidated how patterns of ZO-1 interaction with Cx43 may generate the characteristically large GJs ringing the intercalated disk. GJs at the periphery of the disks in the rat ventricle were found to exhibit greater ZO-1 colocalization at their inward, as compared to their outward facing edges[120]. Disruption of this asymmetric co-localization by expression of a ZO-1 truncation mutant resulted in loss of size distinction between GJs at the peripheral edge of the intercalated disk versus those in the disk interior.

Together with in vitro studies [95, 101], it was proposed that the larger size of GJs at the intercalated disk periphery resulted from preferential accretion of connexons at the outward facing edge (i.e., ZO-1-lacking) of GJ plaques. Therefore, when connexon recruitment is not regulated by the presence of a perinexal ZO-1 network, such as in the case of GJs at the intercalated disk periphery[120] or between Cx43-GFP expressing HeLa cells [100], plaques tend to grow larger.

4.4 ZO-1 Interaction with Cx43: GJ Size and Cx43 Level in Cardiac Disease

While spatial remodeling of GJs is clearly associated with arrhythmic disease, the functional significance of changes in GJ size to pathologic disturbance of conduction in the heart is less clear. Maass and co-workers demonstrated that an increase in GJ size from CT truncation of Cx43 was correlated with a propensity of acutely injured hearts to develop inducible arrhythmias[121]. Whether this effect is due to changes in GJ size or other phenomena associated with CT truncation - such as changes in junctional conductance - remain unclear. Conversely, Nick Severs group showed a decrease in GJ size in ischemic and dilated cardiomyopathies [11]. Interestingly, in some forms of severe epilepsy in humans, decreases in Cx43 GJ size have been reported to accompany an increase in overall GJ area and expression [122]. As is the case with cardiac arrhythmia, epilepsy is characterized by the propagation of chaotic electrical activity.

In 2007, Yamada and co-workers found that ZO-1 and Cx43 levels were reduced concomitant with a reduction in co-localization between these two proteins at intercalated disks in humans at end-stage heart failure [123]. Interestingly, although Cx43 levels were diminished, the percentage of disks positive for Cx43 was not different from that of controls. In a short report, Kostin described a similar diminution in ZO-1 in heart failure patients, but noted that Cx43 was decreased in intercalated disks in which ZO-1 was also reduced [124]. Related analyses were performed in 2007 on a rat model of right ventricular hypertrophy. This study found increased GJ internalization and some lateralization of Cx43, but in apparent contrast with the aforementioned papers, no change in total Cx43 levels or the co-distribution of Cx43/ZO-1 in diseased hearts [125].

As touched upon in the first paragraph of this section, Severs’ group in London performed an in-depth study on ventricular tissues sampled from human patients in end-stage heart failure [11]. In line with earlier results, these workers found that the abundance of Cx43 was significantly reduced in patients compared to controls subjects. Moreover, ZO-1 levels were increased at the intercalated disks in these individuals, and ZO-1 levels showed a negative correlation with Cx43 levels at disks. Furthermore, co-localization and co-immunoprecipitation assays indicated that interaction between Cx43 and ZO-1 was increased in these heart failure patients.

Recent data from the laboratories of Sorgen and Duffy on the border zone of infarcted myocardium in a canine model of coronary occlusion found that reductions in Cx43 level were associated with decreases in Cx43/ZO-1 binding [109]. A novel interaction between activated Src and ZO-1 showed a complementary increase in infarct border zone tissues. Finally, in a report published in 2010, a rat model of pressure-overload hypertrophy over an 8-week period was described in which Cx43 levels initially increased, then declined back to control levels by 8 weeks[126]. A similar variation over time was observed for Cx43 phosphorylated at serine 368 on its primary sequence. ZO-1 co-immunoprecipitation with Cx43 remained unchanged 2 weeks after pressure overload, but declined slightly at 8 weeks.

In sum, the literature on relationships between Cx43/ZO-1 interaction, GJ size, and Cx43 level in diseased myocardium is not amenable to simple interpretation. Further reports of Cx43 remodeling in the diseased ventricle include those by Goldfine et al. and Benes et al. [127, 128]. The different conclusions reached by different groups may be grounded in part in the varying models that have been used. These include two rodent models of hypertrophy, a canine ischemia model, and samples from human patients at various stages of heart failure with differing disease etiologies. Lining up assay of specific microscopy and biochemically based end-points with specific cardiac diseases and disease stages may also be key considerations in uncovering mechanism. Further work in this important area is evidently warranted.

5. GJ Targeting to the Intercalated Disk

How GJs come to be targeted to and retained in the intercalated disk is presently unknown. This being said, two main hypotheses have been proposed to explain how GJs assume polarized distributions at cardiomyocyte termini. The first hypothesis posits that polarization results from intercalated disks having lower rates of loss or turnover of GJs than other parts of the sarcolemma [61, 63, 129]. The second hypothesis proposes that order is generated by direct trafficking of Cx43 to the intercalated disk via cytoskeletal pathways - most notably via Cx43 CT interaction with microtubules. The two propositions are not mutually exclusive and indeed, when eventually understood, the actual mechanism may embody aspects of both.

Axiomatic to the first hypothesis, is that Ca2+-dependent adhesion junctions (adherens junctions and desmosomes) provide mechanically stable domains of sarcolemma, within which GJs are less rapidly turned over. The intercalated disk contains high concentrations of adherens junctions and desmosomes. As such, in this model GJs would tend to be preserved in or near disks in the active tissues of the beating heart. Conversely, GJs in lateral domains of sarcolemma would be mechanically destabilized at higher rates, being more susceptible to shearing stress at lateral interfaces between contracting muscle cells.

The GJ turnover model was first proposed based on descriptive studies of postnatal maturation of the rat and dog heart[61]. During cardiac maturational growth it was observed that the distribution of GJs displayed a transient dissociation from those regions of sarcolemma exhibiting the greatest concentrations of cell-cell adhesion junctions. This lack of GJ-mechanical junction association resulted from GJs accruing from lateral sarcolemma to the intercalated disk more slowly than the bulk of adherens junctions and desmosomal contacts between cells (Figure 3). In rats, the majority of both classes of adhesion junctions (i.e., >60%) become polarized to the intercalated disk between 10 and 20 days following birth [61]. By contrast, the rodent is full grown at 90 postnatal days before the majority of Cx43 is lost from myocyte sides and immunolabeling in the disk matches that of adherens or desmosomal junctions.

Figure 3. Two hypotheses for Gap Junction Recruitment to Intercalated Disks.

Figure 3

1) Following birth, mechanical junctions in the ventricle, including adherens junctions (red), undergo rapid polarization to intercalated disks, while the emergence of polarized distributions of Cx43 GJs (green) from lateral sarcolemma progresses on a slower time course. In Angst et al. [61]. it was proposed that this pattern emerges as GJs tend to be internalized (open green circles) from lateral domains of sarcolemma during postnatal maturation, whereas those GJs co-located with adherens junctions at disks are more liable to be retained in the membrane. 2) Microtubules traffic connexon hemmichannels directly to the Adherens junctions at the intercalated disk for direct assimilation into GJ plaques[142].

Peters et al., found that humans display a similarly extended time course of intercalated disk formation, with lateralized GJs still apparent in children up to 6 years of age [43]. The multi-year time course for establishing polarized distributions of GJs is orders of magnitude slower than GJ trafficking (<5 minutes) kinetics of Cx43. Additionally, these data suggest that over extended periods of time (e.g., such as that occurring over mammalian postnatal development), small deviations in turnover of plaques between disk and lateral sarcolemma could generate large differences in GJ abundance between these domains of membrane.

Electron microscopy of ventricle from normal adult mammals has indicated that internalized or annular GJs are extremely rare [23, 125]. However, a conspicuous increase in the frequency of annular GJs has been described from ultrastructural studies of immature human[43], rabbit [130], and dog [131] ventricle. Importantly, Mühlfeld and Richter found that annular GJ numbers increased sharply between birth, 5 and 14 postnatal days in the rat [132]. The timing of this postnatal uptick in internalization would be consistent with GJ stability being dependent on proximity to mechanical junctions, as it coincides with the developmental phase in this species within which the sarcolemmal distribution of adhesion junctions shows greatest dissociation from that of GJs [61].

Work in vitro and in vivo provides further support for the dependence of GJ spatial distribution on that of adhesion junctions between ventricular myocardial cells. In a study using aligned cultures of neonatal rat ventricular cardiomyocytes, Matsuda et al. recapitulated the maturational dissociation of GJs and adherens junctions, showing that polarization of Cx43 to cell termini was preceded by and required the localization of N-cadherin at intercalated disks [133]. Consistent with biomechanics having a determinative role in the differentiation of the intercalated disk, levels of expression and initial localization of N-cadherin to cardiomyocyte termini have been reported to be dependent on the configuration of active mechanical stress and strain [134-138].

Direct evidence of the requirement for stable cell-cell adhesion junctions in GJ stability has been provided in experiments in which a dominant negative N-cadherin was expressed in cultured cardiomyocytes. This mutant N-cadherin prompted loss of Cx43 from the sarcolemma and disruption of GJ contacts between cardiomyocyte pairs [139]. In line with these in vitro studies, Luo and Radice showed that cardiac-specific knockout of the N-cadherin gene in transgenic mice resulted in reductions of Cx43 GJs [140]. Our group has recently reported that expression of a dominant negative disruptor of adherens junctions in the postnatal rat ventricle leads to increases in lateralized GJs and complementary decreases in Cx43 at intercalated disks [120] -results consistent with adhesion junctions being necessary for establishing and maintaining normal GJ organization.

In the second hypothesis, targeting of Cx43 GJs to the intercalated disk is proposed to involve trafficking via microtubules. Falk, Giepmans and colleagues demonstrated that the CT of Cx43 could directly interact with β-tubulin and based on this interaction, it was proposed that microtubules trafficked connexins directly to the GJ [141]. Subsequent work provided evidence that Cx43 targeting to the plasma membrane involved delivery of connexons to N-cadherin complexes at the plus ends of microtubules [142]. In this respect it is notable that N-cadherin, β-catenin and Cx43 have been reported to co-immunoprecipitate as part of a multi-protein complex[143].

Intercalated disks are enriched with N-cadherin - the Ca2+ dependent cell adhesion molecule found in cardiac adherens junctions. It therefore follows that Cx43 should accumulate preferentially at sites of N-cadherin enrichment where microtubules tether, such as the intercalated disk. Further support for this trafficking-based model has come from more recent work from Shaw’s group in which it was shown that both Cx43 and the microtubule binding protein EB1 were decreased at the intercalated disk of patients with ischemic cardiomyopathy[144]. This observation was associated with a decreased localization of microtubules at adherens junctions and decreased Cx43 trafficking to the disk.

