ArticlePDF Available

Zonula Occludens-1 Alters Connexin43 Gap Junction Size and Organization by Influencing Channel Accretion

Authors:

Abstract and Figures

Regulation of gap junction (GJ) organization is critical for proper function of excitable tissues such as heart and brain, yet mechanisms that govern the dynamic patterning of GJs remain poorly defined. Here, we show that zonula occludens (ZO)-1 localizes preferentially to the periphery of connexin43 (Cx43) GJ plaques. Blockade of the PDS95/dlg/ZO-1 (PDZ)-mediated interaction between ZO-1 and Cx43, by genetic tagging of Cx43 or by a membrane-permeable peptide inhibitor that contains the Cx43 PDZ-binding domain, led to a reduction of peripherally associated ZO-1 accompanied by a significant increase in plaque size. Biochemical data indicate that the size increase was due to unregulated accumulation of gap junctional channels from nonjunctional pools, rather than to increased protein expression or decreased turnover. Coexpression of native Cx43 fully rescued the aberrant tagged-connexin phenotype, but only if channels were composed predominately of untagged connexin. Confocal image analysis revealed that, subsequent to GJ nucleation, ZO-1 association with Cx43 GJs is independent of plaque size. We propose that ZO-1 controls the rate of Cx43 channel accretion at GJ peripheries, which, in conjunction with the rate of GJ turnover, regulates GJ size and distribution.
Content may be subject to copyright.
Molecular Biology of the Cell
Vol. 16, 5686–5698, December 2005
Zonula Occludens-1 Alters Connexin43 Gap Junction Size
and Organization by Influencing Channel Accretion
Andrew W. Hunter, Ralph J. Barker, Ching Zhu, and Robert G. Gourdie
Department of Cell Biology and Anatomy, Cardiovascular Developmental Biology Center, Medical University
of South Carolina, Charleston, SC 29425
Submitted August 9, 2005; Revised September 13, 2005; Accepted September 15, 2005
Monitoring Editor: Asma Nusrat
Regulation of gap junction (GJ) organization is critical for proper function of excitable tissues such as heart and brain, yet
mechanisms that govern the dynamic patterning of GJs remain poorly defined. Here, we show that zonula occludens
(ZO)-1 localizes preferentially to the periphery of connexin43 (Cx43) GJ plaques. Blockade of the PDS95/dlg/ZO-1
(PDZ)-mediated interaction between ZO-1 and Cx43, by genetic tagging of Cx43 or by a membrane-permeable peptide
inhibitor that contains the Cx43 PDZ-binding domain, led to a reduction of peripherally associated ZO-1 accompanied by
a significant increase in plaque size. Biochemical data indicate that the size increase was due to unregulated accumulation
of gap junctional channels from nonjunctional pools, rather than to increased protein expression or decreased turnover.
Coexpression of native Cx43 fully rescued the aberrant tagged-connexin phenotype, but only if channels were composed
predominately of untagged connexin. Confocal image analysis revealed that, subsequent to GJ nucleation, ZO-1 associ-
ation with Cx43 GJs is independent of plaque size. We propose that ZO-1 controls the rate of Cx43 channel accretion at
GJ peripheries, which, in conjunction with the rate of GJ turnover, regulates GJ size and distribution.
INTRODUCTION
The gap junction (GJ) is a plaque-like aggregate of intercel-
lular channels that facilitates cytoplasmic interchange of
ions, second messengers, and other molecules 1 kDa be-
tween cells (Goodenough et al., 1996). Frequent and variably
sized GJ channel aggregates couple most cells in animal
tissues. In excitable organs such as the heart and brain, GJs
show distinctive organizational patterns that configure ex-
tended intercellular pathways for stable and long-term
propagation of action potential (Lo, 2000). The channels
comprising individual GJ plaques are composed of proteins
encoded by the connexin family of genes (Willecke et al.,
2002). Assembly of GJs from connexin monomers is thought
to proceed in a multistep process (Musil and Goodenough,
1991, 1993). First, six connexins oligomerize into a
hemichannel, called a connexon, followed by trafficking to
the plasma membrane. Subsequently, a connexon docks
with a second connexon from the apposed membrane of an
adjacent cell to form an intercellular channel. In a process
that occurs either simultaneous with or after this docking
step, channels aggregate to form the functional organelle of
cell-cell communication—the GJ plaque.
The gating of single channels within a GJ plaque is regu-
lated by various stimuli, including voltage, pH, and phos-
phorylation (Saez et al., 2003). Intercellular communication is
also regulated at the level of the plaque as a whole by factors
that affect the abundance, size, and cellular distribution of GJ
channel aggregates (Hall and Gourdie, 1995; Bukauskas et
al., 2000; Spach et al., 2000; Johnson et al., 2002; Lauf et al.,
2002). Irregularities in the extent and geometry of gap junc-
tional contacts have been implicated in cardiac and neural
electrophysiological pathologies in humans, including isch-
emic heart disease (Smith et al., 1991), hypertrophic cardio-
myopathy (Sepp et al., 1996), heart failure (Dupont et al.,
2001), and medial temporal lobe epilepsy (Fonseca et al.,
2002). At present, the mechanisms governing higher order
aspects of GJ organization and function in health and dis-
ease are poorly understood.
Nonetheless, there is a growing understanding that an
array of proteins interact with connexins, potentially medi-
ating linkage of GJs to other junction types, signal transduc-
tion molecules, and the cytoskeleton (Giepmans, 2004). One
connexin-interacting molecule that has received particular
attention is zonula occludens (ZO)-1. Originally discovered
in association with tight junctions (Stevenson et al., 1986),
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 po-
larity (Anderson, 1996). MAGUKs, such as ZO-1, synaptic
protein PSD95, and the Drosophila tumor suppressor discs
large (dlg), are characterized by an N-terminal array of
protein-binding domains that includes one or more PDS95/
dlg/ZO-1 (PDZ) domains. Previous immunoprecipitation
and yeast two-hybrid studies showed that a short PDZ-
binding motif at the C terminus of connexin43 (Cx43) inter-
acts with the second of three PDZ domains in ZO-1 (Giep-
mans and Moolenaar, 1998; Toyofuku et al., 1998).
Subsequently, other connexins have been reported to inter-
act with ZO-1 via a consensus C-terminal PDZ-binding do-
main similar to that of Cx43 (Jin et al., 2000; Kausalya et al.,
2001; Laing et al., 2001; Nielsen et al., 2003; Li et al., 2004).
Initially, ZO-1 function at GJs was assumed to be analo-
gous to its presumed role at tight junctions. Namely, ZO-1
This article was published online ahead of print in MBC in Press
(http://www.molbiolcell.org/cgi/doi/10.1091/mbc.E05–080737)
on September 29, 2005.
Address correspondence to: Andrew W. Hunter (huntera@
musc.edu) or Robert G. Gourdie (gourdier@musc.edu).
Abbreviations used: AJ, adherens junction; Cx43, connexin43; GJ,
gap junction; MOI, multiplicity of infection; PDZ, PDS95/dlg/ZO-1;
ZO, zonula occludens.
5686 © 2005 by The American Society for Cell Biology
was envisaged as a passive scaffolding molecule that stabi-
lizes GJs through cytoskeletal anchoring. However, subse-
quent reports have indicated that interactions between con-
nexins and ZO-1 may encompass other, more dynamic
functions. In particular, changes in Cx43–ZO-1 interaction
have been noted during remodeling of GJs in cardiomyo-
cytes and other cell types (Defamie et al., 2001; Barker et al.,
2002; Segretain et al., 2004). Moderation of Cx43–ZO-1 inter-
action by c-Src has lead others to hypothesize a role for ZO-1
in the regulation of GJ channel activity (Giepmans et al.,
2001a; Toyofuku et al., 2001; Duffy et al., 2004; Sorgen et al.,
2004). As such, definitive roles for ZO-1 interaction with
connexins remain to be determined. Here, we provide evi-
dence that ZO-1 regulates the cellular distribution of Cx43,
and consequently the size of GJ plaques, by controlling the
rate of channel accretion at plaque perimeters.
MATERIALS AND METHODS
Animals
Hearts were harvested from 2- to 3-d-old Holtzman Sprague Dawley rats.
Animal care was in accordance with institutional guidelines at the Medical
University of South Carolina, Charleston, SC (Animal welfare assurance
#A4328-01).
Antennapedia Peptides
Antennapedia peptides (Lindgren et al., 2000) were generated at Medical
University of South Carolina or by Research Genetics (Huntsville, AL). The
inhibitor peptide includes the Cx43 C-terminal amino acids 374–382 (RPRP-
DDLEI) that encompass the ZO-1-binding sequence. A control peptide was
generated by reversing the Cx43 amino acid sequence (IELDDPRPR). Both
peptides contain a 16-amino acid antennapedia internalization vector (RQP-
KIWFPNRRKPWKK) linked to the N terminus of the Cx43 (or reversed Cx43)
sequence; peptides were N-terminally biotinylated.
Cell Culture
Neonatal cardiomyocytes were isolated with a Neonatal Cardiomyocyte Iso-
lation System (LK003300; Worthington Biochemicals, Freehold, NJ) according
to manufacturer’s instructions. Cultures were generated from 8 to 12 hearts.
Dispersed cardiomyocytes were plated on coverglass or culture dishes coated
with 0.2% gelatin (Sigma-Aldrich, St. Louis, MO). Cytosine arabinoside (Sig-
ma-Aldrich) was used to inhibit fibroblast proliferation. Cardiomyocytes
were cultured as described previously (Barker et al., 2002) as monolayers for
72 h before peptide treatments.
HeLa cells stably expressing either native Cx43 or Cx43-EGFP (Jordan et al.,
1999) were cultured as described previously (Hunter et al., 2003) for 2448 h
before peptide treatments or adenoviral infection. Newly confluent cultures
were used for all experiments.
Adenovirus expressing human Cx43 (Qbiogene, Carlsbad, CA) was added
to cells (at various multiplicities of infection [MOIs]) in Opti-MEM and
allowed to adsorb for several hours before addition of culture medium.
Adenoviral expression progressed for 24 h subsequent to further manipula-
tions.
Peptides were added to cells at 30
M, which falls within the effective range
(1–100
M) for antennapedia peptides (Lindgren et al., 2000). Media contain-
ing peptides were replenished every 24 h.
Determination of Connexin Expression Levels,
Stoichiometry, and Turnover
Cx43 expression was assessed in cardiomyocytes and HeLa Cx43 cells cul-
tured for 48 and 72 h, respectively, with or without peptides. Cells were
solubilized in hot 4% SDS sample buffer for 5 min, sheared, and boiled for 5
min. Samples were resolved by 10% SDS-PAGE and immunoblotted with
Cx43 (71-0700; Zymed Laboratories, South San Francisco, CA) and glyceral-
dehyde-3-phosphate dehydrogenase (GAPDH) (Research Diagnostics,
Flanders, NJ) antibodies. Sample loading was normalized to GAPDH signal.
To determine stoichiometry and turnover in connexin-expressing HeLa
cells, metabolic labeling with [
35
S]methionine (35–100
Ci/ml; PerkinElmer
Life and Analytical Sciences, Boston, MA), pulse-chase analysis, and immu-
noprecipitations using Cx43 antibodies (C6219; Sigma-Aldrich) were per-
formed as described previously (VanSlyke and Musil, 2000). Immunoprecipi-
tated Cx43 and Cx43-GFP were resolved by 10% SDS-PAGE. Dried gels were
exposed to phosphor screens;
35
S-labeled connexins were detected using a
Storm PhosphorImager and quantified using ImageQuant (Molecular Dy-
namics, Sunnyvale, CA). Curve fitting of pulse-chase data were performed
with IGOR Pro (Wavemetrics, Lake Oswego, OR).
Pull-Down Assays
Glutathione S-transferase (GST)-fusion proteins composed of either the PDZ1
(amino acids 1–106) or PDZ2 (amino acids 173–261) domain of human ZO-1
were generated as described previously (Nielsen et al., 2002) and isolated from
isopropyl
-d-thiogalactoside-induced DH5
bacteria according to standard
procedures.
HeLa Cx43 cells were lysed in 50 mM Tris-HCl, pH 7.4, 150 mM NaCl, 2
mM EGTA, 1% NP-40, 0.25% Na-deoxycholate, 100
M phenylmethylsulfonyl
fluoride (PMSF), and 1 Complete protease inhibitors (Roche Diagnostics,
Indianapolis, IN) for 30 min at 4°C and then centrifuged at 16,000 g for 10
min at 4°C. Fusion proteins (2–3
g) coupled to glutathione-Sepharose 4B
beads (GE healthcare, Little Chalfont, Buckinghamshire, United Kingdom)
were added to clarified lysates containing peptides (100
M). Reactions were
incubated sequentially with GST-PDZ1 and GST-PDZ2 beads, each for1hat
4°C. Pelleted beads were washed three times with lysis buffer. Pelleted
material was resolved by 10% (Cx43) and 10 –20% (peptides) SDS-PAGE.
Western blot detection was performed with streptavidin-horseradish perox-
idase (HRP) (GE Healthcare) and Cx43 antibodies (C6219; Sigma-Aldrich).
Triton X-100 Fractionation Assay
Peptide-treated HeLa Cx43 cells were lysed in 1% Triton X-100 in phosphate-
buffered saline (PBS) plus 100
M PMSF and 1 Complete protease inhibitors
for1honicewith occasional vortexing and then centrifuged at 10,000 g for
5 min at 4°C. Triton-soluble fractions (supernatants) were removed, and
Triton-insoluble fractions (pellets) were resuspended in an equal volume of
lysis buffer. Equal volumes of Triton-soluble and -insoluble fractions were
resolved by 10% SDS-PAGE and immunoblotted with Cx43 antibody (71-
0700; Zymed Laboratories). Signal was detected by alkaline phosphatase-
based chemiluminescence (CDP-Star; Tropix, Bedford, MA) exposed to Hy-
perfilm ECL (GE Healthcare). Blots were digitized using a UMAX PowerLook
scanner and VueScan (Hamrick Software, Phoenix, AZ). Quantitative densi-
tometry was performed with ImageJ (National Institutes of Health; http://
rsb.info.nih.gov/ij/); soluble/insoluble ratios were calculated from area-un-
derpeak measurements; insoluble Cx43 isoform ratios were calculated from
peak height measurements.
Fluorescence Labeling
Cells grown on coverglass were fixed in 2% paraformaldehyde in PBS for 10
min, blocked with 1% bovine serum albumin, 0.1% Triton X-100 in PBS, and
immunolabeled with Cx43 (610062; BD Transduction Laboratories, Lexington,
KY; MAB3067; Chemicon International, Temecula, CA) and ZO-1 (61-7300;
Zymed Laboratories) antibodies. Signal was detected directly by green fluo-
rescent protein (GFP) fluorescence or by secondary antibodies conjugated
with Alexa488, Alexa546 (Invitrogen, Carlsbad, CA), or Cy5 (Jackson Immu-
noResearch Laboratories, West Grove, PA). Biotinylated peptides were de-
tected using streptavidin-Cy5 (Jackson ImmunoResearch Laboratories). To
delineate plasma membrane, tetramethylrhodamine B isothiocyanate-wheat
germ agglutinin (TRITC-WGA) (Sigma-Aldrich) was added (5
g/ml) to live
HeLa Cx43 cultures for 5 min before fixation. Samples were mounted in
SlowFade (Invitrogen).
Confocal Microscopy
Optical sections were captured with a TCS SP2 AOBS laser scanning confocal
microscope equipped with a 63/1.4 numerical aperture oil immersion ob-
jective (Leica Microsystems, Deerfield, IL). Lasers used were Ar (488 nm), Kr
(568 nm), and HeNe (633 nm). Pinhole (airy disk), gain, and black level
settings were held constant. Potential overlaps in emission spectra were
eliminated by sequential scanning and tuning of the AOBS. For quantitative
analyses of single optical sections, scan format and zoom were adjusted to
give x-y pixel sizes of 68 68 nm for myocyte images and 116 116 nm for
HeLa cell images. Maximum projections were generated from z-series with
x-y-z voxel dimensions of 68 68 400–730 nm for myocyte images and
124 124 124 nm for HeLa cell images.
Image Analysis
GJ lengths, GJ number, total Cx43 area, and individual plaque areas were
measured from single optical sections as described previously (Green et al.,
1993; Hunter et al., 2003). HeLa cell analysis was modified by limiting mea-
surements to Cx43 signal colocalized with TRITC-WGA signal at cell–cell
borders. Parameter means were calculated from image averages, except in the
case of GJ mean lengths and size distributions, which were determined from
the entire sampled populations. Quantification of ZO-1 colocalization with
individual Cx43 plaques was carried out according to Zhu et al. (2005). Linear
regression analysis of Cx43–ZO-1 colocalization was performed in IGOR Pro.
Statistics
Multiple data sets were compared using analysis of variance (p 0.05
rejected as not significant) and unpaired, two-tailed Bonferroni t tests (p
0.05/3 0.017 rejected as not significant). Size distributions were compared
using nonparametric chi-square and Mann-Whitney rank sum tests. Where
ZO-1 Affects Cx43 Gap Junction Dynamics
Vol. 16, December 2005 5687
appropriate, asterisks (*) indicate significant differences between inhibitor and
no peptide treatments; hash marks (#) indicate significant differences between
inhibitor and reverse peptide treatments.
RESULTS
Native Cx43 Rescues the Abnormal Cx43-GFP Gap
Junction Phenotype in HeLa Cells
HeLa cells do not express significant levels of any known
connexin and thus are considered communication deficient.
Previously, we showed that HeLa cells stably expressing a
GFP-tagged Cx43 form sheetlike GJs of abnormally large
size (Hunter et al., 2003); by contrast, stable expression of
untagged native Cx43 resulted in small, discontinuous GJs
uniformly distributed at cell–cell interfaces (Hunter et al.,
2003). From these observations, we concluded that fusion of
GFP to the C terminus of Cx43 disrupts regulatory mecha-
nisms that control GJ size and cellular distribution.
The fact that Cx43-GFP assembles into smaller GJs when
expressed in communication-competent cells (Jordan et al.,
1999; our unpublished data) suggested to us that the pres-
ence of untagged native Cx43 is sufficient to rescue the
abnormal GJ dynamics of Cx43-GFP. Consistent with our
hypothesis, adenoviral expression of native Cx43 in HeLa
Cx43-GFP cells restored GJs to normal size and organiza-
tional patterns (Figure 1). We quantified this effect by sys-
tematically modulating expression of the two Cx43 species
using varying MOIs of Cx43 adenovirus and then measuring
GJ lengths after 24 h of coexpression. As native Cx43 expres-
sion levels were increased in HeLa Cx43-GFP cells, the fre-
quency of smaller GJs also increased. Complete rescue of the
aberrant Cx43-GFP phenotype occurred at MOI 10 (Figure
1A), as assessed by comparison with the GJ length distribu-
tions in HeLa cells stably expressing only native Cx43
(Hunter et al., 2003) or expressing only native Cx43 after
adenoviral infection at MOI 10 (Figure 1A, bottom).
A shift toward smaller Cx43-GFP GJs with increased lev-
els of native Cx43 is reflected in a plot of mean GJ lengths
versus Cx43 adenovirus MOI (Figure 1B). Relative to the
mean length of pure Cx43-GFP GJs, the mean length in HeLa
Cx43-GFP cells coexpressing native Cx43 at MOI 10 was
reduced approximately threefold and was not significantly
different from the mean length in HeLa cells expressing only
native Cx43 at the same MOI (Figure 1B). This convergence
indicates that GJ size variation, and in particular the abnor-
mally large size of Cx43-GFP junctions in the absence of
native Cx43, is not a simple function of increasing Cx43
expression.
GJ Size Control Is Critically Dependent on Availability of
Free Cx43 C Termini within the Connexon Pool
The partial reduction in mean length observed at MOI 5
(Figure 1) indicates that Cx43-GFP GJ size is normalized
gradually after introduction of native Cx43 and suggests
that the actual amount of native Cx43 sufficient to rescue the
Cx43-GFP phenotype likely falls between that expressed by
viral MOIs of 5 and 10. But exactly how much native Cx43 is
Figure 1. Coexpression of native Cx43 in HeLa Cx43-GFP cells
normalizes gap junction size and distribution. (A) Confocal optical
sections of GFP fluorescence in HeLa Cx43-GFP cells coexpressing
increasing levels of native Cx43 (i.e., increasing viral MOI) and
corresponding length distributions generated from measurements
of membrane-localized Cx43-GFP plaques. Bottom, length distribu-
tion of Cx43 GJs after viral expression of native Cx43 at MOI 10 in
wild-type HeLa cells. Bar, 20
m. (B) Mean length of Cx43-GFP GJs
as a function of native Cx43 viral expression (mean SE). Coex-
pression of native Cx43 at MOI 10 resulted in a 3-fold reduction in
the mean length of mixed GJs relative to pure Cx43-GFP GJs (MOI
0). Asterisk (*) indicates that the GJ size distribution in HeLa Cx43-
GFP cells expressing native Cx43 at MOI 10 was significantly dif-
ferent (p 0.001) from the GJ distribution in HeLa Cx43-GFP cells
expressing native Cx43 at MOI 0 or 5, based on comparisons of GJ
populations (n 1000 for all conditions) using a Mann–Whitney
rank sum test. The difference between length distributions in HeLa
Cx43-GFP and wild-type HeLa at MOI 10 was not significant.
A. W. Hunter et al.
Molecular Biology of the Cell5688
produced at these MOIs? More precisely, how many func-
tional Cx43 C termini are sufficient to normalize GJ size and
organizational pattern in HeLa Cx43-GFP cells?
To quantify the amounts of virally expressed native Cx43
relative to stably expressed Cx43-GFP, HeLa Cx43-GFP cells
were metabolically labeled after infection with increasing
amounts of Cx43 adenovirus. After immunoprecipitation,
35
S-labeled connexins were resolved by SDS-PAGE, and the
relative amounts of the two connexin species were measured
by phosphorimaging. At MOI 10—the Cx43 viral expression
level sufficient for complete rescue of Cx43-GFP GJ size—we
found a native Cx43:Cx43-GFP stoichiometry of 5:1 (Figure
2A). This ratio represents a bulk distribution in rescued cells
of 80% native Cx43 versus 20% Cx43-GFP (Figure 2B),
which is consistent with a population of connexons that on
average contain one or fewer Cx43-GFP molecules.
Three lines of evidence argue in favor of random mixing
of native Cx43 and Cx43-GFP into heteromeric connexons.
First, if tagged and untagged Cx43 segregated exclusively
into homomeric connexons, then we might expect to see
regions of Cx43 immunofluorescence devoid of GFP fluores-
cence; instead, Cx43 immunofluorescence and GFP fluores-
cence displayed nearly identical, overlapping patterns in
cells coexpressing native Cx43 and Cx43-GFP (Figure 2C;
Jordan et al., 1999). Second, motifs governing Cx43 oligomer-
ization are located outside of the cytoplasmic C-terminal
domain (Falk et al., 1997; Maza et al., 2005). And third,
C-terminally tagged Cx43 constructs have been shown to
co-oligomerize with each other and with untagged Cx43
(Lauf et al., 2001; Sarma et al., 2002). Based on knowledge of
the relative expression levels of the two Cx43 species (Figure
2, A and B), and assuming random codistribution of con-
nexins within connexons (Figure 2C), a probability equation
(Burt et al., 2001) was used to generate the expected fre-
quency distributions of Cx43:Cx43-GFP connexon stoichio-
metries (Figure 2D). The probabilistic connexon distribu-
tions suggest that when native Cx43 is expressed in HeLa
Cx43-GFP cells at levels both necessary and sufficient for
normalization of GJ organization (i.e., Cx43 viral expression
at MOIs 5–10; Figure 1), the majority of connexons (65%)
must be composed predominately of native Cx43 (4 per
connexon); furthermore, nearly all connexons (99%) must
contain at least one native Cx43 (Figure 2D). Clearly, a
channel population comprised of connexons with a single
intact Cx43 C terminus is not sufficient to reestablish GJ size
control. A conservative interpretation is that connexons con-
taining even a single Cx43-GFP cannot contribute effectively
to regulation of GJ size.
C-terminal Tagging of Cx43 Does Not Interfere with Gap
Junction Turnover
Connexins turn over at faster rates (t
1/2
1–5 h; Fallon and
Goodenough, 1981; Musil et al., 1990; Laird et al., 1991;
Beardslee et al., 1998) than most membrane proteins (t
1/2
24 h; Hare and Taylor, 1991). One explanation for the for-
mation of large Cx43-GFP GJs is that they are resistant to
degradation and therefore turnover more slowly than GJs
composed of native Cx43. We tested this possibility using
pulse-chase assays of HeLa cells expressing either native
Cx43 or Cx43-GFP, or both proteins (Figure 3). Short pulse
labeling (40 min) of connexin-expressing cells indicated that
native Cx43 and Cx43-GFP turnover at similar rates, with
half-lives (t
1/2
2–3 h) in the range expected for connexins
(Figure 3, A and B).
Previous studies argue against the formation of a meta-
bolically stable pool of Cx43 in GJ plaques (Laird et al., 1995;
Musil et al., 2000). However, if GFP-tagged Cx43 were to
segregate into pools with vastly different turnover kinetics
(e.g., stable junctional versus labile nonjunctional Cx43-
GFP), then short-interval metabolic labeling would not pen-
etrate preexisting, long-lived GJs. We therefore performed
long pulse labeling (15 h) to ensure more uniform
35
S-
labeling in the event that the turnover rate of plaque-incor-
porated Cx43-GFP was appreciably slower than that of non-
junctional Cx43-GFP. However, even after protracted
radiolabeling, the kinetics of Cx43-GFP turnover was not
notably different from those of native Cx43 (Figure, 3, C and
D). In fact, when expressed in HeLa cells, the half-life of
Cx43-GFP (t
1/2
2 h) seemed to be consistently shorter than
that of native Cx43 (t
1/2
3–5 h; Figure 3, A–D)—a result
completely at odds with the hypothesis that the abnormally
large Cx43-GFP GJs arise from resistance to degradation.