Experiments utilizing the expression of a truncated Cx43 raise some considerations for interpreting the role of microtubules in trafficking of connexins to GJs. It has been reported that amino acids 234 to 262 of the Cx43 CT contain the necessary and sufficient sequences to mediate interaction with β–tubulin [141, 145]. As outlined previously in this review, a CT-truncation of Cx43 has been expressed in transgenic mice [64]. This CT mutant was truncated from amino acid 258 and presumably able interact with microtubules normally, as it retained a consensus microtubule-binding domain. The Cx43 mutant protein accumulated at cardiomyocyte termini and formed functional GJs in this transgenic. However, the structure of intercalated disks of mice expressing the truncated Cx43 was disturbed, with GJs being located in domains of sarcolemma lateral to, but not within the body of the disk proper. Moreover, the GJs that formed in these domains were unusually large and showed abnormal associations with ZO-1, N-cadherin and desmosomal proteins.

Taken together, the data suggest that microtubule trafficking may be necessary for general localization of Cx43 at adherens junction-rich regions of sarcolemma. However, these cytoskeletal elements may not be sufficient to maintain Cx43 distribution at the intercalated disk long-term. Additional microtubule independent CT-interactions, or at least Cx43 CT-interactions distal to the presumptive microtubule-binding domain may be required for normal GJ-adherens junction configuration within the disk.

Moreover, consonant with the first hypothesis, numerous lines of evidence support the concept of a hierarchical dependence of GJ pattern and stability on the formation and distribution of adherens junctions. The role of this relationship in establishing the cellular pattern GJ connections may be particularly important in early postnatal life when adherens junctions and desmosomes undergo their rapid polarization to cardiomyocyte termini in response to changes in mechanical activity following birth. This hierarchy is probably also evolutionarily conserved, or at least been converged upon by other species, as cadherin junctions have been reported necessary for membrane retention of innexins, the genetically unrelated GJ proteins of invertebrates [146].

Proper localization of cardiac GJs in the physiologic state appears to be ultimately determined by Ca2+-dependent adhesion junctions, independent of whether GJ stability or targeting by microtubules individually or jointly determine the pattern. Subsequent regulation of GJs via protein-protein interaction and post-translational modification seem likely to be downstream of this requirement. A hypothetical model for how GJs assume localization at the intercalated disk that distils results from various studies is provided in Figure 3.

Questions remain as to if, how and when the different processes hypothesized in the model might act. Given that targeting GJs to the intercalated disk occurs over the entire course of postnatal development, it is conceivable that separate mechanisms may contribute at differing phases or time scales. There are also specific outstanding matters of interest. For example, the molecular mechanism responsible for the spike in GJ internalization during early postnatal life remains to be characterized. Cardiac-specific expression in mouse of a Cx43 mutant protein missing the consensus CT microtubule-binding domain could provide useful insight into the role of trafficking in the differentiation of the intercalated disk.

6. GJ Lateralization in Cardiac Disease

GJ order in the healthy myocardium and pathological disorder in the diseased ventricle appear to show similarities. As has already been broached, in disease GJs often assume lateral distributions that are independent of intercalated disks. The lateralization of Cx43 observed in many pathological scenarios is reminiscent of arrangements observed during maturational growth of the ventricle. This remodeling of GJ contacts in disease is thought responsible for heterogeneities in electrotonic coupling that generate tissue substrates for re-entrant arrhythmias.

In a further notable parallel to postnatal maturation, a number of studies of experimental and clinical cardiomyopathies have shown that Cx43 distribution dissociates from that of N-cadherin[147-151] or desmosomal proteins [152, 153]. In this pathological recapitulation of developmental patterns in the adult, cell-cell adhesion junctions appear to remain prominently localized at the intercalated disk, whereas Cx43 distributes to lateral sarcolemma (Figure 4). Also similar to the situation observed during postnatal growth, endocytozed GJs have been reported to demonstrate an increase in frequency in some of these pathologies [149, 150, 153].

Figure 4. Gap Junctions and Adhesion Junctions in Cardiac Development and Disease.

Figure 4

Immunolabeling of Cx43 and cell-cell adhesion junctions in ventricular myocardium from various reports are shown from normal (undiseased) adult, diseased adult, and normal postnatal ventricle demonstrate how arrangements of intercellular junctions seen in the still developing postnatal heart are recapitulated in adult disease. While these figures have not been modified, the color tables have been adjusted such that Cx43 is consistently indicated by green signal and Ca2+-dependent adhesion junctions are in red A. Normal Adult Human Left Ventricle from Smyth JW, Hong TT, Gao D, Vogan JM, Jensen BC, Fong TS, Simpson PC, Stainier DY, Chi NC, Shaw RM. Limited forward trafficking of connexin 43 reduces cell-cell coupling in stressed human and mouse myocardium. J Clin Invest, 120 (2010) 266-79 used by permision of the American Society for Clinical Investigation. B. S. Kostin, S. Dammer, S. Hein, W.P. Klovekorn, E.P. Bauer, J. Schaper, Connexin 43 expression and distribution in compensated and decompensated cardiac hypertrophy in patients with aortic stenosis, Cardiovasc Res, 62 (2004) 426-436 used by permision of Oxford University Press. C. Reproduced from Sepp R, Severs Nj, Gourdie Rg, Altered Patterns Of Cardiac Intercellular Junction Distribution In Hypertrophic Cardiomyopathy, Heart, 76 (1996) 412-417 with permission from BMJ publishing group. D. Reproduced from A. Nygren, M.L. Olson, K.Y. Chen, T. Emmett, G. Kargacin, Y. Shimoni, Propagation of the cardiac impulse in the diabetic rat heart: reduced conduction reserve, J Physiol, 580 (2007) 543-560. With permission from John, Wiley and Sons. E. Postnatal rat demonstrates uniform Cx43 distribution at the lateral membranes and the disk - taken from [61]. Note that in all cases of disease Cx43 is detected in lateral domains of sarcolemma. Scale bars A = 2.5 μm, B-E= 10 μm.

Mutations of desmosomal proteins have been identified in a number of rare arrhythmic cardiomyopathies including arrhythmic right ventricular cardiomyopathy, Carvajal syndrome and Naxos disease[154-157]. Disruption of mechanical junctions in these syndromes has been noted to be associated with a loss and lateralization of Cx43 in ventricular muscle[158, 159]. The fact that the normal order of the intercalated disk can also be disrupted by mutations in components of desmosomal junctions [160, 161], reinforces the notion of a hierarchical dependence of GJ stability on the normal function of mechanical contacts between cardiomyocytes.

The kinases and phosphatases responsible for regulating Cx43 CT phosphorylation have important roles in determining GJ organization in the sarcolemma in health and disease [97, 98, 162-166]. These Cx43-interacting proteins, and the phosphorylation events that they mediate, have potential in the development of anti-arrhythmia therapies[167, 168]. Phosphorylation of serine 372 and 373 destabilizes Cx43 CT binding to the ZO-1 PDZ2 domain[169], with presumptive downstream consequences for GJ size regulation. In another example with clinical implications, Fishman and colleagues identified a mechanistic link between casein kinase-phosphorylation of the Cx43 CT, and blunting of pathological GJ remodeling, in a pressure overload-induced model of cardiac hypertrophy[168].

Specific patterns of Cx43 phosphorylation appear to be associated with the conferring of resistance to pathological lateralization of GJs. Ek-Vitorin et al reported that in response to transient ischemia, upregulation of PKC-epsilon mediated phosphorylation of Cx43 CT at serine 368 resulted in retention of subpopulations of GJs at intercalated disks, even as total Cx43 underwent remodeling to lateral cardiomyocyte surfaces [170]. Similarly, Kardami and co-workers demonstrated that ischemic-preconditioning increased PKC-epsilon-mediated Cx43 phosphorylation of serines at positions 262 and 368, and inhibited Cx43 lateralization following an ischemic insult [171]. Studies that examined cell-cycle regulation of Cx43 have shown that during S-phase GJ assembly is decreased and is correlated with increased phosphorylation of serine 368 [164]. Phosphorylation may also have affects on intramolecular interaction between the Cx43 CT and CL [71] and the targeting of Cx43 to the intercalated disk[64, 168]. Maass et al., Circ Res., 2007).

The Cx43 CT-interacting kinase PKC-ε also seems to have promise as a therapeutic target. Dhein and colleagues determined that a short peptide, AAP10, inhibited Cx43 dephosphorylation at the PKC-ε consensus serine 368 and remodeling of GJs away from intercalated disks following induction of ischemia in a rabbit model[172]. Our laboratory has demonstrated that a peptide mimetic of the Cx43 CT can activate PKC-ε phosphorylation in vitro. Treatment of cardiac injuries with this peptide in a mouse model resulted in increases in the levels of Cx43 at intercalated disks and consistent with experiments in vitro, prompted serine 368 phosphorylation of Cx43 [173].

Interestingly, Matrix metalloproteinase-7 (MMP-7) has been shown capable of cleaving peptides from the extreme CT of Cx43[174] (Figure 2.) It is unclear at this point if this can occur physiologically, as MMP-7 is canonically extracellular whereas the Cx43 CT in cytoplasmic. Nonetheless, this interesting result points to the possibility that endogenous Cx43 may itself be a source of bioactive cleavage products. Given the data on the actions of synthetic peptides based on the Cx43 CT [95, 101, 173, 175], any endogenously generated peptides of related sequence would have the potential to regulate GJ size and configuration in the sarcolemma via modulation of PKC-e, ZO-1 and other interacting proteins.

11. Concluding Remarks

The Cx43 CT, through its protein-protein interactions and post-translational modifications, plays a direct role in regulating Cx43 GJ size, localization and sub-organization at the intercalated disk in the developing and adult ventricle. The proper arrangement of Cx43 GJs at the disk is critical to the maintenance of a stable cardiac rhythm and further understanding of Cx43 CT interactions could provide new paths for therapies that prevent arrhythmia and sudden cardiac death. Exciting future work would usefully focus on further elucidating the processes that result in perturbations of Cx43 localization and function in disease. By identifying the network of Cx43-binding partners, and understanding their dynamic patterns of interaction, a fuller understanding of the molecular determinants of pathological GJ remodeling will emerge.

Highlights.

  • >

    Gap junctions are key to electrical signaling in the heart

  • >

    Disruptions of gap junction organization in disease of the heart can lead to arrhythmia and sudden death

  • >

    Here, we review the current state of knowledge on the molecular processes regulating gap junction organization in the normal and diseased heart

  • >

    Understanding these molecular processes could provide new ways to treat heart disease.