When native Cx43 was coexpressed in HeLa Cx43-GFP
cells, the turnover rate of both Cx43 species fell in the range
of rates measured when native Cx43 was expressed alone
(Figure 3E). Thus, the half-life of Cx43-GFP was increased
slightly in the presence of native Cx43 at levels sufficient to
normalize GJ size. The convergence of the half-life of Cx43-
GFP to that of native Cx43 in coexpressing cells provides
further evidence of random coassembly of native Cx43 and
Cx43-GFP into heteromeric connexons. More importantly,
the increase in Cx43-GFP half-life is opposite to what might
have been expected if coexpression of native Cx43 rescued
GJ size in HeLa Cx43-GFP cells by increasing the turnover
rate of Cx43-GFP. These observations confirm that C-termi-
nal tagging of Cx43 does not interfere with GJ turnover and
allow us to draw an important conclusion: the abnormal size
Figure 2. Reduction of Cx43-GFP plaque size by coexpression of
native Cx43 requires a majority of connexons be composed predom-
inately of native Cx43. (A) Representative gel showing changes in
expression levels of Cx43-GFP and native Cx43 in HeLa Cx43-GFP
cells after infection with increasing Cx43 adenovirus MOI. Connexin
stoichiometries (mean SE; n 3) at each MOI were determined by
measuring band intensities with a PhosphorImager. (B) Connexin
stoichiometries presented as bulk expression levels of Cx43-GFP
and native Cx43 relative to total connexin. (C) Representative sets of
confocal images showing Cx43-GFP and native Cx43 codistribute in
HeLa Cx43-GFP cells coexpressing Cx43 at viral MOI 10. Note the
lack of Cx43 domains devoid of GFP fluorescence. Bar, 10
m. (D)
Distribution of connexon stoichiometries expected when native
Cx43 is coexpressed in HeLa Cx43-GFP at levels necessary (MOI 5)
and sufficient (MOI 10) to normalize GJ size and distribution.
ZO-1 Affects Cx43 Gap Junction Dynamics
Vol. 16, December 2005 5689
of Cx43-GFP GJs cannot be explained by the formation of
degradation-resistant plaques.
A Peptide Inhibitor of Cx43–ZO-1 Interaction Based on
the PDZ-binding Domain of Cx43
Our results with Cx43-GFP suggested that tagging of Cx43
interfered with a regulatory mechanism that controls GJ
size. C-terminal tagging of Cx43, including with GFP, has
been shown by immunoprecipitation to profoundly disrupt
binding of ZO-1 (Giepmans and Moolenaar, 1998; Giepmans
et al., 2001b), suggesting a role for ZO-1 in regulating the size
and organization of Cx43 GJs. However, manipulation of
Cx43 through genetic tagging might affect the binding of
proteins other than ZO-1. Therefore, to directly target ZO-1
activity, we synthesized a peptide comprising the PDZ-
binding domain of Cx43 linked to an antennapedia internal-
ization sequence (Figure 4A).
At concentrations between 10 and 100
M, antennapedia
peptides were detected at relatively uniform levels in cells
up to 24 h after addition to culture media (Figures 4B and 5,
E–F and H–I). Once inside cells, we hypothesized that the
inhibitor peptide should compete for binding to ZO-1,
thereby reducing its availability for interaction with the
PDZ-binding domain of endogenous Cx43. As a control
peptide, we used the same antennapedia sequence but re-
versed the Cx43 PDZ-binding sequence; thus, the reverse
control peptide no longer has a free C-terminal isoleucine
(Figure 4A), a Cx43 residue necessary for interaction with
the second PDZ domain in ZO-1 (Giepmans and Moolenaar,
1998). In accord with our expectations, the inhibitor peptide
but not the reverse peptide was pulled down by a recombi-
nant polypeptide comprising the PDZ2 domain of ZO-1
(Figure 4C). The specificity of this interaction was corrobo-
rated by the observation that the PDZ1 domain of ZO-1 did
not pull down the inhibitor or reverse peptide (Figure 4C).
Having demonstrated specificity of targeting to the PDZ2
domain, we next tested whether the inhibitor peptide dis-
rupted ZO-1 interaction with Cx43. Inhibitor or control pep-
tides were added to lysates from HeLa Cx43 cells and then
either the PDZ1 or PDZ2 domains of ZO-1 were used to pull
down Cx43. Consistent with interference of Cx43–ZO-1
binding, the amount of Cx43 pulled down by PDZ2 in the
presence of inhibitor peptide was significantly reduced rel-
ative to Cx43 pulled down in the presence of control peptide
or no peptide (Figure 4D). As expected, no Cx43 was pulled
down by PDZ1 irrespective of the presence or absence of
peptides. Based on these data, we conclude that the inhibitor
peptide disrupts the interaction between Cx43 and ZO-1.
Disruption of Cx43–ZO-1 Interaction Increases the Size of
Cx43 GJs
To characterize the effects of the inhibitor peptide on GJs,
HeLa Cx43 cells were treated with the inhibitor, reverse, or
no peptide for 72 h. At the end of the culture period, living
cells were first exposed to TRITC-WGA to label cell mem-
branes and then fixed and immunolabeled for Cx43 (Figure
5, A–F). As shown qualitatively in Figure 5, A–C, exposure
to the inhibitor peptide resulted in increased levels of Cx43
at WGA-defined cell membranes relative to controls. In-
creased cell border-localized Cx43 also was observed in
primary cultures of neonatal cardiomyocytes treated with
inhibitor peptide (Figure 5I).
Quantitatively, treatment with inhibitor peptide increased
the average dimensions of individual Cx43 plaques in both
HeLa Cx43 cells and neonatal cardiomyocytes relative to
controls (Figure 6A). This increase in mean GJ size in re-
sponse to inhibitor peptide was highly significant with
larger GJs occurring at greater frequency (at the expense of
smaller GJs) in both cell types (Figure 6B). In addition to
increased GJ size, the total area of membrane-localized
plaques in HeLa Cx43 cells was significantly increased by
exposure to inhibitor peptide relative to no peptide controls
(Figure 6C); and consistent with increased plaque fusion (or
decreased fission) events, HeLa Cx43 cells showed a slight
reduction in the number of membrane-localized GJs in in-
hibitor-treated versus control cells (Figure 6D), although we
could not establish this as a definitive trend.
Figure 3. Native Cx43 and Cx43-GFP turnover at similar rates in
HeLa cells. HeLa cells expressing native Cx43 (A and C) or Cx43-
GFP (B and D) or both connexin species (E) were pulse labeled with
[
35
S]methionine for 40 min (A, B, and E) or 15 h (C and D),
followed by chase with unlabeled methionine. At the designated
times during the chase period,
35
S-labeled connexins were quanti
-
tatively immunoprecipitated from HeLa cells lysates and resolved
using SDS-PAGE. Band intensities were measured with a Phospho-
rImager. Plots show levels of
35
S-labeled connexins remaining, nor
-
malized to the amount of
35
S-labeled connexin at time zero, as a
function of chase time (mean SD; n 2). Curves are exponential
fits to the data (weighted by their uncertainties) which yielded the
half-life (t
1/2
)of
35
S-labeled connexins (mean SE).
A. W. Hunter et al.
Molecular Biology of the Cell5690
We can rule out the possibility that the observed increases
in plaque size (Figure 6A) and overall levels of immunode-
tectable Cx43 GJs at cell–cell borders (Figure 6C) were due
simply to an increase in Cx43 expression, because the inhib-
itor peptide did not significantly affect the total amount of
Cx43 in HeLa cells or cardiomyocytes (Figure 6E). This
agrees with the recent observation that Cx43 abundance was
unaltered after disruption of Cx43–ZO-1 interaction in os-
teoblastic cells (Laing et al., 2005). Moreover, this accords
with our finding that the size of Cx43-GFP plaques was
reduced after increasing the expression of native Cx43 (Fig-
ures 1 and 2).
Collectively, the data presented in Figure 6 suggest that
the observed increase in GJ plaque size in the presence of the
inhibitor peptide is due mainly to unregulated accretion of
connexons from nonjunctional pools, with a possible contri-
bution from increased plaque fusions.
Peptide Inhibition of Cx43–ZO-1 Interaction Leads to
Redistribution of Cx43 from Nonjunctional Pools to GJ
Plaques
Unaltered levels of total Cx43 in peptide-treated cells (Figure
6E) suggested that, rather than accumulation of newly up-
regulated Cx43, the increased dimensions of individual
Cx43 plaques at cell borders might be due to a redistribution
of preexisting Cx43. We tested this possibility using a bio-
chemical fractionation assay that distinguishes junctional
from nonjunctional connexins. HeLa Cx43 cells were treated
with peptides for 72 h and then lysed in ice-cold buffer
containing 1% Triton X-100. Under these conditions non-
junctional connexins are solubilized, whereas GJ plaques
remain insoluble (Musil and Goodenough, 1991). After lysis,
the two populations of connexins were fractionated by cen-
trifugation and resolved using SDS-PAGE. The Triton-solu-
ble fraction resolves predominately as an indistinguishable
collection of faster-migrating isoforms (NP/P0) that repre-
sent nonjunctional Cx43, which is largely (although not ex-
clusively) unphosphorylated (Figure 7A; Musil and Good-
enough, 1991; Solan et al., 2003). In contrast, the Triton-
insoluble gap junctional fraction resolves as at least three
distinct Cx43 isoforms, NP/P0 and two phosphorylated spe-
cies, P1 and P2 (Figure 7A).
Generally, it is thought that unphosphorylated Cx43 as-
sembles into connexons en route to the plasma membrane
(Musil and Goodenough, 1991; Musil and Goodenough,
1993). Phosphorylation events that result in gel mobility
shifts occur subsequent to delivery to the plasma membrane
and are closely linked to accretion of connexons into Triton-
insoluble GJ plaques (Musil and Goodenough, 1991). The
banding patterns shown in Figure 7A follow the well-estab-
lished relationship between detergent solubility, gel mobil-
ity, and phosphorylation state with respect to the life cycle of
connexins (Musil and Goodenough, 1991). Band intensity
profiles reveal a shift in the relative amounts of the various
Cx43 isoforms after treatment with inhibitor peptide (Figure
7B). Quantitative densitometry showed that the inhibitor
peptide caused a significant reduction in the level of soluble
relative to insoluble Cx43 (Figure 7C), which suggests
that the amount of endogenous Cx43 pulled down from HeLa
lysates by GST-PDZ2 beads was significantly reduced by the pres-
ence of inhibitor peptide. Peptide inhibition of PDZ2-Cx43 interac-
tion was reproducible (n 3). Comparison of nonspecific GST
fusion protein band intensities confirms equal loading across con-
ditions.
Figure 4. A peptide inhibitor that binds specifically to the second
PDZ domain of ZO-1 and blocks interaction with Cx43. (A) Domain
organization of full-length Cx43 and antennapedia peptides. (B)
Detection of biotinylated peptide (0–100
M) in HeLa cells after 2 h.
Streptavidin-Cy5 fluorescence intensity shows a dose-dependent
increase with increasing peptide concentration. Cell nuclei are
stained with DAPI. Bar, 15
m. (C) Western blots showing that
inhibitor and reverse peptides (3.5 kDa) both were detected by
streptavidin-HRP (top blot), but only inhibitor peptide was detected
by an antibody that recognizes the C terminus of Cx43 (bottom blot).
Only inhibitor peptide was pulled down by purified ZO-1 GST-
PDZ2 fusion protein-coupled beads; neither peptide was pulled
down by GST-PDZ1 beads. (D) Representative Cx43 blot showing
ZO-1 Affects Cx43 Gap Junction Dynamics
Vol. 16, December 2005 5691
greater incorporation of connexons into GJ plaques. Further-
more, within the insoluble fraction, we detected a significant
increase in the amounts of the P1 and P2 phosphoisoforms
relative to controls (Figure 7D). The increases in Cx43-P1
and -P2 seemed to be linked because the ratio of P2/P1 was
not affected by peptide treatment (Figure 7D). The increase
in Triton-insoluble Cx43, in conjunction with the appearance
of elevated levels of phosphorylated Cx43 species known to
populate mature GJs, supports the idea that the inhibitor
peptide increases GJ size by promoting the accretion of
nonjunctional connexons into gap junctional plaques.
ZO-1 Localization at GJ Peripheries Depends on PDZ-
mediated Interaction with Cx43
In addition to increasing the extent of Cx43 GJs, both C-
terminal tagging of Cx43 and the inhibitor peptide had
Figure 5. Peptide inhibition of Cx43–ZO-1 interaction increases the extent of membrane-localized Cx43 GJ plaques in HeLa cells and
cardiomyocytes. Confocal optical sections of HeLa Cx43 cells (A–F) and neonatal cardiomyocytes (G–I) cultured for 72 or 24 h, respectively,
without peptide (A, D, and G), or with reverse (B, E, and H) or inhibitor peptide (C, F, and I). All cells were immunolabeled for Cx43 (A–I).
Peptides were detected with streptavidin-Cy5 (D–I). WGA-TRITC delineates HeLa cell membranes (A–C). Arrowheads (C and I) denote
accumulation of membrane-localized Cx43 plaques in inhibitor-treated cells. Bars, 20
m (A–F); 80
m (G–I).
A. W. Hunter et al.
Molecular Biology of the Cell5692
effects on the pattern of colocalization between ZO-1 and
Cx43 (Figure 8). Interestingly, ZO-1 was often seen associ-
ated at the peripheries of GJs in HeLa Cx43 cells (Figure 8A).
By contrast, ZO-1 localization in HeLa Cx43-GFP cells
seemed to be independent of GJs (Figure 8B). Although
ZO-1 staining was seen along the edges of some large Cx43-
GFP plaques (Figure 8B, insets), this localization seemed to
be coincidental, with no evidence of actual overlap. How-
ever, when native Cx43 was coexpressed in HeLa Cx43-GFP
cells, colocalization of ZO-1 at plaque peripheries was rees-
tablished (Figure 8C, insets). This suggests that the mecha-
nistic link between restoration of ZO-1 localization at GJ
peripheries and size reduction of aberrantly large Cx43-GFP
GJs is the formation—after introduction of native Cx43—of
heteromeric connexons that are competent to bind ZO-1.
Similar to the HeLa Cx43-GFP phenotype, isolated cardi-
omyocytes treated with the inhibitor peptide produced
larger GJs that seemed to have less colocalized ZO-1 than
GJs in control cells (Figure 8, D–F). Again, most of the ZO-1
that colocalized with Cx43 in cardiomyocytes seemed re-
Figure 6. Quantification of gap junction patterning in HeLa cells
and cardiomyocytes after peptide inhibition of Cx43–ZO-1 interac-
tion. (A) The area of individual Cx43 plaques increased significantly
(*
#
p 0.001) after 72 h (HeLa Cx43) and 48 h (cardiomyocytes)
treatments with inhibitor peptide relative to controls (means SE;
n 30). (B) Size distributions of Cx43 plaques in cells treated with
inhibitor or no peptide. Inhibitor distributions are significantly dif-
ferent (chi-square test yielded *
#
p 0.001; n 780 for all conditions)
from control distributions (reverse peptide distributions omitted for
clarity). (C) Total area of membrane-localized Cx43 plaques was
increased significantly (*p 0.013) by inhibitor peptide (means
SE; n 30). (D) Cells treated with inhibitor peptide displayed a
slight reduction in GJ number (means SE; n 30). (E) Western
blots of HeLa and cardiomyocyte lysates show that peptide treat-
ment does not alter Cx43 protein expression levels. GAPDH blots
serve as loading controls.
Figure 7. The formation of larger GJ plaques after disruption of
Cx43–ZO-1 interaction results from a redistribution of Cx43. (A)
Representative gel of Triton fractionation assay. (B) Corresponding
plot of band intensities generated by densitometric analysis. Note
the change in peak profiles in the inhibitor condition. (C) Shifts in
subcellular Cx43 distribution are expressed as ratios of fractionated
Cx43 isoforms. Exposure to inhibitor peptide for 72 h significantly
reduced (*p 0.0001;
#
p 0.015) the level of Triton-soluble relative
to Triton-insoluble Cx43 in HeLa cells. (D) The shift of Cx43 from
soluble to insoluble pools induced by inhibitor peptide was accom-
panied by significant increases in Cx43-P1 (*p 0.012;
#
p 0.006)
and -P2 (*p 0.003;
#
p 0.005) phosphoisoforms (relative to
NP/P0) within the insoluble fraction.
ZO-1 Affects Cx43 Gap Junction Dynamics
Vol. 16, December 2005 5693
stricted to the periphery of GJ plaques rather than their
interior. In fact, some plaques displayed near continuous
runs of ZO-1 that circumscribed the perimeter of the GJ;
however, only a fraction of ZO-1 at plaque perimeters
seemed to overlap directly with Cx43 (Figure 8, D–F, arrow-
heads).
Quantitative analysis of confocal images confirmed that
the inhibitor peptide reduced ZO-1 colocalization with Cx43
Figure 8. Disruption of Cx43–ZO-1 interaction reduces the preferential association of ZO-1 with the perimeter of Cx43 GJ plaques. (A–C)
Maximum projection z-series of HeLa cells expressing native Cx43 (A), Cx43-GFP (B), or both GFP-tagged and native Cx43 (C). After 24 h,
cells were fixed and immunolabeled for Cx43 and ZO-1. Note that ZO-1 localizes to the periphery of small, discontinuous native Cx43 GJ
plaques (A, arrowhead) but is excluded from large, sheetlike Cx43-GFP plaques (B). Coexpression of native Cx43 (viral MOI 5) in HeLa
Cx43-GFP cells reduces plaque size and restores ZO-1 colocalization at plaques perimeters (C, arrowhead). (D–F) Maximum projection
z-series of neonatal cardiomyocytes treated for 48 h with the inhibitor (F), reverse (E), or no peptide (D). Note the striking pattern of ZO-1
colocalization with Cx43 plaque edges (D–F, arrowheads), which is reduced in inhibitor-treated cells. Asterisks (*) in top panels (A–F) denote
image areas enlarged in lower panels (the second set of bottom panels in D–F are enlargements from different images). Bars, 10
m (A–C,
top); 5
m (A–C, bottom); 5
m(D–F, top); 2.5
m (D–F, bottom).
A. W. Hunter et al.
Molecular Biology of the Cell5694
GJs. The average area of individual Cx43 plaques colocal-
ized with ZO-1 was decreased significantly in isolated car-
diomyocytes treated for 48 h with the inhibitor (17.5 1.3%)
compared with reverse (23.1 1.6%) and no peptide (25.5
2.0%) controls (Figure 9A). Fitting a line to a plot of the area
of Cx43 colocalized with ZO-1 versus GJ size (Figure 9B)
revealed a positive correlation (r 0.7 for each condition
when data were weighted by their uncertainties) between GJ
size and the area of colocalized ZO-1 that was independent
of peptide treatment. However, the differential effects of the
peptide treatments were manifest as a significant decrease in
the slope (equal to ZO-1 colocalization as a percentage GJ
area) of the line fitted to the inhibitor data (21.08 0.008%)
relative to the slopes derived from fits to the reverse (28.23
0.009%) and no peptide (28.43 0.011%) control data (Fig-
ure 9B). Plaques 3.44
m
2
were excluded from the fitting
process due to a single nonlinear outlier in the no peptide
data at these GJ sizes. Nevertheless, the colocalization values
associated with GJs 3.44
m
2
in cells treated with either
the inhibitor or reverse peptide fell within the 95% confi-
dence interval for the linear fits shown in Figure 9B. Because
the data were weighted by their uncertainties for linear
regression analysis (Figure 9B), mean colocalization values
obtained from linear fits differ slightly from those calculated
by unweighted averaging (Figure 9A).
Linear extrapolation of the data for the different peptide
conditions did not produce significantly different intercepts;
interestingly, however, the intercepts for all three conditions
were significantly different from zero (Inh p 0.002, Rev p
0.005, and NP p 0.001) and were consistently negative
values, which correspond to positive ordinal intercepts (Fig-
ure 9B). This suggests that, rather than being coassembled
with Cx43, ZO-1 is recruited to GJs after they reach a critical
size. Taking the average of the ordinal intercepts (assuming
that the intercepts from the three conditions are equivalent)
yielded a critical GJ size of 9300 400 nm
2
100 channels,
assuming 90-Å center-to-center hexagonal packing.
Together, the similar disruptive effects of the inhibitor
peptide and C-terminal tagging of Cx43 show that localiza-
tion of ZO-1 to GJ plaques is likely dependent on a binding
interaction between the C terminus of Cx43 and the PDZ2
domain of ZO-1.
DISCUSSION
The shape, extent, and distribution of gap junctional plaques
are recognized as important variables in the regulation of
intercellular coupling (Evans and Martin, 2002) and cou-
pling-dependent processes such as action potential spread in
heart and brain (Severs et al., 2004). Here, we provide evi-
dence that reduction in ZO-1 interaction with Cx43 is asso-
ciated with increases in the size of GJ plaques. First, we
demonstrated that disruption of ZO-1 binding by genetic
tagging of the C terminus of Cx43 with GFP leads to the
formation of abnormally large GJs when the tagged con-
struct is expressed in connexin-deficient HeLa cells; the ab-
normal Cx43-GFP phenotype was rescued by coexpression
of significant levels of native Cx43 (80% of the total con-
nexin). Second, we showed that targeted disruption of
Cx43–ZO-1 interaction using a peptide inhibitor based on
the PDZ-binding domain of Cx43 also caused a significant
increase in the mean size of Cx43 GJs in two cell types,
primary neonatal cardiomyocytes and a HeLa cell line ex-
pressing native Cx43. Third, we found that the increase in GJ
size arose from a redistribution of Cx43 from nonjunctional
pools to GJ plaques, with a possible contribution from in-
creased plaque fusion events. And finally, both GFP tagging
and the inhibitor peptide caused a significant reduction in
the amount of ZO-1 associated at the peripheries of GJ
plaques. Together, these data suggest a function for ZO-1 in
regulating the growth of Cx43 GJs by controlling the rate at
which connexons are incorporated into GJ plaques at their
peripheries.
Previously, we demonstrated that ZO-1 shows low-to-
moderate levels of colocalization with Cx43 GJs in adult rat
heart (Barker et al., 2002). The present study confirms and
extends this finding, demonstrating quantitatively in iso-
lated cardiomyocytes that colocalization between ZO-1 and
Cx43 in individual GJ plaques is limited to 28% overlap
(Figure 9B). Likewise, Nielsen et al. (2003) found that the
overall extent of colocalization between ZO-1 and Cx46 or
Cx50 in the eye lens was limited. Duffy et al. (2004) have also
confirmed in cultured astrocytes that colocalization of ZO-1
with Cx43 at cell borders is confined to a partial overlap.
Figure 9. Quantitative image analysis of ZO-1 association with
Cx43 GJs after peptide inhibition. (A) The average level of ZO-1
colocalization within individual plaques Cx43 (means SE) was
significantly reduced (*p 0.001;
#
p 0.006) in inhibitor-treated
cardiomyocytes. (B) Cx43 area colocalized with ZO-1 (means SE)
as a function of GJ size in myocytes. GJ sizes are means derived
from binned data. Lines are linear regression fits weighted by
uncertainties in GJ area colocalized with ZO-1; black dotted line is
the fit to the reverse peptide data (omitted for clarity). Slopes
correspond to the fractional area of individual Cx43 plaques colo-
calized with ZO-1, which was significantly reduced (*
#
p 0.001) in
inhibitor-treated cells.
ZO-1 Affects Cx43 Gap Junction Dynamics
Vol. 16, December 2005 5695
Overall, the data suggest that in astrocytes, lens and heart a
significant proportion of the connexin pool localized at cell
borders does not associate with ZO-1.
What then is the functional significance of the limited
interaction between ZO-1 and connexins? The high resolu-
tion confocal images presented here provide a detailed per-
spective of the spatial relationship between ZO-1 and Cx43
within GJ plaques. These images reveal that the majority of
ZO-1 does not overlap directly with plaque interiors; rather,
ZO-1 seems to associate preferentially with the edges of GJs.
EM data from Li et al. (2004) also support the idea that ZO-1
associates preferentially with the perimeter of GJ plaques.
That the perimeter is a specialized microdomain of GJ
plaques was established by studies that demonstrated that
addition of C-terminal-tagged Cx43 occurred at the periph-
ery of GJ plaques (Gaietta et al., 2002; Lauf et al., 2002).
Although there seems to be a general spatial correlation
between ZO-1 and sites of connexin recruitment into GJs, it
is unlikely that ZO-1 enhances GJ assembly because C-
terminal-truncated forms of Cx43 lacking the ZO-1 binding
domain are able to form functional GJs (Fishman et al., 1991;
Liu et al., 1993). More specifically, mutations targeting the
PDZ-binding domain of Cx43 and other connexins known to
interact with ZO-1 suggest that ZO-1 is dispensable for the
formation or ongoing aggregation of connexons into func-
tional GJs (Toyofuku et al., 2001; Nielsen et al., 2003; Li et al.,
2004). Likewise, fusion of epitope tags or GFP to the C
terminus of Cx43, which results in loss of ZO-1 binding
(Giepmans and Moolenaar, 1998; Giepmans et al., 2001b),
does not impede the formation of GJ plaques (Jordan et al.,
1999; Bukauskas et al., 2000; Gaietta et al., 2002; Lauf et al.,
2002; Hunter et al., 2003). Furthermore, as we show here,
reduction of Cx43–ZO-1 interaction increases rather than
decreases the size of Cx43 plaques. Collectively, these data
indicate that a role for ZO-1 in promoting the formation and
growth of GJ plaques, although not ruled out, is unlikely.