  • >

    The evidence for a novel gap junctional domain - the Perinexus - is discussed

Acknowledgements

The support of NIH F30 HL095320 (PI:JAP), NIH T32 HL007260 (Trainee:JMR), NASA Pre-Doctoral Fellowship from the South Carolina Space Grant Consortium (Trainee:JMR), NIH grants HL082802 (PI:RGG) and 1R01DE019355-01 (PD: M.J. Yost), and AHA Grant-in-Aid 87651 (PI:RGG) are acknowledged with gratitude. The generous support of Dr Phil Saul MD (Chief, Pediatric Cardiology, MUSC) is also acknowledged with gratitude.

Abbreviations

aa

amino acid

CT

Carboxyl-Terminal

CL

Cytoplasmic loop

Cx43

Connexin 43

dlg

Disks Large

GJ

Gap Junction

LSCM

Laser Scanning Confocal Microscope

NT

Amino Terminal

MAGUK

Membrane-Associated Guanylate Kinase

ODDD

oculodentodigital dysplasia

PDZ

Postsynaptic density/Disks large/ZO-1

UKS

Unknown Significance

ZO-1

Zonula-Occludens-1

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Goodenough DA, Goliger JA, Paul DL. Connexins, connexons, and intercellular communication. Annual Review of Biochemistry. 1996;65:475–502. doi: 10.1146/annurev.bi.65.070196.002355. [DOI] [PubMed] [Google Scholar]
  • [2].Herve JC, Bourmeyster N, Sarrouilhe D, Duffy HS. Gap junctional complexes: from partners to functions. Prog Biophys Mol Biol. 2007;94:29–65. doi: 10.1016/j.pbiomolbio.2007.03.010. [DOI] [PubMed] [Google Scholar]
  • [3].Bruzzone R, White TW, Paul DL. Connections with connexins: the molecular basis of direct intercellular signaling. European Journal of Biochemistry. 1996;238:1–27. doi: 10.1111/j.1432-1033.1996.0001q.x. [DOI] [PubMed] [Google Scholar]
  • [4].Willecke K, Eiberger J, Degen J, Eckardt D, Romualdi A, Guldenagel M, Deutsch U, Sohl G. Structural and functional diversity of connexin genes in the mouse and human genome. Biol Chem. 2002;383:725–737. doi: 10.1515/BC.2002.076. [DOI] [PubMed] [Google Scholar]
  • [5].Musil LS, Goodenough DA. Biochemical analysis of connexin43 intracellular transport, phosphorylation, and assembly into gap junctional plaques. Journal of Cell Biology. 1991;115:1357–1374. doi: 10.1083/jcb.115.5.1357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [6].Laird DW. Life cycle of connexins in health and disease. Biochem J. 2006;394:527–543. doi: 10.1042/BJ20051922. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Sosinsky GE, Nicholson BJ. Structural organization of gap junction channels. Biochim Biophys Acta. 2005;1711:99–125. doi: 10.1016/j.bbamem.2005.04.001. [DOI] [PubMed] [Google Scholar]
  • [8].Paznekas WA, Karczeski B, Vermeer S, Lowry RB, Delatycki M, Laurence F, Koivisto PA, Van Maldergem L, Boyadjiev SA, Bodurtha JN, Jabs EW. GJA1 mutations, variants, and connexin 43 dysfunction as it relates to the oculodentodigital dysplasia phenotype. Hum Mutat. 2009;30:724–733. doi: 10.1002/humu.20958. [DOI] [PubMed] [Google Scholar]
  • [9].Severs NJ, Coppen SR, Dupont E, Yeh H, Ko Y, Matsushita T. Gap junction alterations in human cardiac disease. Cardiovascular Research. 2004;62:368–377. doi: 10.1016/j.cardiores.2003.12.007. [DOI] [PubMed] [Google Scholar]
  • [10].Miquerol L, Dupays L, Theveniau-Ruissy M, Alcolea S, Jarry-Guichard T, Abran P, Gros D. Gap junctional connexins in the developing mouse cardiac conduction system. Novartis Found Symp. 2003;250:80–98. doi: 10.1002/0470868066.ch6. discussion 98-109, 276-109. [DOI] [PubMed] [Google Scholar]
  • [11].Bruce AF, Rothery S, Dupont E, Severs NJ. Gap junction remodelling in human heart failure is associated with increased interaction of connexin43 with ZO-1. Cardiovasc Res. 2008;77:757–765. doi: 10.1093/cvr/cvm083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [12].Severs NJ, Bruce AF, Dupont E, Rothery S. Remodelling of gap junctions and connexin expression in diseased myocardium. Cardiovasc Res. 2008;80:9–19. doi: 10.1093/cvr/cvn133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [13].Wit AL. Remodeling of cardiac gap junctions: the relationship to the genesis of ventricular tachycardia. 2001;34(Suppl):77–83. doi: 10.1054/jelc.2001.28834. [DOI] [PubMed] [Google Scholar]
  • [14].Danik SB, Liu F, Zhang J, Suk HJ, Morley GE, Fishman GI, Gutstein DE. Modulation of cardiac gap junction expression and arrhythmic susceptibility. Circ Res. 2004;95:1035–1041. doi: 10.1161/01.RES.0000148664.33695.2a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Gutstein DE, Morley GE, Vaidya D, Liu F, Chen FL, Stuhlmann H, Fishman GI. Heterogeneous expression of Gap junction channels in the heart leads to conduction defects and ventricular dysfunction. Circulation. 2001;104:1194–1199. doi: 10.1161/hc3601.093990. [DOI] [PubMed] [Google Scholar]
  • [16].Je Saffitz. Gap junctions: functional effects of molecular structure and tissue distribution. Advances in Experimental Medicine & Biology. 1997;430:291–301. doi: 10.1007/978-1-4615-5959-7_25. [DOI] [PubMed] [Google Scholar]
  • [17].Poelzing S, Rosenbaum DS. Altered connexin43 expression produces arrhythmia substrate in heart failure. Am J Physiol Heart Circ Physiol. 2004;287:H1762–1770. doi: 10.1152/ajpheart.00346.2004. [DOI] [PubMed] [Google Scholar]
  • [18].Yao J, Hussain W, Patel P, Peters NS, Boyden PA, Wit AL. remodeling of gap junctional channel function in epicardial border zone of healing canine infarct. Circ. Res. 2003;92:437–443. doi: 10.1161/01.RES.0000059301.81035.06. [DOI] [PubMed] [Google Scholar]
  • [19].de Groot JR, Coronel R. Acute ischemia-induced gap junctional uncoupling and arrhythmogenesis. Cardiovascular Research. 2004;62:323–334. doi: 10.1016/j.cardiores.2004.01.033. [DOI] [PubMed] [Google Scholar]
  • [20].Rohr S. Role of gap junctions in the propagation of the cardiac action potential. Cardiovascular Research. 2004;62:309–322. doi: 10.1016/j.cardiores.2003.11.035. [DOI] [PubMed] [Google Scholar]
  • [21].Dhein S, Hagen A, Jozwiak J, Dietze A, Garbade J, Barten M, Kostelka M, Mohr FW. Improving cardiac gap junction communication as a new antiarrhythmic mechanism: the action of antiarrhythmic peptides. Naunyn Schmiedebergs Arch Pharmacol. 2010;381:221–34. doi: 10.1007/s00210-009-0473-1. [DOI] [PubMed] [Google Scholar]
  • [22].Delmar M. The intercalated disk as a single functional unit. Heart Rhythm. 2004;1:12–13. doi: 10.1016/j.hrthm.2004.02.003. [DOI] [PubMed] [Google Scholar]
  • [23].Severs NJ. Gap junction shape and orientation at the cardiac intercalated disk. Circ Res. 1989;65:1458–1462. doi: 10.1161/01.res.65.5.1458. [DOI] [PubMed] [Google Scholar]
  • [24].Palatinus JA, Gourdie RG. Xin and the art of intercalated disk maintenance. Am J Physiol Heart Circ Physiol. 2007;293:H2626–2628. doi: 10.1152/ajpheart.00954.2007. [DOI] [PubMed] [Google Scholar]
  • [25].Palatinus JA, Rhett JM, Gourdie RG. Translational lessons from scarless healing of cutaneous wounds and regenerative repair of the myocardium. J Mol Cell Cardiol. 2010;48:550–557. doi: 10.1016/j.yjmcc.2009.06.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Forbes MS, Sperelakis N. Intercalated discs of mammalian heart: a review of structure and function. Tissue Cell. 1985;17:605–648. doi: 10.1016/0040-8166(85)90001-1. [DOI] [PubMed] [Google Scholar]
  • [27].Sjostrand FS, Andersson E. Electron microscopy of the intercalated discs of cardiac muscle tissue. Experientia. 1954;10:369–370. doi: 10.1007/BF02160542. [DOI] [PubMed] [Google Scholar]
  • [28].Da Meddoff, Page E. Extensions of the cardiac plasma membrane from the intercalated disk of one cell into the transverse tubules of the adjacent cell. Journal of Ultrastructure Research. 1968;24:508–521. doi: 10.1016/s0022-5320(68)80051-6. [DOI] [PubMed] [Google Scholar]
  • [29].Saffitz JE, Davis LM, Darrow BJ, Kanter HL, Laing JG, Beyer EC. The molecular basis of anisotropy: role of gap junctions. J Cardiovasc Electrophysiol. 1995;6:498–510. doi: 10.1111/j.1540-8167.1995.tb00423.x. [DOI] [PubMed] [Google Scholar]
  • [30].Hoyt RH, Cohen ML, Saffitz JE. Distribution and three-dimensional structure of intercellular junctions in canine myocardium. Circ Res. 1989;64:563–574. doi: 10.1161/01.res.64.3.563. [DOI] [PubMed] [Google Scholar]
  • [31].Saffitz JE, Schuessler RB, Yamada KA. Mechanisms of remodeling of gap junction distributions and the development of anatomic substrates of arrhythmias. Cardiovasc. Res. 1999;42:309–317. doi: 10.1016/s0008-6363(99)00023-1. [DOI] [PubMed] [Google Scholar]
  • [32].Luke RA, Beyer EC, Hoyt RF, Saffitz JE. Quantitative analysis of intercellular connections by immunohistochemistry of the cardiac gap junction protein connexin43. Circ Res. 1989;65:1450–1457. doi: 10.1161/01.res.65.5.1450. [DOI] [PubMed] [Google Scholar]
  • [33].Green CR, Severs NJ. Gap junction connexon configuration in rapidly frozen myocardium and isolated intercalated disks. J Cell Biol. 1984;99:453–463. doi: 10.1083/jcb.99.2.453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Severs NJ, Gourdie RG, Harfst E, Peters NS, Green CR. Intercellular junctions and the application of microscopical techniques: the cardiac gap junction as a case model. Journal of Microscopy. 1993;169:299–328. doi: 10.1111/j.1365-2818.1993.tb03308.x. [DOI] [PubMed] [Google Scholar]