Based on increased Cx43–ZO-1 colocalization after enzy-
matically or chemically induced GJ internalization, we and
others have postulated that reduction of ZO-1 binding to
Cx43 inhibits GJ internalization via formation of so-called
annular GJs. (Barker et al., 2002; Segretain et al., 2004). How-
ever, our analysis of Cx43-GFP turnover in HeLa cells ar-
gues strongly against an explanation for the redistribution of
Cx43 and plaque expansion resulting from peptide inhibi-
tion of Cx43–ZO-1 interaction based solely on suppression
of GJ internalization. Cx43-GFP, a connexin incompetent to
interact with ZO-1, was found to turnover at a rate nearly
identical to that of native Cx43 (Figure 3). Thus, GJs com-
prised exclusively of Cx43-GFP, despite being abnormally
large, continue to turnover with a half-life (t
1/2
2h)
consistent with turnover rates in vivo (Fallon and Good-
enough, 1981; Beardslee et al., 1998) and with other systems
expressing tagged Cx43 (Gaietta et al., 2002; Segretain and
Falk, 2004). Moreover, cells expressing Cx43-GFP generate
annular GJs (Laird et al., 2001; Segretain and Falk, 2004), a
finding we have confirmed using live imaging of HeLa
Cx43-GFP cells (our unpublished data). And although Duffy
et al. (2004) confirmed coassociation between Cx43 and ZO-1
during enzymatically induced Cx43 internalization in astro-
cytes, pH-induced internalization of Cx43 was preceded by
a loss of Cx43–ZO-1 interaction. Therefore, increased Cx43–
ZO-1 colocalization after certain membrane internalization
events remains an unresolved issue, but it seems to be
independent of normal Cx43 turnover. Finally, if ZO-1 were
to play a prominent role in GJ internalization we might
expect to see ZO-1 localize preferentially to the center of
GJs—the major site of plaque removal—rather than the GJ
periphery—the major site of plaque assembly. Together,
these findings provide strong evidence against the direct
involvement of ZO-1 in the machinery of GJ internalization
and/or degradation.
Disruption of Cx43–ZO-1 interaction does seem to be as-
sociated with unconstrained growth of Cx43 channel aggre-
gates. We propose that ZO-1 actively constrains GJ size by
regulating the rate of channel accretion. Qualitatively, the
preferential localization at the perimeter of GJ plaques
places ZO-1 in a prime location—the site of plaque assem-
bly—to play a role in control of connexon incorporation into
GJs. But the most compelling evidence for ZO-1 control of
connexon accretion comes from observations that the inhib-
itor peptide caused a quantitative reduction in the amount
of ZO-1 colocalized within individual plaques, concomitant
with quantitative increases in Triton-insoluble Cx43 (at the
expense of the Triton-soluble pool) and plaque size. We
think these events are linked causally, which makes sense.
Even a slight reduction in ZO-1 binding at plaque perime-
ters might allow connexon accretion to proceed at a faster
rate. If the turnover rate remained constant, then an increase
in the rate of channel aggregation would be expected over
time to result in larger GJ plaques. Likewise, if the rate of
connexon accretion increased but connexin expression re-
mained constant, then we would expect to see an increase in
junctional (i.e., Triton-insoluble) Cx43 relative to nonjunc-
tional (i.e., Triton-soluble) Cx43. Our combined data sets fit
both the conditions and the expectations of the above-men-
tioned suppositions. The appearance of increased levels of
Triton-insoluble Cx43-P1 and -P2 isoforms, both of which
are constituents of mature GJs (Musil and Goodenough,
1991), in inhibitor-treated cells lends further support to our
proposal that ZO-1 plays a role in connexon accretion.
Importantly, although we think ZO-1 controls the rate of
connexon accretion and therefore exerts a level of control
over plaque expansion, ZO-1 does not sense GJ size. We
infer this from the linear relationship between GJ size and
the level of ZO-1 associated with individual plaques: ZO-1
continues to associate with larger GJs at levels equivalent (in
terms of percentage of plaque coverage) to that of smaller
GJs (Figure 8H). This suggests that subsequent to plaque
nucleation the mechanism of ZO-1 recruitment to GJs oper-
ates irrespective of plaque size. Thus, ZO-1 does not seem to
cap GJ size in absolute terms. Instead, the average GJ size at
steady state is set by a combination of the rates of connexon
accretion and GJ turnover.
Our conclusion that ZO-1 constrains the growth of GJs is
seemingly at odds with the findings of Laing et al. (2005),
who reported reduced appositional Cx43 immunostaining
and decreased GJ-mediated dye transfer after overexpres-
sion of a dominant-negative ZO-1 mutant in osteoblastic
cells. It is noteworthy that the truncated mutant used by
Laing et al. (2005) included N-terminal regions of ZO-1 that
extend beyond the PDZ2 domain, including the entire PDZ1
domain. The N-terminal half of ZO-1 is known to be a
functional component in cadherin-based cell adhesion (Itoh
et al., 1997). In fact, a truncated ZO-1 mutant similar to that
of Laing et al. (2005) was shown to disrupt cadherin-based
adherens junctions (AJs; Ryeom et al., 2000). Because AJs are
prerequisite for the formation and maintenance of Cx43 GJs
(Meyer et al., 1992; Hertig et al., 1996; Wei et al., 2005), the
effects of the Laing et al. (2005) mutant could be interpreted
as secondary due to decreased GJ stability, rather than
caused directly by loss of ZO-1 binding to Cx43.
Left unanswered is the question of how ZO-1 arrives at
plaque perimeters. Much of the ZO-1 that surrounds GJs
does not colocalize with Cx43. Presumably, ZO-1 not asso-
A. W. Hunter et al.
Molecular Biology of the Cell5696
ciated directly with Cx43 at plaque edges is complexed with
another cell junction type, the obvious candidate being the
AJ. The recent discovery that Cx43 coassembles with N-
cadherin (Wei et al., 2005) raises the possibility that AJs
nucleate GJs, with ZO-1 recruited first to AJs then passed
along to Cx43 once the GJ achieves a critical density of
channels, as suggested by our colocalization analysis (Figure
9B). Thus, it is unsurprising that overexpression of ZO-1 can
lead to enhanced GJ-mediated dye transfer (Laing et al.,
2005) because ZO-1 might promote the formation of AJs,
which in turn serve as GJ nucleation sites. The important
point is that ZO-1 does not inhibit the formation of GJs, we
think it merely constrains the rate of GJ growth after they
form.
The recent finding that ZO-2 also binds to the C terminus
of Cx43 (Singh et al., 2005) presents the intriguing prospect
that ZO-1 and ZO-2 compete for influence over Cx43 GJ
patterning. Because both ZO proteins interact with the PDZ-
binding motif in Cx43, the effects of our genetic tagging and
peptide inhibition on GJ size control might be influenced by
alterations in the activity of ZO-2 as well as ZO-1. At
present, we cannot exclude this possibility, but given that
Cx43 interacts predominately with ZO-1 in quiescent cells
(Singh et al., 2005), the effects of peptide inhibition in post-
mitotic cardiomyocytes reported here are likely due to the
specific disruption of ZO-1 activity. Nevertheless, distin-
guishing the activities of ZO-1 and ZO-2 with respect to
Cx43 GJ dynamics poses a formidable challenge, especially
considering the potential for functional redundancy and
compensation (Umeda et al., 2004).
Finally, control of Cx43 channel accretion has important
consequences for the regulation of GJ-mediated cell–cell
communication versus cell communication with the extra-
cellular environment via hemichannels (Goodenough and
Paul, 2003). Channel clustering is thought to lead to activa-
tion of individual intercellular channels within GJ plaques
(Bukauskas et al., 2000). A natural assumption is that by
controlling the rate of channel aggregation, ZO-1 directly
influences the level of cell–cell communication. But by lim-
iting connexon incorporation into GJ plaques, ZO-1 might
also effectively maintain a pool of nonjunctional Cx43
hemichannels that remain free to communicate with the
extracellular space. Changes in the rate of connexon accre-
tion by modulating ZO-1 action at plaque perimeters might
then be expected to provide a control point for dynamic
switching between junctional and nonjunctional communi-
cation mediated by Cx43 channels. In conjunction with fast
turnover rates, this type of control offers the possibility of
shifting the balance from one type of communication to the
other on relatively rapid time scales, thus allowing cells to
respond quickly to changes in their environment.
ACKNOWLEDGMENTS
We thank Dr. Ben Giepmans for ZO-1 GST-PDZ fusion proteins, assistance
with turnover assays, and helpful discussion; Jane Jourdan and Dr. Yuhua
Zhang for technical assistance; Dr. Stewart Denslow for advice on statistics;
Dr. Klaus Willecke for HeLa Cx43 cells; and Dr. Dale Laird for the Cx43-GFP
construct. This work was supported by National Institutes of Health Grants
HL07260 (to A.W.H.) and HL56728, HL36059, and HD39946 (to R.G.G.).
REFERENCES
Anderson, J. M. (1996). Cell signalling: MAGUK magic. Curr. Biol. 6, 382–384.
Barker, R. J., Price, R. L., and Gourdie, R. G. (2002). Increased association of
ZO-1 with connexin43 during remodeling of cardiac gap junctions. Circ. Res.
90, 317–324.
Beardslee, M. A., Laing, J. G., Beyer, E. C., and Saffitz, J. E. (1998). Rapid
turnover of connexin43 in the adult rat heart. Circ. Res. 83, 629635.
Bukauskas, F. F., Jordan, K., Bukauskiene, A., Bennett, M. V., Lampe, P. D.,
Laird, D. W., and Verselis, V. K. (2000). Clustering of connexin 43-enhanced
green fluorescent protein gap junction channels and functional coupling in
living cells. Proc. Natl. Acad. Sci. USA 97, 2556 –2561.
Burt, J. M., Fletcher, A. M., Steele, T. D., Wu, Y., Cottrell, G. T., and Kurjiaka,
D. T. (2001). Alteration of Cx 43, Cx40 expression ratio in A7r5 cells. Am. J.
Physiol. 280, C500–C508.
Defamie, N., Mograbi, B., Roger, C., Cronier, L., Malassine, A., Brucker-Davis,
F., Fenichel, P., Segretain, D., and Pointis, G. (2001). Disruption of gap
junctional intercellular communication by lindane is associated with aberrant
localization of connexin43 and zonula occludens-1 in 42GPA9 Sertoli cells.
Carcinogenesis 22, 1537–1542.
Duffy, H. S., Ashton, A. W., O’Donnell, P., Coombs, W., Taffet, S. M., Delmar,
M., and Spray, D. C. (2004). Regulation of connexin43 protein complexes by
intracellular acidification. Circ. Res. 94, 215–222.
Dupont, E., Matsushita, T., Kaba, R. A., Vozzi, C., Coppen, S. R., Khan, N.,
Kaprielian, R., Yacoub, M. H., and Severs, N. J. (2001). Altered connexin
expression in human congestive heart failure. J. Mol. Cell Cardiol. 33, 359
371.
Evans, W. H., and Martin, P. E. (2002). Gap junctions: structure and function
(Review). Mol. Membr. Biol. 19, 121–136.
Falk, M. M., Buehler, L. K., Kumar, N. M., and Gilula, N. B. (1997). Cell-free
synthesis and assembly of connexins into functional gap junction membrane
channels. EMBO J. 16, 2703–2716.
Fallon, R. F., and Goodenough, D. A. (1981). Five-hour half-life of mouse liver
gap-junction protein. J. Cell Biol. 90, 521–526.
Fishman, G. I., Hertzberg, E. L., Spray, D. C., and Leinwand, L. A. (1991).
Expression of connexin43 in the developing rat heart. Circ. Res. 68, 782–787.
Fonseca, C. G., Green, C. R., and Nicholson, L. F. (2002). Upregulation in
astrocytic connexin 43 gap junction levels may exacerbate generalized sei-
zures in mesial temporal lobe epilepsy. Brain Res. 929, 105–116.
Gaietta, G., Deerinck, T. J., Adams, S. R., Bouwer, J., Tour, O., Laird, D. W.,
Sosinsky, G. E., Tsien, R. Y., and Ellisman, M. H. (2002). Multicolor and
electron microscopic imaging of connexin trafficking. Science 296, 503–507.
Giepmans, B. N. (2004). Gap junctions and connexin-interacting proteins.
Cardiovasc. Res. 62, 233–245.
Giepmans, B. N., Hengeveld, T., Postma, F. R., and Moolenaar, W. H. (2001a).
Interaction of c-Src with gap junction protein connexin-43. Role in the regu-
lation of cell-cell communication. J. Biol. Chem. 276, 8544 8549.
Giepmans, B. N., and Moolenaar, W. H. (1998). The gap junction protein
connexin43 interacts with the second PDZ domain of the zona occludens-1
protein. Curr. Biol. 8, 931–934.
Giepmans, B. N., Verlaan, I., and Moolenaar, W. H. (2001b). Connexin-43
interactions with ZO-1 and
- and
-tubulin. Cell Commun. Adhes. 8, 219
223.
Goodenough, D. A., Goliger, J. A., and Paul, D. L. (1996). Connexins, connex-
ons, and intercellular communication. Annu. Rev. Biochem. 65, 475–502.
Goodenough, D. A., and Paul, D. L. (2003). Beyond the gap: functions of
unpaired connexon channels. Nat. Rev. Mol. Cell. Biol. 4, 285–294.
Green, C. R., Peters, N. S., Gourdie, R. G., Rothery, S., and Severs, N. J. (1993).
Validation of immunohistochemical quantification in confocal scanning laser
microscopy: a comparative assessment of gap junction size with confocal and
ultrastructural techniques. J. Histochem. Cytochem. 41, 1339 –1349.
Hall, J. E., and Gourdie, R. G. (1995). Spatial organization of cardiac gap
junctions can affect access resistance. Microsc. Res. Tech. 31, 446451.
Hare, J. F., and Taylor, K. (1991). Mechanisms of plasma membrane protein
degradation: recycling proteins are degraded more rapidly than those con-
fined to the cell surface. Proc. Natl. Acad. Sci. USA 88, 5902–5906.
Hertig, C. M., Butz, S., Koch, S., Eppenberger-Eberhardt, M., Kemler, R., and
Eppenberger, H. M. (1996). 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. 109, 11–20.
Hunter, A. W., Jourdan, J., and Gourdie, R. G. (2003). Fusion of GFP to the
carboxyl terminus of connexin43 increases gap junction size in HeLa cells.
Cell Commun. Adhes. 10, 211–214.
Itoh, M., Nagafuchi, A., Moroi, S., and Tsukita, S. (1997). Involvement of ZO-1
in cadherin-based cell adhesion through its direct binding to
catenin and
actin filaments. J. Cell Biol. 138, 181–192.
ZO-1 Affects Cx43 Gap Junction Dynamics
Vol. 16, December 2005 5697
Jin, C., Lau, A. F., and Martyn, K. D. (2000). Identification of connexin-
interacting proteins: application of the yeast two-hybrid screen. Methods 20,
219–231.
Johnson, R. G., Meyer, R. A., Li, X. R., Preus, D. M., Tan, L., Grunenwald, H.,
Paulson, A. F., Laird, D. W., and Sheridan, J. D. (2002). Gap junctions assemble
in the presence of cytoskeletal inhibitors, but enhanced assembly requires
microtubules. Exp. Cell Res. 275, 67–80.
Jordan, K., Solan, J. L., Dominguez, M., Sia, M., Hand, A., Lampe, P. D., and
Laird, D. W. (1999). Trafficking, assembly, and function of a connexin43-green
fluorescent protein chimera in live mammalian cells. Mol. Biol. Cell 10,
2033–2050.
Kausalya, P. J., Reichert, M., and Hunziker, W. (2001). Connexin45 directly
binds to ZO-1 and localizes to the tight junction region in epithelial MDCK
cells. FEBS Lett. 505, 92–96.
Laing, J. G., Chou, B. C., and Steinberg, T. H. (2005). ZO-1 alters the plasma
membrane localization and function of Cx43 in osteoblastic cells. J. Cell Sci.
118, 2167–2176.
Laing, J. G., Manley-Markowski, R. N., Koval, M., Civitelli, R., and Steinberg,
T. H. (2001). Connexin45 interacts with zonula occludens-1 and connexin43 in
osteoblastic cells. J. Biol. Chem. 276, 23051–23055.
Laird, D. W., Castillo, M., and Kasprzak, L. (1995). Gap junction turnover,
intracellular trafficking, and phosphorylation of connexin43 in brefeldin A-
treated rat mammary tumor cells. J. Cell Biol. 131, 1193–1203.
Laird, D. W., Jordan, K., and Shao, Q. (2001). Expression and imaging of
connexin-GFP chimeras in live mammalian cells. Methods Mol. Biol. 154,
135–142.
Laird, D. W., Puranam, K. L., and Revel, J. P. (1991). Turnover and phosphor-
ylation dynamics of connexin43 gap junction protein in cultured cardiac
myocytes. Biochem. J. 273, 67–72.
Lauf, U., Giepmans, B. N., Lopez, P., Braconnot, S., Chen, S. C., and Falk,
M. M. (2002). Dynamic trafficking and delivery of connexons to the plasma
membrane and accretion to gap junctions in living cells. Proc. Natl. Acad. Sci.
USA 99, 10446–10451.
Lauf, U., Lopez, P., and Falk, M. M. (2001). Expression of fluorescently tagged
connexins: a novel approach to rescue function of oligomeric DsRed-tagged
proteins. FEBS Lett. 498, 11–15.
Li, X., Olson, C., Lu, S., Kamasawa, N., Yasumura, T., Rash, J. E., and Nagy,
J. I. (2004). Neuronal connexin36 association with zonula occludens-1 protein
(ZO-1) in mouse brain and interaction with the first PDZ domain of ZO-1. Eur.
J. Neurosci. 19, 2132–2146.
Lindgren, M., Hallbrink, M., Prochiantz, A., and Langel, U. (2000). Cell-
penetrating peptides. Trends Pharmacol. Sci. 21, 99 –103.
Liu, S., Taffet, S., Stoner, L., Delmar, M., Vallano, M. L., and Jalife, J. (1993). A
structural basis for the unequal sensitivity of the major cardiac and liver gap
junctions to intracellular acidification: the carboxyl tail length. Biophys. J. 64,
1422–1433.
Lo, C. W. (2000). Role of gap junctions in cardiac conduction and develop-
ment: insights from the connexin knockout mice 87, 346 –348.
Maza, J., Das Sarma, J., and Koval, M. (2005). Defining a minimal motif
required to prevent connexin oligomerization in the endoplasmic reticulum.
J. Biol. Chem. 280, 21115–21121.
Meyer, R. A., Laird, D. W., Revel, J. P., and Johnson, R. G. (1992). Inhibition
of gap junction and adherens junction assembly by connexin and A-CAM
antibodies. J. Cell Biol. 119, 179–189.
Musil, L. S., Cunningham, B. A., Edelman, G. M., and Goodenough, D. A.
(1990). Differential phosphorylation of the gap junction protein connexin43 in
junctional communication-competent and -deficient cell lines. J. Cell Biol. 111,
2077–2088.
Musil, L. S., and Goodenough, D. A. (1991). Biochemical analysis of con-
nexin43 intracellular transport, phosphorylation, and assembly into gap junc-
tional plaques. J. Cell Biol. 115, 1357–1374.
Musil, L. S., and Goodenough, D. A. (1993). Multisubunit assembly of an
integral plasma membrane channel protein, gap junction connexin43, occurs
after exit from the ER. Cell 74, 1065–1077.
Musil, L. S., Le, A. C., VanSlyke, J. K., and Roberts, L. M. (2000). Regulation
of connexin degradation as a mechanism to increase gap junction assembly
and function. J. Biol. Chem. 275, 25207–25215.
Nielsen, P. A., Baruch, A., Shestopalov, V. I., Giepmans, B. N., Dunia, I.,
Benedetti, E. L., and Kumar, N. M. (2003). Lens connexins
3Cx46 and
8Cx50
interact with zonula occludens protein-1 (ZO-1). Mol. Biol. Cell 14, 2470–2481.
Nielsen, P. A., Beahm, D. L., Giepmans, B. N., Baruch, A., Hall, J. E., and
Kumar, N. M. (2002). Molecular cloning, functional expression, and tissue
distribution of a novel human gap junction-forming protein, connexin-31.9.
Interaction with zona occludens protein-1. J. Biol. Chem. 277, 38272–38283.
Ryeom, S. W., Paul, D., and Goodenough, D. A. (2000). Truncation mutants of
the tight junction protein ZO-1 disrupt corneal epithelial cell morphology.
Mol. Biol. Cell 11, 1687–1696.
Saez, J. C., Berthoud, V. M., Branes, M. C., Martinez, A. D., and Beyer, E. C.
(2003). Plasma membrane channels formed by connexins: their regulation and
functions. Physiol. Rev. 83, 1359–1400.
Sarma, J. D., Wang, F., and Koval, M. (2002). Targeted gap junction protein
constructs reveal connexin-specific differences in oligomerization. J. Biol.
Chem. 277, 20911–20918.
Segretain, D., and Falk, M. M. (2004). Regulation of connexin biosynthesis,
assembly, gap junction formation, and removal. Biochim. Biophys. Acta 1662,
3–21.
Segretain, D., Fiorini, C., Decrouy, X., Defamie, N., Prat, J. R., and Pointis, G.
(2004). A proposed role for ZO-1 in targeting connexin 43 gap junctions to the
endocytic pathway. Biochimie 86, 241–244.
Sepp, R., Severs, N. J., and Gourdie, R. G. (1996). Altered patterns of cardiac
intercellular junction distribution in hypertrophic cardiomyopathy. Heart 76,
412–417.
Severs, N. J., Dupont, E., Coppen, S. R., Halliday, D., Inett, E., Baylis, D., and
Rothery, S. (2004). Remodelling of gap junctions and connexin expression in
heart disease. Biochim. Biophys. Acta 1662, 138 –148.
Singh, D., Solan, J. L., Taffet, S. M., Javier, R., and Lampe, P. D. (2005).
Connexin 43 interacts with zona occludens-1 and -2 proteins in a cell cycle
stage-specific manner. J. Biol. Chem. 280, 30416 –30421.
Smith, J. H., Green, C. R., Peters, N. S., Rothery, S., and Severs, N. J. (1991).
Altered patterns of gap junction distribution in ischemic heart disease. An
immunohistochemical study of human myocardium using laser scanning
confocal microscopy. Am. J. Pathol. 139, 801– 821.
Solan, J. L., Fry, M. D., TenBroek, E. M., and Lampe, P. D. (2003). Connexin43
phosphorylation at S368 is acute during S and G2/M and in response to
protein kinase C activation. J. Cell Sci. 116, 2203–2211.
Sorgen, P. L., Duffy, H. S., Sahoo, P., Coombs, W., Delmar, M., and Spray,
D. C. (2004). 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. 279, 54695–54701.
Spach, M. S., Heidlage, J. F., Dolber, P. C., and Barr, R. C. (2000). Electrophys-
iological effects of remodeling cardiac gap junctions and cell size: experimen-
tal and model studies of normal cardiac growth. 86, 302–311.
Stevenson, B. R., Siliciano, J. D., Mooseker, M. S., and Goodenough, D. A.
(1986). Identification of ZO-1, a high molecular weight polypeptide associated
with the tight junction (zonula occludens) in a variety of epithelia. J. Cell Biol.
103, 755–766.
Toyofuku, T., Akamatsu, Y., Zhang, H., Kuzuya, T., Tada, M., and Hori, M.
(2001). c-Src regulates the interaction between connexin-43 and ZO-1 in car-
diac myocytes. J. Biol. Chem. 276, 1780 –1788.
Toyofuku, T., Yabuki, M., Otsu, K., Kuzuya, T., Hori, M., and Tada, M. (1998).
Direct association of the gap junction protein connexin-43 with ZO-1 in
cardiac myocytes. J. Biol. Chem. 273, 12725–12731.
Umeda, K., Matsui, T., Nakayama, M., Furuse, K., Sasaki, H., Furuse, M., and
Tsukita, S. (2004). Establishment and characterization of cultured epithelial
cells lacking expression of ZO-1. J. Biol. Chem. 279, 44785–44794.
VanSlyke, J. K., and Musil, L. S. (2000). Analysis of connexin intracellular
transport and assembly. Methods 20, 156–164.
Wei, C. J., Francis, R., Xu, X., and Lo, C. W. (2005). Connexin43 associated with
an N-cadherin-containing multiprotein complex is required for gap junction
formation in NIH3T3 cells. J. Biol. Chem. 280, 19925–19936.
Willecke, K., Eiberger, J., Degen, J., Eckardt, D., Romualdi, A., Guldenagel, M.,
Deutsch, U., and Sohl, G. (2002). Structural and functional diversity of con-
nexin genes in the mouse and human genome. Biol. Chem. 383, 725–737.
Zhu, C., Barker, R. J., Hunter, A. W., Zhang, Y., Jourdan, J., and Gourdie, R.
G. (2005). Quantitative analysis of ZO-1 colocalization with Cx43 gap junction
plaques in cultures of rat neonatal cardiomyocytes. Microsc. Microanal. 11,
244–248.
A. W. Hunter et al.
Molecular Biology of the Cell5698
... To analyze endogenous Cx43 activity at gap junctions without modifying expression levels, Rhett et al. [54] used a distinctive 25-amino-acid peptide (αCT1) designed to mimic the regulatory cytoplasmic domain of Cx43. This peptide redirects uncoupled Cx43 hemichannels into gap junctions, thereby reducing gap junction turnover and altering gap junction aggregation without impacting Cx43 protein levels [54,55]. Consequently, αCT1 provides the desired characteristics, as we were able to separate the expression levels of Cx43 from its function of Cx43. ...
... Previous studies have established that the mechanism of action for αCT1 is disrupting the Cx43 interaction with the zonula occludens junction protein 1 (ZO-1) [54,55] because ZO-1 downregulates Cx43 and prevents gap junction formation. ...
... The therapeutic α-connexin carboxyl-terminal peptide (αCT1) and its reverse sequence peptide (R-PEP) were synthesized by the American Peptide Company (Sunnyvale, CA, USA). The αCT1 peptide corresponds to a short Cx43 C-terminal sequence linked to an internalization sequence called antennapedia (RQPKIWFPNRRKPWKKRPRPDDLEI), as described and characterized by Hunter et al. (2005). The antennapedia-internalizing peptide sequence was RQPKIWFPNRRKPWKK. ...