  • [35].Gros D, Mocquard JP, Challice CE, Schrevel J. Formation and growth of gap junctions in mouse myocardium during ontogenesis: a freeze-cleave study. J Cell Sci. 1978;30:45–61. doi: 10.1242/jcs.30.1.45. [DOI] [PubMed] [Google Scholar]
  • [36].Shibata Y, Nakata K, Page E. Ultrastructural changes during development of gap junctions in rabbit left ventricular myocardial cells. Journal of Ultrastructure Research. 1980;71:258–271. doi: 10.1016/s0022-5320(80)90078-7. [DOI] [PubMed] [Google Scholar]
  • [37].Paul DL. Molecular cloning of cDNA for rat liver gap junction protein. J Cell Biol. 1986;103:123–134. doi: 10.1083/jcb.103.1.123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [38].Beyer EC, Paul DL, Goodenough DA. Connexin43: a protein from rat heart homologous to a goap junction protein from liver. J Cell Biol. 1987;105:2621–2629. doi: 10.1083/jcb.105.6.2621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Beyer EC, Kistler J, Paul DL, Goodenough DA. Antisera directed against connexin43 peptides react with a 43-kD protein localized to gap junctions in myocardium and other tissues. J Cell Biol. 1989;108:595–605. doi: 10.1083/jcb.108.2.595. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [40].Fromaget C, el Aoumari A, Gros D. Distribution pattern of connexin 43, a gap junctional protein, during the differentiation of mouse heart myocytes. Differentiation. 1992;51:9–20. doi: 10.1111/j.1432-0436.1992.tb00675.x. [DOI] [PubMed] [Google Scholar]
  • [41].Gourdie RG, Green CR, Severs NJ. Gap junction distribution in adult mammalian myocardium revealed by an anti-peptide antibody and laser scanning confocal microscopy. Journal of Cell Science. 1991;99:41–55. doi: 10.1242/jcs.99.1.41. [DOI] [PubMed] [Google Scholar]
  • [42].Gourdie RG, Green CR, Severs NJ, Thompson RP. Three-dimensional reconstruction of gap junction arrangement in developing and adult rat hearts. Trans Roy Microscop Soc. 1990;1:417–420. [Google Scholar]
  • [43].Peters NS, Severs NJ, Rothery SM, Lincoln C, Yacoub MH, Green CR. Spatiotemporal relation between gap junctions and fascia adherens junctions during postnatal development of human ventricular myocardium [see comments] Circulation. 1994;90:713–725. doi: 10.1161/01.cir.90.2.713. [DOI] [PubMed] [Google Scholar]
  • [44].Kaprielian Rr, Gunning M, Dupont E, Sheppard Mn, Rothery Sm, Underwood R, Pennell Dj, Fox K, Pepper J, Poolewilson Pa, Severs Nj. Downregulation Of Immunodetectable Connexin43 and Decreased Gap Junction Size In the Pathogenesis Of Chronic Hibernation In the Human Left Ventricle. Circulation. 1998;97:651–660. doi: 10.1161/01.cir.97.7.651. [DOI] [PubMed] [Google Scholar]
  • [45].Gourdie RG, Harfst E, Severs NJ, Green CR. Cardiac gap junctions in rat ventricle: localization using site-directed antibodies and laser scanning confocal microscopy. Cardioscience. 1990;1:75–82. [PubMed] [Google Scholar]
  • [46].Green CR, Peters NS, Gourdie RG, Rothery S, Severs NJ. Validation of immunohistochemical quantification in confocal scanning laser microscopy: a comparative assessment of gap junction size with confocal and ultrastructural techniques. J Histochem Cytochem. 1993;41:1339–1349. doi: 10.1177/41.9.8354875. [DOI] [PubMed] [Google Scholar]
  • [47].Gourdie RG, Green CR, Severs NJ, Thompson RP. Immunolabelling patterns of gap junction connexins in the developing and mature rat heart. Anat Embryol (Berl) 1992;185:363–378. doi: 10.1007/BF00188548. [DOI] [PubMed] [Google Scholar]
  • [48].Spach MS, Dolber PC. Relating extracellular potentials and their derivates to anisotropic propagation at a microscopic level in human cardiac muscle. Evidence for electrical uncoupling of side-to-side fiber connections with increasing age. Circulation Research. 1986;58:356–371. doi: 10.1161/01.res.58.3.356. [DOI] [PubMed] [Google Scholar]
  • [49].Litchenberg WH, Norman LW, Holwell AK, Martin KL, Hewett KW, Gourdie RG. The rate and anisotropy of impulse propagation in the postnatal terminal crest are correlated with remodeling of Cx43 gap junction pattern. Cardiovascular Research. 2000;45:379–387. doi: 10.1016/s0008-6363(99)00363-6. [DOI] [PubMed] [Google Scholar]
  • [50].Spach MS, Heidlage JF, Dolber PC, Barr RC. Electrophysiological effects of remodeling cardiac gap junctions and cell size: experimental and model studies of normal cardiac growth. Circ Res. 2000;86:302–311. doi: 10.1161/01.res.86.3.302. [DOI] [PubMed] [Google Scholar]
  • [51].Smith JH, Green CR, Peters NS, Rothery S, Severs NJ. Altered patterns of gap junction distribution in ischemic heart disease. An immunohistochemical study of human myocardium using laser scanning confocal microscopy. Am J Pathol. 1991;139:801–821. [PMC free article] [PubMed] [Google Scholar]
  • [52].Luke RA, Saffitz JE. Remodeling of ventricular conduction pathways in healed canine infarct border zones. J Clin Invest. 1991;87:1594–1602. doi: 10.1172/JCI115173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Spach MS, Heidlage JF, Dolber PC, Barr RC. Changes in anisotropic conduction caused by remodeling cell size and the cellular distribution of gap junctions and Na(+) channels. J Electrocardiol. 2001;34(Suppl):69–76. doi: 10.1054/jelc.2001.28833. [DOI] [PubMed] [Google Scholar]
  • [54].Poelzing S, Rosenbaum DS. Nature, significance, and mechanisms of electrical heterogeneities in ventricle. Anat Rec A Discov Mol Cell Evol Biol. 2004;280:1010–1017. doi: 10.1002/ar.a.20103. [DOI] [PubMed] [Google Scholar]
  • [55].Wit AL, Duffy HS. Drug development for treatment of cardiac arrhythmias: targeting the gap junctions. Am J Physiol Heart Circ Physiol. 2008;294:H16–18. doi: 10.1152/ajpheart.01031.2007. [DOI] [PubMed] [Google Scholar]
  • [56].Cabo C, Yao J, Boyden PA, Chen S, Hussain W, Duffy HS, Ciaccio EJ, Peters NS, Wit AL. Heterogeneous gap junction remodeling in reentrant circuits in the epicardial border zone of the healing canine infarct. Cardiovasc Res. 2006;72:241–249. doi: 10.1016/j.cardiores.2006.07.005. [DOI] [PubMed] [Google Scholar]
  • [57].Kasi VS, Xiao HD, Shang LL, Iravanian S, Langberg J, Witham EA, Jiao Z, Gallego CJ, Bernstein KE, Dudley SC., Jr. Cardiac-restricted angiotensin-converting enzyme overexpression causes conduction defects and connexin dysregulation. Am J Physiol Heart Circ Physiol. 2007;293:H182–192. doi: 10.1152/ajpheart.00684.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [58].Peters NS, Coromilas J, Severs NJ, Wit AL. Disturbed connexin43 gap junction distribution correlates with the location of reentrant circuits in the epicardial border zone of healing canine infarcts that cause ventricular tachycardia. Circulation. 1997;95:988–996. doi: 10.1161/01.cir.95.4.988. [DOI] [PubMed] [Google Scholar]
  • [59].Kleber AG, Rudy Y. Basic mechanisms of cardiac impulse propagation and associated arrhythmias. Physiol Rev. 2004;84:431–488. doi: 10.1152/physrev.00025.2003. [DOI] [PubMed] [Google Scholar]
  • [60].Matsushita T, Oyamada M, Fujimoto K, Yasuda Y, Masuda S, Wada Y, Oka T, Takamatsu T. Remodeling of cell-cell and cell-extracellular matrix interactions at the border zone of rat myocardial infarcts. Circ Res. 1999;85:1046–1055. doi: 10.1161/01.res.85.11.1046. [DOI] [PubMed] [Google Scholar]
  • [61].Angst BD, Khan LU, Severs NJ, Whitely K, Rothery S, Thompson RP, Magee AI, Gourdie RG. Dissociated spatial patterning of gap junctions and cell adhesion junctions during postnatal differentiation of ventricular myocardium. Circ Res. 1997;80:88–94. doi: 10.1161/01.res.80.1.88. [DOI] [PubMed] [Google Scholar]
  • [62].Geisler SB, Green KJ, Isom LL, Meshinchi S, Martens JR, Delmar M, Russell MW. Ordered assembly of the adhesive and electrochemical connections within newly formed intercalated disks in primary cultures of adult rat cardiomyocytes. J Biomed Biotechnol. 2010;2010:624719. doi: 10.1155/2010/624719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Hirschy A, Schatzmann F, Ehler E, Perriard JC. Establishment of cardiac cytoarchitecture in the developing mouse heart. Dev Biol. 2006;289:430–441. doi: 10.1016/j.ydbio.2005.10.046. [DOI] [PubMed] [Google Scholar]
  • [64].Maass K, Shibayama J, Chase SE, Willecke K, Delmar M. C-terminal truncation of connexin43 changes number, size, and localization of cardiac gap junction plaques. Circ Res. 2007;101:1283–1291. doi: 10.1161/CIRCRESAHA.107.162818. [DOI] [PubMed] [Google Scholar]
  • [65].Sorgen PL, Duffy HS, Sahoo P, Coombs W, Delmar M, Spray DC. Structural changes in the carboxyl terminus of the gap junction protein connexin43 indicates signaling between binding domains for c-Src and zonula occludens-1. J Biol Chem. 2004;279:54695–54701. doi: 10.1074/jbc.M409552200. [DOI] [PubMed] [Google Scholar]
  • [66].Giepmans BN. Gap junctions and connexin-interacting proteins. Cardiovasc Res. 2004;62:233–245. doi: 10.1016/j.cardiores.2003.12.009. [DOI] [PubMed] [Google Scholar]
  • [67].Sorgen PL, Duffy HS, Spray DC, Delmar M. pH-dependent dimerization of the carboxyl terminal domain of Cx43. Biophys J. 2004;87:574–581. doi: 10.1529/biophysj.103.039230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Homma N, Alvarado JL, Coombs W, Stergiopoulos K, Taffet SM, Lau AF, Delmar M. A particle-receptor model for the insulin-induced closure of connexin43 channels. Circ Res. 1998;83:27–32. doi: 10.1161/01.res.83.1.27. [DOI] [PubMed] [Google Scholar]
  • [69].Morley GE, Taffet SM, Delmar M. Intramolecular interactions mediate pH regulation of connexin43 channels. Biophysical Journal. 1996;70:1294–1302. doi: 10.1016/S0006-3495(96)79686-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [70].Verma V, Larsen BD, Coombs W, Lin X, Spagnol G, Sorgen PL, Taffet SM, Delmar M. Novel Pharmacophores of Connexin43 Based on the “RXP” Series of Cx43-Binding Peptides. Circ Res. 2009;105:176–84. doi: 10.1161/CIRCRESAHA.109.200576. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Ponsaerts R, De Vuyst E, Retamal M, D’Hondt C, Vermeire D, Wang N, De Smedt H, Zimmermann P, Himpens B, Vereecke J, Leybaert L, Bultynck G. Intramolecular loop/tail interactions are essential for connexin 43-hemichannel activity. Faseb J. 2010;24:4378–95. doi: 10.1096/fj.09-153007. [DOI] [PubMed] [Google Scholar]
  • [72].Toyofuku T, Yabuki M, Otsu K, Kuzuya T, Hori M, Tada M. Direct association of the gap junction protein connexin-43 with ZO-1 in cardiac myocytes. J Biol Chem. 1998;273:12725–12731. doi: 10.1074/jbc.273.21.12725. [DOI] [PubMed] [Google Scholar]
  • [73].Giepmans BN, Moolenaar WH. The gap junction protein connexin43 interacts with the second PDZ domain of the zona occludens-1 protein. Current Biology. 1998;8:931–934. doi: 10.1016/s0960-9822(07)00375-2. [DOI] [PubMed] [Google Scholar]
  • [74].Segretain D, Fiorini C, Decrouy X, Defamie N, Prat JR, Pointis G. A proposed role for ZO-1 in targeting connexin 43 gap junctions to the endocytic pathway. Biochimie. 2004;86:241–244. doi: 10.1016/j.biochi.2004.05.003. [DOI] [PubMed] [Google Scholar]
  • [75].Nielsen PA, Baruch A, Shestopalov VI, Giepmans BN, Dunia I, Benedetti EL, Kumar NM. Lens connexins alpha3Cx46 and alpha8Cx50 interact with zonula occludens protein-1 (ZO-1) Mol Biol Cell. 2003;14:2470–2481. doi: 10.1091/mbc.E02-10-0637. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [76].Jinn Y, Inase N. Connexin43, E-Cadherin, {beta}-Catenin and ZO-1 Expression, and Aberrant Methylation of the Connexin43 Gene in NSCLC. Anticancer Res. 2010;30:2271–2278. [PubMed] [Google Scholar]
  • [77].Barker RJ, Price RL, Gourdie RG. Increased co-localization of connexin43 and ZO-1 in dissociated adult myocytes. Cell Commun Adhes. 2001;8:205–208. doi: 10.3109/15419060109080724. [DOI] [PubMed] [Google Scholar]
  • [78].Giepmans BN, Verlaan I, Moolenaar WH. Connexin-43 interactions with ZO-1 and alpha- and beta-tubulin. Cell Commun Adhes. 2001;8:219–223. doi: 10.3109/15419060109080727. [DOI] [PubMed] [Google Scholar]
  • [79].Barker RJ, Gang C, Jourdan J, Zhang Y, Price RL, Bond J, Harris BS, Hewett KW, G RG. Connexin and ZO1 Interactions During Gap Junction Remodeling. Einthoven 2002: 100 Years of Electrocardiography. 2002:161–165. [Google Scholar]
  • [80].Stevenson BR, Siliciano JD, Mooseker MS, Goodenough DA. Identification of ZO-1: a high molecular weight polypeptide associated with the tight junction (zonula occludens) in a variety of epithelia. J Cell Biol. 1986;103:755–766. doi: 10.1083/jcb.103.3.755. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Anderson JM. Cell signalling: MAGUK magic. Curr Biol. 1996;6:382–384. doi: 10.1016/s0960-9822(02)00501-8. [DOI] [PubMed] [Google Scholar]
  • [82].Hartsock A, Nelson WJ. Adherens and tight junctions: structure, function and connections to the actin cytoskeleton. Biochim Biophys Acta. 2008;1778:660–669. doi: 10.1016/j.bbamem.2007.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Fanning AS, Lye MF, Anderson JM, Lavie A. Domain swapping within PDZ2 is responsible for dimerization of ZO proteins. J Biol Chem. 2007;282:37710–37716. doi: 10.1074/jbc.M707255200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [84].Xiao F, Weng J, Fan K, Wang W. Detailed Regulatory Mechanism of the Interaction between ZO-1 PDZ2 and Connexin43 Revealed by MD Simulations. PLoS One. 2011;6:e21527. doi: 10.1371/journal.pone.0021527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Kausalya PJ, Reichert M, Hunziker W. Connexin45 directly binds to ZO-1 and localizes to the tight junction region in epithelial MDCK cells. FEBS Letters. 2001;505:92–96. doi: 10.1016/s0014-5793(01)02786-7. [DOI] [PubMed] [Google Scholar]
  • [86].Laing JG, Chou BC, Steinberg TH. ZO-1 alters the plasma membrane localization and function of Cx43 in osteoblastic cells. J Cell Sci. 2005;118:2167–2176. doi: 10.1242/jcs.02329. [DOI] [PubMed] [Google Scholar]
  • [87].Nielsen PA, Beahm DL, Giepmans BN, Baruch A, Hall JE, Kumar NM. Molecular cloning, functional expression, and tissue distribution of a novel human gap junction-forming protein, connexin-31.9. Interaction with zona occludens protein-1. Journal of Biological Chemistry. 2002;277:38272–38283. doi: 10.1074/jbc.M205348200. [DOI] [PubMed] [Google Scholar]
  • [88].Flores CE, Li X, Bennett MV, Nagy JI, Pereda AE. Interaction between connexin35 and zonula occludens-1 and its potential role in the regulation of electrical synapses. Proc Natl Acad Sci U S A. 2008;105:12545–12550. doi: 10.1073/pnas.0804793105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [89].Jin C, Lau AF, Martyn KD. Identification of connexin-interacting proteins: application of the yeast two-hybrid screen. Methods. 2000;20:219–231. doi: 10.1006/meth.1999.0939. [DOI] [PubMed] [Google Scholar]
  • [90].Singh D, Solan JL, Taffet SM, Javier R, Lampe PD. Connexin 43 interacts with zona occludens-1 and -2 proteins in a cell cycle stage-specific manner. J Biol Chem. 2005;280:30416–30421. doi: 10.1074/jbc.M506799200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [91].Fanning AS, Ma TY, Anderson JM. Isolation and functional characterization of the actin binding region in the tight junction protein ZO-1. Faseb J. 2002;16:1835–1837. doi: 10.1096/fj.02-0121fje. [DOI] [PubMed] [Google Scholar]
  • [92].Barker RJ, Price RL, Gourdie RG. Increased association of ZO-1 with connexin43 during remodeling of cardiac gap junctions. Circ Res. 2002;90:317–324. doi: 10.1161/hh0302.104471. [DOI] [PubMed] [Google Scholar]
  • [93].Barker RJ, Gourdie RG. In: Connexin Interacting Proteins, Gap junctional Communication in Health and Disease. 2nd Edition De Mello WC, editor. Kluwer: 2003. pp. 25–53. [Google Scholar]
  • [94].Girao H, Pereira P. The proteasome regulates the interaction between Cx43 and ZO-1. J Cell Biochem. 2007;102:719–728. doi: 10.1002/jcb.21351. [DOI] [PubMed] [Google Scholar]
  • [95].Rhett JM, Jourdan J, Gourdie RG. Connexin 43 connexon to gap junction transition is regulated by zonula occludens-1. Mol Biol Cell. 2011;22:1516–1528. doi: 10.1091/mbc.E10-06-0548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [96].Akoyev V, Takemoto DJ. ZO-1 is required for protein kinase C gamma-driven disassembly of connexin 43. Cell Signal. 2007;19:958–967. doi: 10.1016/j.cellsig.2006.11.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [97].Toyofuku T, Akamatsu Y, Zhang H, Kuzuya T, Tada M, Hori M. c-Src regulates the interaction between connexin-43 and ZO-1 in cardiac myocytes. J Biol Chem. 2001;276:1780–1788. doi: 10.1074/jbc.M005826200. [DOI] [PubMed] [Google Scholar]
  • [98].Gilleron J, Fiorini C, Carette D, Avondet C, Falk MM, Segretain D, Pointis G. Molecular reorganization of Cx43, ZO-1 and Src complexes during the endocytosis of gap junction plaques in response to a non-genomic carcinogen. J Cell Sci. 2008;121:4069–78. doi: 10.1242/jcs.033373. [DOI] [PubMed] [Google Scholar]
  • [99].Derangeon M, Bourmeyster N, Plaisance I, Pinet-Charvet C, Chen Q, Duthe F, Popoff MR, Sarrouilhe D, Herve JC. RhoA GTPase and F-actin dynamically regulate the permeability of Cx43-made channels in rat cardiac myocytes. J Biol Chem. 2008;283:30754–30765. doi: 10.1074/jbc.M801556200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Hunter AW, Jourdan J, Gourdie RG. Fusion of GFP to the carboxyl terminus of connexin43 increases gap junction size in HeLa cells. Cell Commun Adhes. 2003;10:211–214. doi: 10.1080/cac.10.4-6.211.214. [DOI] [PubMed] [Google Scholar]