Article
Full-text available
Simple Summary Cell-to-cell communication is important for tissue homeostasis and is, in general, decreased in cancer cells. The αCT1 peptide mimics the regulatory cytoplasmic domain of connexin 43 (Cx43), a protein that forms the gap junctions, and redirects uncoupled Cx43 hemichannels into gap junctions, reducing gap junction turnover and changing its aggregation without affecting Cx43 protein levels. In this study, we investigated the effects of the αCT1 peptides on the viability of canine benign and malignant mammary epithelial cells. The αCT1 peptide exerted differential effects on the viability of canine adenoma and adenocarcinoma cells while preserving the normal canine mammary epithelial cells. αCT1 may be possibly included in therapeutic protocols for canine mammary neoplasms. Abstract Mammary cancer is highly prevalent in non-castrated female dogs. Cell-to-cell communication is an important mechanism to maintain homeostasis, and connexins are proteins that assemble to form the communicating gap junctions. In many cancers, communication capacity is reduced; several approaches are being tested in order to increase the communication capacity in cancer cells and, therefore, alter their viability. This study analyzed the effects of the alpha-connexin carboxyl-terminal peptide (αCT1) on canine mammary non-neoplastic and neoplastic epithelial cells. Seven canine epithelial mammary cell lines were used. Among these, one was a normal canine epithelial mammary cell line (LOEC-NMG), two canine mammary adenomas (LOEC-MAd1 and LOEC-MAd2), and four canine mammary adenocarcinomas (LOEC-MCA1, LOEC-MCA2, LOEC-MCA3 and CF41). The αCT1 corresponds to a short Cx43 C-terminal sequence linked to an internalization sequence called the antennapedia. After 24 h of incubation, the medium containing different αCT1 peptide concentrations was added to the cells, and only the culture medium was used for control. The 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) test was used to quantify cell viability before treatment and 48, 72, and 96 h after the treatment. Results showed that the normal mammary epithelial cell line (LOEC-NMG) was resistant to treatment with αCT1, which is consistent with a previous study on human mammary cell lines. One of the adenoma cell lines (LOEC-MAd2) was also resistant to treatment with αCT1, although the other (LOEC-MAd1) was susceptible to treatment, mostly at 72 h after treatment. Regarding the four canine adenocarcinoma cell lines, they differ regarding the susceptibility to the treatment with αCT1. Three cell lines, canine mixed adenocarcinoma (LOEC-MCA1), canine complex adenocarcinoma (LOEC-MCA2), and commercial canine mammary adenocarcinoma cell line CF41, were susceptible to treatment with αCT1, while one canine mammary adenocarcinoma cell line (LOEC-MCA3) was resistant to treatment. In most αCT1 treated cell lines, Cx43 was strongly detected in cell membranes by immunofluorescence. We propose that αCT1 restored the cell-to-cell communication capacity of neoplastic cells and induced inhibitory effects on cell viability.
... α-connexin carboxyl-terminal (aCT1) is a therapeutic peptide modeled on the sequence of the C-terminal tail of Cx43. Application of aCT1 was shown to increase gap junction size and GJIC [12,44]. Prior work shows the anti-cancer effects of aCT1 in the brain, melanoma, and breast cancer [12,45,46]. ...
... The aCT1 sequence is modeled after the C-terminal sequence of Cx43, which is reported to interact with the tight junction protein ZO-1 [57][58][59]. The reported mechanism of action for aCT1 leading to increased GJIC is binding to ZO-1, thereby physically impairing ZO-1 from preventing gap junction accretion [44]. Additionally, the c-terminus of Cx43 is known to interact with a large number of proteins including src, caveolin, β-catenin, p120-catenin, and tubulin [8,[60][61][62]. ...
... The aCT1 sequence is modeled after the C-terminal sequence of Cx43, which is reported to interact with the tight junction protein ZO-1 [57][58][59]. The reported mechanism of action for aCT1 leading to increased GJIC is binding to ZO-1, thereby physically impairing ZO-1 from preventing gap junction accretion [44]. Additionally, the c-terminus of Cx43 is known to interact with a large number of proteins including src, caveolin, βcatenin, p120-catenin, and tubulin [8,[60][61][62]. ...
Article
Full-text available
Simple Summary Breast cancer is a leading cause of death for women worldwide. This study aims to examine the mechanisms for the anti-cancer effects of a therapeutic peptide agent targeted to connexin 43 (Cx43) called alpha-connexin carboxyl-terminal peptide (aCT1). Findings from this study confirm that aCT1 promotes gap junction intercellular communication and Cx43 protein stabilization and that the peptide binds to Zonal Occludens-1 (ZO-1), consistent with prior reports on the mechanism of action of this agent. New findings from this study identify novel aCT1-interacting proteins and the effect of aCT1 on breast cancer signaling. Abstract Connexin 43 (Cx43) is a protein encoded by the GJA1 gene and is a component of cell membrane structures called gap junctions, which facilitate intercellular communication. Prior evidence indicates that elevated GJA1 expression in the HER2-positive (HER2+) subtype of breast cancer is associated with poor prognosis. Prior evidence also suggests that HER2+ breast cancers that have become refractory to HER2-targeted agents have a loss of Cx43 gap junction intercellular communication (GJIC). In this study, a Cx43-targeted agent called alpha-connexin carboxyl-terminal peptide (aCT1) is examined to determine whether GJIC can be rescued in refractory HER2+ breast cancer cells. A proposed mechanism of action for aCT1 is binding to the tight junction protein Zonal Occludens-1 (ZO-1). However, the true scope of activity for aCT1 has not been explored. In this study, mass spectrometry proteomic analysis is used to determine the breadth of aCT1-interacting proteins. The NanoString nCounter Breast Cancer 360 panel is also used to examine the effect of aCT1 on cancer signaling in HER2+ breast cancer cells. Findings from this study show a dynamic range of binding partners for aCT1, many of which regulate gene expression and RNA biology. nCounter analysis shows that a number of pathways are significantly impacted by aCT1, including upregulation of apoptotic factors, leading to the prediction and demonstration that aCT1 can boost the cell death effects of cisplatin and lapatinib in HER2+ breast cancer cells that have become resistant to HER2-targeted agents.
... 53 The mechanism of action of a-CT1 is that it inhibits the interaction between the tight junction associated protein Zonula Occludens 1 (ZO-1) and the C-terminus of Cx43. 54 Previous studies have shown that treatment with a-CT1 in either mouse cardiac injury models or scratch wounded HeLa cells increases S368 phosphorylation of Cx43. 16 Since our previous data demonstrated that Cx43-S368 phosphorylation is lower in PCKm inhibited HaCaT cells, we examined if a-CT1 can increase Cx43-S368 phosphorylation in an in vitro wound healing assay in HaCaT cells. ...
Article
Full-text available
Downregulation of intercellular communication through suppression of gap junctional conductance is necessary during wound healing. Connexin 43 (Cx43), a prominent gap junction protein in skin, is downregulated following wounding to restrict communication between keratinocytes. Previous studies found that PKCμ, a novel PKC isozyme, regulates efficient cutaneous wound healing. However, the molecular mechanism by which PKCμ regulates wound healing remains unknown. We have identified that PKCμ suppresses intercellular communication and enhances cell migration in an in vitro wound healing model by regulating Cx43 containing gap junctions. PKCμ can directly interact with and phosphorylate Cx43 at S368, which leads to Cx43 internalization and downregulation. Finally, utilizing phosphomimetic and non-phosphorylatable S368 substitutions and gap junction inhibitors, we confirmed that PKCμ regulates intercellular communication and in vitro wound healing by controlling Cx43-S368 phosphorylation. These results define PKCμ as a critical regulator of Cx43 phosphorylation to control cell migration and wound healing in keratinocytes.
... An important concern when targeting Cx43 in the endothelium is the potential impact on Cx43-mediated GJ coupling between cardiomyocytes. However, aCT11 inhibits Cx43 HCs by promoting their accretion into GJ plaques, avoiding any GJ uncoupling.[72][73][74][75] Although small-molecule drugs are easier to deliver in vivo, their reduced specificity often yields off-target effects. ...
Article
Full-text available
The type I transmembrane protein BT-IgSF is predominantly localized in the brain and testes. It belongs to the CAR subgroup of Ig cell adhesion proteins, that are hypothesized to regulate connexin expression or localization. Here, we studied the putative link between BT-IgSF and connexins in astrocytes, ependymal cells and neurons of the mouse. Global knockout of BT-IgSF caused an increase in the clustering of connexin43 (Gja1), but not of connexin30 (Gjb6), on astrocytes and ependymal cells. Additionally, knockout animals displayed reduced expression levels of connexin43 protein in the cortex and hippocampus. Importantly, analysis of biocytin spread in hippocampal or cortical slices from mature mice of either sex revealed a decrease in astrocytic cell-cell coupling in the absence of BT-IgSF. Blocking either protein biosynthesis or proteolysis showed that the lysosomal pathway increased connexin43 degradation in astrocytes. Localization of connexin43 in subcellular compartments was not impaired in astrocytes of BT-IgSF mutants. In contrast to connexin43 the localization and expression of connexin36 (Gjd2) on neurons was not affected by the absence of BT-IgSF. Overall, our data indicate that the IgCAM BT-IgSF is essential for correct gap junction-mediated astrocyte-to-astrocyte cell communication. Significance Statement Astrocytes regulate a variety of physiological processes in the developing and adult brain that are essential for proper brain function. Astrocytes form extensive networks in the brain and communicate via gap junctions. Disruptions of gap junction coupling are found in several diseases such as neurodegeneration or epilepsy. Here, we demonstrate that the cell adhesion protein BT-IgSF is essential for gap junction mediated coupling between astrocytes in the cortex and hippocampus.
Article
Full-text available
Fibrosis initially appears as a normal response to damage, where activated fibroblasts produce large amounts of the extracellular matrix (ECM) during the wound healing process to assist in the repair of injured tissue. However, the excessive accumulation of the ECM, unresolved by remodeling mechanisms, leads to organ dysfunction. Connexins, a family of transmembrane channel proteins, are widely recognized for their major roles in fibrosis, the epithelial–mesenchymal transition (EMT), and wound healing. Efforts have been made in recent years to identify novel mediators and targets for this regulation. Connexins form gap junctions and hemichannels, mediating communications between neighboring cells and inside and outside of cells, respectively. Recent evidence suggests that connexins, beyond forming channels, possess channel-independent functions in fibrosis, the EMT, and wound healing. One crucial channel-independent function is their role as the primary functional component for cell adhesion. Other channel-independent functions of connexins involve their roles in mitochondria and exosomes. This review summarizes the latest advances in the channel-dependent and independent roles of connexins in fibrosis, the EMT, and wound healing, with a particular focus on eye diseases, emphasizing their potential as novel, promising therapeutic targets.
Article
Triple-negative breast cancer (TNBC) represents the most lethal and treatment-resistant breast cancer subtype with limited treatment options. We previously identified a protein complex unique to TNBC composed of the gap junction protein connexin 26 (Cx26), the pluripotency transcription factor NANOG, and focal adhesion kinase (FAK). We sought to determine whether a peptide mimetic of the interaction region of Cx26 attenuated tumor growth in pre-clinical models. We designed peptides based on Cx26 juxtamembrane domains and performed binding experiments with NANOG and FAK using surface plasmon resonance. Binding studies revealed that the Cx26 C-terminal tail and intracellular loop bound to NANOG and FAK with submicromolar-to-micromolar affinity and that a 5-amino acid sequence in the C-terminal tail of Cx26 (RYCSG) was sufficient for binding. Peptides with high affinity were engineered with a cell-penetrating antennapedia sequence and assessed in functional assays including cell proliferation, tumorsphere formation, and in vivo tumor growth, and downstream signaling changes were measured. The cell-penetrating Cx26 peptide (aCx26-pep) disrupted self-renewal while reducing nuclear FAK and NANOG and inhibiting NANOG target gene expression in TNBC cells but not luminal mammary epithelial cells. In vivo, aCx26-pep reduced tumor growth and proliferation and induced cell death. Here, we provide proof-of-concept that a Cx26 peptide-based strategy inhibits growth and alters NANOG activity specifically in TNBC, indicating the therapeutic potential of this targeting approach.
Article
Cerebral cavernous malformation type-3 (CCM3) is a type of brain vascular malformation caused by mutations in programmed cell death protein-10 (PDCD10). It is characterized by early life occurrence of hemorrhagic stroke and profound blood-brain barrier defects. The pathogenic mechanisms responsible for microvascular hyperpermeability and lesion progression in CCM3 are still largely unknown. The current study examined brain endothelial barrier structural defects formed in the absence of CCM3 in vivo and in vitro that may lead to CCM3 lesion leakage. We found significant upregulation of a 20 kDa isoform of connexin 43 (GJA1-20 k) in brain endothelial cells (BEC) in both non-leaky and leaky lesions, as well as in an in vitro CCM3 knockdown model (CCM3KD-BEC). Morphological, biochemical, FRET, and FRAP analyses of CCM3KD-BEC found GJA1-20 k regulates full-length GJA1 biogenesis, prompting uncontrolled gap junction growth. Furthermore, by binding to a tight junction scaffolding protein, ZO-1, GJA1-20 k interferes with Cx43/ZO-1 interactions and gap junction/tight junction crosstalk, promoting ZO-1 dissociation from tight junction complexes and diminishing claudin-5/ZO-1 interaction. As a consequence, the tight junction complex is destabilized, allowing "replacement" of tight junctions with gap junctions leading to increased brain endothelial barrier permeability. Modifying cellular levels of GJA1-20 k rescued brain endothelial barrier integrity re-establishing the spatial organization of gap and tight junctional complexes. This study highlights generation of potential defects at the CCM3-affected brain endothelial barrier which may underlie prolonged vascular leakiness.
Article
Full-text available
We examined the roles of the extracellular domains of a gap junction protein and a cell adhesion molecule in gap junction and adherens junction formation by altering cell interactions with antibody Fab fragments. Using immunoblotting and immunocytochemistry we demonstrated that Novikoff cells contained the gap junction protein, connexin43 (Cx43), and the cell adhesion molecule, A-CAM (N-cadherin). Cells were dissociated in EDTA, allowed to recover, and reaggregated for 60 min in media containing Fab fragments prepared from a number of antibodies. We observed no cell-cell dye transfer 4 min after microinjection in 90% of the cell pairs treated with Fab fragments of antibodies for the first or second extracellular domain of Cx43, the second extracellular domain of connexin32 (Cx32) or A-CAM. Cell-cell dye transfer was detected within 30 s in cell pairs treated with control Fab fragments (pre-immune serum, antibodies to the rat major histocompatibility complex or the amino or carboxyl termii of Cx43). We observed no gap junctions by freeze-fracture EM and no adherens junctions by thin section EM between cells treated with the Fab fragments that blocked cell-cell dye transfer. Gap junctions were found on approximately 50% of the cells in control samples using freeze-fracture EM. We demonstrated with reaggregated Novikoff cells that: (a) functional interactions of the extracellular domains of the connexins were necessary for the formation of gap junction channels; (b) cell interactions mediated by A-CAM were required for gap junction assembly; and (c) Fab fragments of antibodies for A-CAM or connexin extracellular domains blocked adherens junction formation.
Article
Full-text available
Arrhythmias are a common and potentially life-threatening complication of myocardial ischemia and infarction in humans. The structural pathways for the rapid intercellular conduction of the electrical impulse that stimulates coordinated contraction in the myocardium are formed by the gap junctions situated at intercalated disks. By raising antibodies to cardiac gap-junctional protein, and using these antibodies in an immunohistochemical procedure in combination with the technique of laser scanning confocal microscopy, we have succeeded in localizing gap junctions, with a clarity not previously possible, through thick volumes of human myocardial tissue. To explore the structural basis for ischemia and infarction-related arrhythmogenesis, antibody labeling and laser scanning confocal microscopy were applied to study the organization, distribution, and other characteristics of gap junctions in the explanted hearts of patients undergoing cardiac transplantation for advanced ischemic heart disease. In areas of myocardium free from histologically detectable structural damage, there was no significant difference in the size of distribution of labeled gap junctions, or in their number per intercalated disk, between left ventricular tissue (in which functional impairment was severe) and right ventricular tissue (in which functional impairment was minimal). However, in myocytes at the border of healed infarcts--zones to which the slow conduction responsible for reentry arrhythmias has been localized--the organization of gap junctions was markedly disordered; instead of being aggregated into discrete intercalated disks, gap-junctional immunostaining was spread extensively over myocyte surfaces. Some infarct zones were bridged by continuous strands of myocytes, coupled to one another by gap junctions, thereby linking healthy myocardium on either side. At their thinnest, these bridges were in some instances no wider than a single attenuated myocyte. The conclusions are 1) a widespread, generalized derangement of gap junction organization does not appear to underlie functional impairment in the ischemic heart, 2) a disorderly arrangement typifies gap junctions in myocytes of the infarct border zone, and this may contribute to alterations in conduction that are capable of precipitating reentry arrhythmias, and 3) delicate chains of myocytes traverse some healed infarcts, apparently forming electrically coupled bridges across what would otherwise constitute blocked zones. The weakest link in this chain can be a single, degenerating myocyte; avoidance of arrhythmia may therefore depend on the continued survival of this single cell.
Article
Full-text available
We previously demonstrated that the gap junction protein connexin43 is translated as a 42-kD protein (connexin43-NP) that is efficiently phosphorylated to a 46,000-Mr species (connexin43-P2) in gap junctional communication-competent, but not in communication-deficient, cells. In this study, we used a combination of metabolic radiolabeling and immunoprecipitation to investigate the assembly of connexin43 into gap junctions and the relationship of this event to phosphorylation of connexin43. Examination of the detergent solubility of connexin43 in communication-competent NRK cells revealed that processing of connexin43 to the P2 form was accompanied by acquisition of resistance to solubilization in 1% Triton X-100. Immunohistochemical localization of connexin43 in Triton-extracted NRK cells demonstrated that connexin43-P2 (Triton-insoluble) was concentrated in gap junctional plaques, whereas connexin43-NP (Triton-soluble) was predominantly intracellular. Using either a 20 degrees C intracellular transport block or cell-surface protein biotinylation, we determined that connexin43 was transported to the plasma membrane in the Triton-soluble connexin43-NP form. Cell-surface biotinylated connexin43-NP was processed to Triton-insoluble connexin43-P2 at 37 degrees C. Connexin43-NP was also transported to the plasma membrane in communication defective, gap junction-deficient S180 and L929 cells but was not processed to Triton-insoluble connexin43-P2. Taken together, these results demonstrate that gap junction assembly is regulated after arrival of connexin43 at the plasma membrane and is temporally associated with acquisition of insolubility in Triton X-100 and phosphorylation to the connexin43-P2 form.
Article
Connexin43 is a member of the highly homologous connexin family of gap junction proteins. We have studied how connexin monomers are assembled into functional gap junction plaques by examining the biosynthesis of connexin43 in cell types that differ greatly in their ability to form functional gap junctions. Using a combination of metabolic radiolabeling and immunoprecipitation, we have shown that connexin43 is synthesized in gap junctional communication-competent cells as a 42-kD protein that is efficiently converted to a approximately 46-kD species (connexin43-P2) by the posttranslational addition of phosphate. Surprisingly, certain cell lines severely deficient in gap junctional communication and known cell-cell adhesion molecules (S180 and L929 cells) also expressed 42-kD connexin43. Connexin43 in these communication-deficient cell lines was not, however, phosphorylated to the P2 form. Conversion of S180 cells to a communication-competent phenotype by transfection with a cDNA encoding the cell-cell adhesion molecule L-CAM induced phosphorylation of connexin43 to the P2 form; conversely, blocking junctional communication in ordinarily communication-competent cells inhibited connexin43-P2 formation. Immunohistochemical localization studies indicated that only communication-competent cells accumulated connexin43 in visible gap junction plaques. Together, these results establish a strong correlation between the ability of cells to process connexin43 to the P2 form and to produce functional gap junctions. Connexin43 phosphorylation may therefore play a functional role in gap junction assembly and/or activity.
Article
Intercellular gap junction channels are thought to form when oligomers of connexins from one cell (connexons) register and pair with connexons from a neighboring cell en route to forming tightly packed arrays (plaques). In the current study we used the rat mammary BICR-M1Rk tumor cell line to examine the trafficking, maturation, and kinetics of connexin43 (Cx43). Cx43 was conclusively shown to reside in the Golgi apparatus in addition to sites of cell-cell apposition in these cells and in normal rat kidney cells. Brefeldin A (BFA) blocked Cx43 trafficking to the surface of the mammary cells and also prevented phosphorylation of the 42-kD form of Cx43 to 44- and 46-kD species. However, phosphorylation of Cx43 occurred in the presence of BFA while it was still a resident of the ER or Golgi apparatus yielding a 43-kD form of Cx43. Moreover, the 42- and 43-kD forms of Cx43 trapped in the ER/Golgi compartment were available for gap junction assembly upon the removal of BFA. Mammary cells treated with BFA for 6 h lost preexisting gap junction "plaques," as well as the 44- and 46-kD forms of Cx43 and functional coupling. These events were reversible 1 h after the removal of BFA and not dependent on protein synthesis. In summary, we provide strong evidence that in BICR-M1Rk tumor cells: (a) Cx43 is transiently phosphorylated in the ER/Golgi apparatus, (b) Cx43 trapped in the ER/Golgi compartment is not subject to rapid degradation and is available for the assembly of new gap junction channels upon the removal of BFA, (c) the rapid turnover of gap junction plaques is correlated with the loss of the 44- and 46-kD forms of Cx43.
Article
We examined the roles of the extracellular domains of a gap junction protein and a cell adhesion molecule in gap junction and adherens junction formation by altering cell interactions with antibody Fab fragments. Using immunoblotting and immunocytochemistry we demonstrated that Novikoff cells contained the gap junction protein, connexin43 (Cx43), and the cell adhesion molecule, A-CAM (N-cadherin). Cells were dissociated in EDTA, allowed to recover, and reaggregated for 60 min in media containing Fab fragments prepared from a number of antibodies. We observed no cell-cell dye transfer 4 min after microinjection in 90% of the cell pairs treated with Fab fragments of antibodies for the first or second extracellular domain of Cx43, the second extracellular domain of connexin32 (Cx32) or A-CAM. Cell-cell dye transfer was detected within 30 s in cell pairs treated with control Fab fragments (pre-immune serum, antibodies to the rat major histocompatibility complex or the amino or carboxyl termii of Cx43). We observed no gap junctions by freeze-fracture EM and no adherens junctions by thin section EM between cells treated with the Fab fragments that blocked cell-cell dye transfer. Gap junctions were found on approximately 50% of the cells in control samples using freeze-fracture EM. We demonstrated with reaggregated Novikoff cells that: (a) functional interactions of the extracellular domains of the connexins were necessary for the formation of gap junction channels; (b) cell interactions mediated by A-CAM were required for gap junction assembly; and (c) Fab fragments of antibodies for A-CAM or connexin extracellular domains blocked adherens junction formation.
Article
A novel, brilliantly red fluorescent protein, DsRed has become available recently opening up a wide variety of experimental opportunities for double labeling and fluorescence resonance electron transfer experiments in combination with green fluorescent protein (GFP). Unlike in the case of GFP, proteins tagged with DsRed were often found to aggregate within the cell. Here we report a simple method that allows rescuing the function of an oligomeric protein tagged with DsRed. We demonstrate the feasibility of this approach on the subunit proteins of an oligomeric membrane channel, gap junction connexins. Additionally, DsRed fluorescence was easily detected 12–16 h post transfection, much earlier than previously reported, and could readily be differentiated from co-expressed GFP. Thus, this approach can eliminate the major drawbacks of this highly attractive autofluorescent protein.
Article
Congestive heart failure is associated with a high risk of life-threatening ventricular re-entrant arrhythmias. Down-regulation of the principal gap-junctional protein of the ventricular myocytes, connexin43, has previously been implicated in arrhythmia in ischaemic heart disease, but it is not known whether connexin43 is similarly reduced in heart failure due to idiopathic dilated cardiomyopathy, whether disease-related connexin43 down-regulation occurs at the level of transcription or translation, or whether the expression of other connexin isotypes is altered in congestive heart failure. We therefore investigated the expression of the four connexins expressed in the heart—connexins 43, 40, 45 and 37—at the mRNA and protein levels in explanted hearts from transplant patients with end-stage heart failure (NYHA class 4) by immunoconfocal analysis, and northern and western blotting. Connexin43 mRNA and protein were markedly downregulated in the left ventricle in end-stage heart failure due both to ischaemic cardiomyopathy and idiopathic dilated cardiomyopathy. Connexin43 content was spatially heterogeneous in the diseased ventricle. Connexin40 mRNA was increased in the ischaemic group, more so in the left ventricle than the right. This correlated with an increased depth of connexin40 protein expression in myocytes at the endocardial surface. Connexin45 mRNA and protein, present only in very low quantities, followed a similar trend to connexin43, while connexin37 (exclusively expressed in endothelium) showed no change. Our findings show that congestive heart failure is associated with significantly reduced levels of the principal gap junction protein, connexin43, in the left ventricle, potentially contributing to enhanced arrhythmogenicity and contractile dysfunction. This down-regulation is due predominantly to a reduced transcript steady-state level. Elevated connexin40 may represent a compensatory response that improves the spread of depolarization in the otherwise compromised ischaemic ventricle.
Article
Connexin43 is the predominant gap junction protein expressed in the heart. To determine the relation between cardiac maturation and gap junction gene expression, the developmental profiles of connexin43 mRNA and protein were examined in the rat heart. Connexin43 mRNA levels accumulate progressively (eightfold) during embryonic and early neonatal stages, accompanied by a parallel, but temporally delayed, accumulation of connexin43 protein (15-fold). As the heart matures further, both mRNA and protein levels subsequently decline, to about 50% and 30% of their maximum levels, respectively. These observations suggest that increases in intercellular coupling that characterize cardiac development do not depend solely on modulation of connexin43 gene expression, but rather are likely to involve organization of gap junction channels into the intercalated disc.