  • [101].Hunter AW, Barker RJ, Zhu C, Gourdie RG. Zonula occludens-1 alters connexin43 gap junction size and organization by influencing channel accretion. Mol Biol Cell. 2005;16:5686–5698. doi: 10.1091/mbc.E05-08-0737. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Hunter AW, Gourdie RG. The second PDZ domain of zonula occludens-1 is dispensable for targeting to connexin 43 gap junctions. Cell Commun Adhes. 2008;15:55–63. doi: 10.1080/15419060802014370. [DOI] [PubMed] [Google Scholar]
  • [103].Baker SM, Kim N, Gumpert AM, Segretain D, Falk MM. Acute internalization of gap junctions in vascular endothelial cells in response to inflammatory mediator-induced G-protein coupled receptor activation. FEBS Lett. 2008;582:4039–4046. doi: 10.1016/j.febslet.2008.10.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Severs NJ, Shovel KS, Slade AM, Powell T, Twist VW, Green CR. Fate of gap junctions in isolated adult mammalian cardiomyocytes. Circ Res. 1989;65:22–42. doi: 10.1161/01.res.65.1.22. [DOI] [PubMed] [Google Scholar]
  • [105].Crow DS, Beyer EC, Paul DL, Kobe SS, Lau AF. Phosphorylation of connexin43 gap junction protein in uninfected and Rous sarcoma virus-transformed mammalian fibroblasts. Mol Cell Biol. 1990;10:1754–1763. doi: 10.1128/mcb.10.4.1754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [106].Filson AJ, Azarnia R, Beyer EC, Loewenstein WR, Brugge JS. Tyrosine phosphorylation of a gap junction protein correlates with inhibition of cell-to-cell communication. Cell Growth Differ. 1990;1:661–668. [PubMed] [Google Scholar]
  • [107].Lin R, Warn-Cramer BJ, Kurata WE, Lau AF. v-Src-mediated phosphorylation of connexin43 on tyrosine disrupts gap junctional communication in mammalian cells. Cell Commun Adhes. 2001;8:265–269. doi: 10.3109/15419060109080735. [DOI] [PubMed] [Google Scholar]
  • [108].Toyofuku T, Yabuki M, Otsu K, Kuzuya T, Tada M, Hori M. Functional role of c-Src in gap junctions of the cardiomyopathic heart. Circ. Res. 1999;85:672–681. doi: 10.1161/01.res.85.8.672. [DOI] [PubMed] [Google Scholar]
  • [109].Kieken F, Mutsaers N, Dolmatova E, Virgil K, Wit AL, Kellezi A, Hirst-Jensen BJ, Duffy HS, Sorgen PL. Structural and Molecular Mechanisms of Gap Junction Remodeling in Epicardial Border Zone Myocytes following Myocardial Infarction. Circ Res. 2009;104:1103–12. doi: 10.1161/CIRCRESAHA.108.190454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [110].Gilleron J, Carette D, Fiorini C, Dompierre J, Macia E, Denizot JP, Segretain D, Pointis G. The large GTPase dynamin2: A new player in connexin 43 gap junction endocytosis, recycling and degradation. Int J Biochem Cell Biol. 2011;43:1208–17. doi: 10.1016/j.biocel.2011.04.014. [DOI] [PubMed] [Google Scholar]
  • [111].van Zeijl L, Ponsioen B, Giepmans BN, Ariaens A, Postma FR, Varnai P, Balla T, Divecha N, Jalink K, Moolenaar WH. Regulation of connexin43 gap junctional communication by phosphatidylinositol 4,5-bisphosphate. J Cell Biol. 2007;177:881–891. doi: 10.1083/jcb.200610144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [112].Piehl M, Lehmann C, Gumpert A, Denizot JP, Segretain D, Falk MM. Internalization of large double-membrane intercellular vesicles by a clathrin-dependent endocytic process. Mol Biol Cell. 2007;18:337–347. doi: 10.1091/mbc.E06-06-0487. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Duffy HS, Ashton AW, O’Donnell P, Coombs W, Taffet SM, Delmar M, Spray DC. Regulation of connexin43 protein complexes by intracellular acidification. Circ Res. 2004;94:215–222. doi: 10.1161/01.RES.0000113924.06926.11. [DOI] [PubMed] [Google Scholar]
  • [114].Hunter A, Gourdie R. Mictotubules and actin differentially influence remodeling of connexin43 gap junctions. The FASEB Journal. 2008;22:800–804. [Google Scholar]
  • [115].Zhu C, Barker RJ, Hunter AW, Zhang Y, Jourdan J, Gourdie RG. Quantitative analysis of ZO-1 colocalization with Cx43 gap junction plaques in cultures of rat neonatal cardiomyocytes. Microsc Microanal. 2005;11:244–248. doi: 10.1017/S143192760505049X. [DOI] [PubMed] [Google Scholar]
  • [116].Lauf U, Giepmans BN, Lopez P, Braconnot S, Chen SC, Falk MM. Dynamic trafficking and delivery of connexons to the plasma membrane and accretion to gap junctions in living cells. Proc Natl Acad Sci U S A. 2002;99:10446–10451. doi: 10.1073/pnas.162055899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [117].Gaietta G, Deerinck TJ, Adams SR, Bouwer J, Tour O, Laird DW, Sosinsky GE, Tsien RY, Ellisman MH. Multicolor and electron microscopic imaging of connexin trafficking. Science. 2002;296:503–507. doi: 10.1126/science.1068793. [DOI] [PubMed] [Google Scholar]
  • [118].Lim BK, Xiong D, Dorner A, Youn TJ, Yung A, Liu TI, Gu Y, Dalton ND, Wright AT, Evans SM, Chen J, Peterson KL, McCulloch AD, Yajima T, Knowlton KU. Coxsackievirus and adenovirus receptor (CAR) mediates atrioventricular-node function and connexin 45 localization in the murine heart. J Clin Invest. 2008;118:2758–2770. doi: 10.1172/JCI34777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [119].Grikscheit K, Thomas N, Bruce AF, Rothery S, Chan J, Severs NJ, Dupont E. Coexpression of connexin 45 with connexin 43 decreases gap junction size. Cell Commun Adhes. 2008;15:185–193. doi: 10.1080/15419060802013943. [DOI] [PubMed] [Google Scholar]
  • [120].Palatinus JA, O’Quinn MP, Barker RJ, Harris BS, Jourdan J, Gourdie RG. ZO-1 Determines Adherens and Gap Junction Localization at Intercalated Disks. Am J Physiol Heart Circ Physiol. 2010;300:H583–94. doi: 10.1152/ajpheart.00999.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [121].Maass K, Chase SE, Lin X, Delmar M. Cx43 CT Domain Influences Infarct Size and Susceptibility to Ventricular Tachyarrhythmias in Acute Myocardial Infarction. Cardiovasc Res. 2009;84:361–7. doi: 10.1093/cvr/cvp250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [122].Fonseca CG, Green CR, Nicholson LF. Upregulation in astrocytic connexin 43 gap junction levels may exacerbate generalized seizures in mesial temporal lobe epilepsy. Brain Res. 2002;929:105–116. doi: 10.1016/s0006-8993(01)03289-9. [DOI] [PubMed] [Google Scholar]
  • [123].Laing JG, Saffitz JE, Steinberg TH, Yamada KA. Diminished zonula occludens-1 expression in the failing human heart. Cardiovasc Pathol. 2007;16:159–164. doi: 10.1016/j.carpath.2007.01.004. [DOI] [PubMed] [Google Scholar]
  • [124].Kostin S. Zonula occludens-1 and connexin 43 expression in the failing human heart. J Cell Mol Med. 2007;11:892–895. doi: 10.1111/j.1582-4934.2007.00063.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [125].Sasano C, Honjo H, Takagishi Y, Uzzaman M, Emdad L, Shimizu A, Murata Y, Kamiya K, Kodama I. Internalization and dephosphorylation of connexin43 in hypertrophied right ventricles of rats with pulmonary hypertension. Circ J. 2007;71:382–389. doi: 10.1253/circj.71.382. [DOI] [PubMed] [Google Scholar]
  • [126].Jin H, Chemaly ER, Lee A, Kho C, Hadri L, Hajjar RJ, Akar FG. Mechanoelectrical remodeling and arrhythmias during progression of hypertrophy. Faseb J. 2009;24:451–63. doi: 10.1096/fj.09-136622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Goldfine SM, Walcott B, Brink PR, Magid NM, Borer JS. Myocardial connexin43 expression in left ventricular hypertrophy resulting from aortic regurgitation. Cardiovasc Pathol. 1999;8:1–6. doi: 10.1016/s1054-8807(98)00011-8. [DOI] [PubMed] [Google Scholar]
  • [128].Benes J, Kazdova L, Drahota Z, Houstek J, Medrikova D, Kopecky J, Kovarova N, Vrbacky M, Sedmera D, Strnad H, Kolar M, Petrak J, Benada O, Skaroupkova P, Cervenka L, Melenovsky V. Effect of metformin therapy on cardiac function and survival in a volume-overload model of heart failure in rats. Clin Sci (Lond) 2011;121:29–41. doi: 10.1042/CS20100527. [DOI] [PubMed] [Google Scholar]
  • [129].Saffitz JE, Green KG, Kraft WJ, Schechtman KB, Yamada KA. Effects of diminished expression of connexin43 on gap junction number and size in ventricular myocardium. Am J Physiol Heart Circ Physiol. 2000;278:H1662–1670. doi: 10.1152/ajpheart.2000.278.5.H1662. [DOI] [PubMed] [Google Scholar]
  • [130].Chen L, Ge Goings, Upshaw-Earley J, Page E. Cardiac gap junctions and gap junction-associated vesicles: ultrastructural comparison of in situ negative staining with conventional positive staining. Circulation Research. 1989;64:501–514. doi: 10.1161/01.res.64.3.501. [DOI] [PubMed] [Google Scholar]
  • [131].Legato MJ. Cellular mechanisms of normal growth in the mammalian heart. I. Qualitative and quantitative features of ventricular architecture in the dog from birth to five months of age. Circ Res United States. 1979;44:250–262. doi: 10.1161/01.res.44.2.250. [DOI] [PubMed] [Google Scholar]
  • [132].Muhlfeld C, Richter J. High-pressure freezing and freeze substitution of rat myocardium for immunogold labeling of connexin 43. Anat Rec A Discov Mol Cell Evol Biol. 2006;288:1059–1067. doi: 10.1002/ar.a.20380. [DOI] [PubMed] [Google Scholar]
  • [133].Matsuda T, Fujio Y, Nariai T, Ito T, Yamane M, Takatani T, Takahashi K, Azuma J. N-cadherin signals through Rac1 determine the localization of connexin 43 in cardiac myocytes. J Mol Cell Cardiol. 2006;40:495–502. doi: 10.1016/j.yjmcc.2005.12.010. [DOI] [PubMed] [Google Scholar]
  • [134].de Jonge HW, Dekkers DH, Houtsmuller AB, Sharma HS, Lamers JM. Differential Signaling and Hypertrophic Responses in Cyclically Stretched vs Endothelin-1 Stimulated Neonatal Rat Cardiomyocytes. Cell Biochem Biophys. 2007;47:21–32. doi: 10.1385/cbb:47:1:21. [DOI] [PubMed] [Google Scholar]
  • [135].Matsuda T, Takahashi K, Nariai T, Ito T, Takatani T, Fujio Y, Azuma J. N-cadherin-mediated cell adhesion determines the plasticity for cell alignment in response to mechanical stretch in cultured cardiomyocytes. Biochem Biophys Res Commun. 2005;326:228–232. doi: 10.1016/j.bbrc.2004.11.019. [DOI] [PubMed] [Google Scholar]
  • [136].Salameh A, Wustmann A, Karl S, Blanke K, Apel D, Rojas-Gomez D, Franke H, Mohr FW, Janousek J, Dhein S. Cyclic Mechanical Stretch Induces Cardiomyocyte Orientation and Polarization of the Gap Junction Protein Connexin43. Circ Res. 2010 doi: 10.1161/CIRCRESAHA.109.214429. [DOI] [PubMed] [Google Scholar]
  • [137].Yamada K, Green KG, Samarel AM, Saffitz JE. Distinct pathways regulate expression of cardiac electrical and mechanical junction proteins in response to stretch. Circ Res. 2005;97:346–353. doi: 10.1161/01.RES.0000178788.76568.8a. [DOI] [PubMed] [Google Scholar]
  • [138].Gopalan SM, Flaim C, Bhatia SN, Hoshijima M, Knoell R, Chien KR, Omens JH, McCulloch AD. Anisotropic stretch-induced hypertrophy in neonatal ventricular myocytes micropatterned on deformable elastomers. Biotechnol Bioeng. 2003;81:578–587. doi: 10.1002/bit.10506. [DOI] [PubMed] [Google Scholar]
  • [139].Hertig CM, Butz S, Koch S, Eppenberger-Eberhardt M, Kemler R, Eppenberger HM. N-cadherin in adult rat cardiomyocytes in culture. II. Spatio-temporal appearance of proteins involved in cell-cell contact and communication. Formation of two distinct N-cadherin/catenin complexes. J Cell Sci. 1996;109(Pt 1):11–20. doi: 10.1242/jcs.109.1.11. [DOI] [PubMed] [Google Scholar]
  • [140].Li J, Levin MD, Xiong Y, Petrenko N, Patel VV, Radice GL. N-cadherin haploinsufficiency affects cardiac gap junctions and arrhythmic susceptibility. J Mol Cell Cardiol. 2008;44:597–606. doi: 10.1016/j.yjmcc.2007.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [141].Giepmans BN, Verlaan I, Hengeveld T, Janssen H, Calafat J, Falk MM, Moolenaar WH. Gap junction protein connexin-43 interacts directly with microtubules. Curr Biol. 2001;11:1364–1368. doi: 10.1016/s0960-9822(01)00424-9. [DOI] [PubMed] [Google Scholar]
  • [142].Shaw RM, Fay AJ, Puthenveedu MA, von Zastrow M, Jan YN, Jan LY. Microtubule plus-end-tracking proteins target gap junctions directly from the cell interior to adherens junctions. Cell. 2007;128:547–560. doi: 10.1016/j.cell.2006.12.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [143].Wei CJ, Francis R, Xu X, Lo CW. Connexin43 associated with an N-cadherin-containing multiprotein complex is required for gap junction formation in NIH3T3 cells. J Biol Chem. 2005;280:19925–19936. doi: 10.1074/jbc.M412921200. [DOI] [PubMed] [Google Scholar]
  • [144].Smyth JW, Hong TT, Gao D, Vogan JM, Jensen BC, Fong TS, Simpson PC, Stainier DY, Chi NC, Shaw RM. Limited forward trafficking of connexin 43 reduces cell-cell coupling in stressed human and mouse myocardium. J Clin Invest. 2010;120:266–279. doi: 10.1172/JCI39740. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [145].Dai P, Nakagami T, Tanaka H, Hitomi T, Takamatsu T. Cx43 mediates TGF-beta signaling through competitive Smads binding to microtubules. Mol Biol Cell. 2007;18:2264–2273. doi: 10.1091/mbc.E06-12-1064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [146].Bauer R, Lehmann C, Martini J, Eckardt F, Hoch M. Gap junction channel protein innexin 2 is essential for epithelial morphogenesis in the Drosophila embryo. Mol Biol Cell. 2004;15:2992–3004. doi: 10.1091/mbc.E04-01-0056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [147].Akar FG, Spragg DD, Tunin RS, Kass DA, Tomaselli GF. Mechanisms underlying conduction slowing and arrhythmogenesis in nonischemic dilated cardiomyopathy. Circ Res. 2004;95:717–725. doi: 10.1161/01.RES.0000144125.61927.1c. [DOI] [PubMed] [Google Scholar]
  • [148].Akar FG, Nass RD, Hahn S, Cingolani E, Shah M, Hesketh GG, DiSilvestre D, Tunin RS, Kass DA, Tomaselli GF. Dynamic changes in conduction velocity and gap junction properties during development of pacing-induced heart failure. Am J Physiol Heart Circ Physiol. 2007;293:H1223–1230. doi: 10.1152/ajpheart.00079.2007. [DOI] [PubMed] [Google Scholar]
  • [149].Hesketh GG, Shah MH, Halperin VL, Cooke CA, Akar FG, Yen TE, Kass DA, Machamer CE, Van Eyk JE, Tomaselli GF. Ultrastructure and Regulation of Lateralized Connexin43 in the Failing Heart. Circ Res. 2010 doi: 10.1161/CIRCRESAHA.108.182147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [150].Nygren A, Olson ML, Chen KY, Emmett T, Kargacin G, Shimoni Y. Propagation of the cardiac impulse in the diabetic rat heart: reduced conduction reserve. J Physiol. 2007;580:543–560. doi: 10.1113/jphysiol.2006.123729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [151].Kostin S, Dammer S, Hein S, Klovekorn WP, Bauer EP, Schaper J. Connexin 43 expression and distribution in compensated and decompensated cardiac hypertrophy in patients with aortic stenosis. Cardiovasc Res. 2004;62:426–436. doi: 10.1016/j.cardiores.2003.12.010. [DOI] [PubMed] [Google Scholar]
  • [152].Sepp R, Severs Nj, Gourdie Rg. Altered Patterns Of Cardiac Intercellular Junction Distribution In Hypertrophic Cardiomyopathy. Heart. 1996;76:412–417. doi: 10.1136/hrt.76.5.412. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Uzzaman M, Honjo H, Takagishi Y, Emdad L, Magee AI, Severs NJ, Kodama I. Remodeling of gap junctional coupling in hypertrophied right ventricles of rats with monocrotaline-induced pulmonary hypertension. Circ Res. 2000;86:871–878. doi: 10.1161/01.res.86.8.871. [DOI] [PubMed] [Google Scholar]
  • [154].Gehmlich K, Syrris P, Peskett E, Evans A, Ehler E, Asimaki A, Anastasakis A, Tsatsopoulou A, Vouliotis AI, Stefanadis C, Saffitz JE, Protonotarios N, McKenna WJ. Mechanistic insights into arrhythmogenic right ventricular cardiomyopathy caused by desmocollin-2 mutations. Cardiovasc Res. 2011;90:77–87. doi: 10.1093/cvr/cvq353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [155].Bolling MC, Jonkman MF. Skin and heart: une liaison dangereuse. Exp Dermatol. 2009;18:658–668. doi: 10.1111/j.1600-0625.2009.00908.x. [DOI] [PubMed] [Google Scholar]
  • [156].Li J, Radice GL. A new perspective on intercalated disc organization: implications for heart disease. Dermatol Res Pract. 2010;2010:207835. doi: 10.1155/2010/207835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [157].Lombardi R, Marian AJ. Arrhythmogenic right ventricular cardiomyopathy is a disease of cardiac stem cells. Curr Opin Cardiol. 2010 doi: 10.1097/HCO.0b013e3283376daf. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Saffitz J, Arrhythmogenic E. cardiomyopathy and abnormalities of cell-to-cell coupling. Heart rhythm: the official journal of the Heart Rhythm Society. 2009;6:S62–S65. doi: 10.1016/j.hrthm.2009.03.003. [DOI] [PubMed] [Google Scholar]
  • [159].Fidler LM, Wilson GJ, Liu F, Cui X, Scherer SW, Taylor GP, Hamilton RM. Abnormal connexin43 in arrhythmogenic right ventricular cardiomyopathy caused by plakophilin-2 mutations. J Cell Mol Med. 2009;13:4219–4228. doi: 10.1111/j.1582-4934.2008.00438.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [160].Oxford EM, Musa H, Maass K, Coombs W, Taffet SM, Delmar M. Connexin43 remodeling caused by inhibition of plakophilin-2 expression in cardiac cells. Circ Res. 2007;101:703–711. doi: 10.1161/CIRCRESAHA.107.154252. [DOI] [PubMed] [Google Scholar]
  • [161].Saffitz JE. Dependence of electrical coupling on mechanical coupling in cardiac myocytes: insights gained from cardiomyopathies caused by defects in cell-cell connections. Ann N Y Acad Sci. 2005;1047:336–344. doi: 10.1196/annals.1341.030. [DOI] [PubMed] [Google Scholar]
  • [162].Ai X, Pogwizd SM. Connexin 43 downregulation and dephosphorylation in nonischemic heart failure is associated with enhanced colocalized protein phosphatase type 2A. Circ Res. 2005;96:54–63. doi: 10.1161/01.RES.0000152325.07495.5a. [DOI] [PubMed] [Google Scholar]
  • [163].Lampe PD, Lau AF. Regulation of gap junctions by phosphorylation of connexins. Arch Biochem Biophys. 2000;384:205–215. doi: 10.1006/abbi.2000.2131. [DOI] [PubMed] [Google Scholar]
  • [164].Solan JL, Fry MD, TenBroek EM, Lampe PD. Connexin43 phosphorylation at S368 is acute during S and G2/M and in response to protein kinase C activation. J Cell Sci. 2003;116:2203–2211. doi: 10.1242/jcs.00428. [DOI] [PubMed] [Google Scholar]
  • [165].Kalcheva N, Qu J, Sandeep N, Garcia L, Zhang J, Wang Z, Lampe PD, Suadicani SO, Spray DC, Fishman GI. Gap junction remodeling and cardiac arrhythmogenesis in a murine model of oculodentodigital dysplasia. Proc Natl Acad Sci U S A. 2007;104:20512–20516. doi: 10.1073/pnas.0705472105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [166].TenBroek EM, Lampe PD, Solan JL, Reynhout JK, Johnson RG. Ser364 of connexin43 and the upregulation of gap junction assembly by cAMP. J Cell Biol. 2001;155:1307–1318. doi: 10.1083/jcb.200102017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [167].Qu J, Volpicelli FM, Garcia LI, Sandeep N, Zhang J, Marquez-Rosado L, Lampe PD, Fishman GI. Gap junction remodeling and spironolactone-dependent reverse remodeling in the hypertrophied heart. Circ Res. 2009;104:365–371. doi: 10.1161/CIRCRESAHA.108.184044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [168].Remo BF, Qu J, Volpicelli FM, Giovannone S, Shin D, Lader J, Liu FY, Zhang J, Lent DS, Morley GE, Fishman GI. Phosphatase-Resistant Gap Junctions Inhibit Pathological Remodeling and Prevent Arrhythmias. Circ Res. 2011;108:1459–66. doi: 10.1161/CIRCRESAHA.111.244046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [169].Chen J, Pan L, Wei Z, Zhao Y, Zhang M. Domain-swapped dimerization of ZO-1 PDZ2 generates specific and regulatory connexin43-binding sites. Embo J. 2008;27:2113–2123. doi: 10.1038/emboj.2008.138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [170].Ek-Vitorin JF, King TJ, Heyman NS, Lampe PD, Burt JM. Selectivity of connexin 43 channels is regulated through protein kinase C-dependent phosphorylation. Circ Res. 2006;98:1498–1505. doi: 10.1161/01.RES.0000227572.45891.2c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [171].Srisakuldee W, Jeyaraman MM, Nickel BE, Tanguy S, Jiang ZS, Kardami E. Phosphorylation of connexin-43 at serine 262 promotes a cardiac injury-resistant state. Cardiovasc Res. 2009;83:672–681. doi: 10.1093/cvr/cvp142. [DOI] [PubMed] [Google Scholar]
  • [172].Jozwiak J, Dhein S. Local effects and mechanisms of antiarrhythmic peptide AAP10 in acute regional myocardial ischemia: electrophysiological and molecular findings. Naunyn Schmiedebergs Arch Pharmacol. 2008;378:459–470. doi: 10.1007/s00210-008-0317-4. [DOI] [PubMed] [Google Scholar]
  • [173].O’Quinn MP, Palatinus JA, Harris BS, Hewett KW, Gourdie RG. A Peptide Mimetic of the Connexin43 Carboxyl Terminus Reduces Gap Junction Remodeling and Induced Arrhythmia Following Ventricular Injury. Circ Res. 2011;108:704–15. doi: 10.1161/CIRCRESAHA.110.235747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [174].Lindsey ML, Escobar GP, Mukherjee R, Goshorn DK, Sheats NJ, Bruce JA, Mains IM, Hendrick JK, Hewett KW, Gourdie RG, Matrisian LM, Spinale FG. Matrix metalloproteinase-7 affects connexin-43 levels, electrical conduction, and survival after myocardial infarction. Circulation. 2006;113:2919–2928. doi: 10.1161/CIRCULATIONAHA.106.612960. [DOI] [PubMed] [Google Scholar]
  • [175].Palatinus JA, Rhett JM, Gourdie RG. Enhanced PKCepsilon mediated phosphorylation of connexin43 at serine 368 by a carboxyl-terminal mimetic peptide is dependent on injury. Channels (Austin) 2011;5:236–40. doi: 10.4161/chan.5.3.15834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [176].Huang RY, Laing JG, Kanter EM, Berthoud VM, Bao M, Rohrs HW, Townsend RR, Yamada KA. Identification of CaMKII Phosphorylation Sites in Connexin43 by High-Resolution Mass Spectrometry. J Proteome Res. 2011;10:1098–109. doi: 10.1021/pr1008702. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [177].Lau AF, Kurata WE, Kanemitsu MY, Loo LW, Warn-Cramer BJ, Eckhart W, Lampe PD. Regulation of connexin43 function by activated tyrosine protein kinases. J Bioenerg Biomembr. 1996;28:359–368. doi: 10.1007/BF02110112. [DOI] [PubMed] [Google Scholar]
  • [178].Sirnes S, Kjenseth A, Leithe E, Rivedal E. Interplay between PKC and the MAP kinase pathway in Connexin43 phosphorylation and inhibition of gap junction intercellular communication. Biochem Biophys Res Commun. 2009;382:41–45. doi: 10.1016/j.bbrc.2009.02.141. [DOI] [PubMed] [Google Scholar]
  • [179].Warn-Cramer BJ, Lampe PD, Kurata WE, Kanemitsu MY, Loo LW, Eckhart W, Lau AF. Characterization of the mitogen-activated protein kinase phosphorylation sites on the connexin-43 gap junction protein. J Biol Chem. 1996;271:3779–3786. doi: 10.1074/jbc.271.7.3779. [DOI] [PubMed] [Google Scholar]
  • [180].Lan Z, Kurata WE, Martyn KD, Jin C, Lau AF. Novel rab GAP-like protein, CIP85, interacts with connexin43 and induces its degradation. Biochemistry. 2005;44:2385–2396. doi: 10.1021/bi048306w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [181].Doble BW, Dang X, Ping P, Fandrich RR, Nickel BE, Jin Y, Cattini PA, Kardami E. Phosphorylation of serine 262 in the gap junction protein connexin-43 regulates DNA synthesis in cell-cell contact forming cardiomyocytes. J Cell Sci. 2004;117:507–514. doi: 10.1242/jcs.00889. [DOI] [PubMed] [Google Scholar]
  • [182].Giepmans BN, Hengeveld T, Postma FR, Moolenaar WH. Interaction of c-Src with gap junction protein connexin-43. Role in the regulation of cell-cell communication. J Biol Chem. 2001;276:8544–8549. doi: 10.1074/jbc.M005847200. [DOI] [PubMed] [Google Scholar]
  • [183].Solan JL, Lampe PD. Connexin 43 in LA-25 cells with active v-src is phosphorylated on Y247, Y265, S262, S279/282, and S368 via multiple signaling pathways. Cell Commun Adhes. 2008;15:75–84. doi: 10.1080/15419060802014016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [184].Kanemitsu My, Loo Lwm, Simon S, Lau Af, Eckhart W. Tyrosine Phosphorylation Of Connexin 43 By V-Src Is Mediated By Sh2 and Sh3 Domain Interactions. Journal of Biological Chemistry. 1997;272:22824–22831. doi: 10.1074/jbc.272.36.22824. [DOI] [PubMed] [Google Scholar]
  • [185].Leykauf K, Salek M, Bomke J, Frech M, Lehmann WD, Durst M, Alonso A. Ubiquitin protein ligase Nedd4 binds to connexin43 by a phosphorylation-modulated process. J Cell Sci. 2006;119:3634–3642. doi: 10.1242/jcs.03149. [DOI] [PubMed] [Google Scholar]
  • [186].Cooper CD, Lampe PD. Casein kinase 1 regulates connexin-43 gap junction assembly. J Biol Chem. 2002;277:44962–44968. doi: 10.1074/jbc.M209427200. [DOI] [PubMed] [Google Scholar]
  • [187].Solan JL, Lampe PD. Key connexin 43 phosphorylation events regulate the gap junction life cycle. J Membr Biol. 2007;217:35–41. doi: 10.1007/s00232-007-9035-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [188].Solan JL, Marquez-Rosado L, Sorgen PL, Thornton PJ, Gafken PR, Lampe PD. Phosphorylation at S365 is a gatekeeper event that changes the structure of Cx43 and prevents down-regulation by PKC. J Cell Biol. 2007;179:1301–1309. doi: 10.1083/jcb.200707060. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [189].Lampe PD, TenBroek EM, Burt JM, Kurata WE, Johnson RG, Lau AF. Phosphorylation of connexin43 on serine368 by protein kinase C regulates gap junctional communication. J Cell Biol. 2000;149:1503–1512. doi: 10.1083/jcb.149.7.1503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [190].Yogo K, Ogawa T, Akiyama M, Ishida-Kitagawa N, Sasada H, Sato E, Takeya T. PKA implicated in the phosphorylation of Cx43 induced by stimulation with FSH in rat granulosa cells. J Reprod Dev. 2006;52:321–328. doi: 10.1262/jrd.17107. [DOI] [PubMed] [Google Scholar]
  • [191].Saez Jc, Nairn Ac, Czernik Aj, Fishman Gi, Spray Dc, Hertzberg El. Phosphorylation Of Connexin43 and the Regulation Of Neonatal Rat Cardiac Myocyte Gap Junctions. Journal of Molecular & Cellular Cardiology. 1997;29:2131–2145. doi: 10.1006/jmcc.1997.0447. [DOI] [PubMed] [Google Scholar]
  • [192].Park DJ, Freitas TA, Wallick CJ, Guyette CV, Warn-Cramer BJ. Molecular dynamics and in vitro analysis of Connexin43: A new 14-3-3 mode-1 interacting protein. Protein Sci. 2006;15:2344–2355. doi: 10.1110/ps.062172506. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [193].Fu CT, Bechberger JF, Ozog MA, Perbal B, Naus CC. CCN3 (NOV) interacts with connexin43 in C6 glioma cells: possible mechanism of connexin-mediated growth suppression. J Biol Chem. 2004;279:36943–36950. doi: 10.1074/jbc.M403952200. [DOI] [PubMed] [Google Scholar]
  • [194].Liu L, Li Y, Lin J, Liang Q, Sheng X, Wu J, Huang R, Liu S. Connexin43 interacts with Caveolin-3 in the heart. Mol Biol Rep. 2009;37:1685–91. doi: 10.1007/s11033-009-9584-5. [DOI] [PubMed] [Google Scholar]
  • [195].Langlois S, Cowan KN, Shao Q, Cowan BJ, Laird DW. Caveolin-1 and -2 interact with connexin43 and regulate gap junctional intercellular communication in keratinocytes. Mol Biol Cell. 2008;19:912–928. doi: 10.1091/mbc.E07-06-0596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [196].Butkevich E, Hulsmann S, Wenzel D, Shirao T, Duden R, Majoul I. Drebrin is a novel connexin-43 binding partner that links gap junctions to the submembrane cytoskeleton. Curr Biol. 2004;14:650–658. doi: 10.1016/j.cub.2004.03.063. [DOI] [PubMed] [Google Scholar]

RESOURCES