ArticlePDF Available

Connexin43 phosphorylation at S368 is acute during S and G2/M and in response to protein kinase C activation

Authors:

Abstract and Figures

Phorbol esters such as 12-O-tetradeconylphorbol-13-acetate (TPA) activate protein kinase C, increase Connexin43 (Cx43) phosphorylation, and decrease cell-cell communication via gap junctions in many cell types. Previous work has implicated protein kinase C (PKC) in the direct phosphorylation of Cx43 at S368, which results in a change in single channel behavior that contributes to a decrease in intercellular communication. We have examined Cx43 phosphorylation in several cell lines with an antibody specific for phosphorylated S368. We show that this antibody detects Cx43 only when it is phosphorylated at S368 and, consistent with previous results, TPA treatment causes a dramatic increase in phosphorylation at S368. However, in some cell types, the increased phosphorylation at S368 did not cause a detectable shift in migration as compared with the nonphosphorylated Cx43. Immunofluorescence showed increased S368 immunolabeling in cytoplasmic and plasma membrane structures in response to TPA. Immunoblot analysis of synchronized cells showed increased phosphorylation at S368 during S and G2/M phases of the cell cycle. S-phase cells contained more total Cx43 but assembled fewer functional gap junctional channels than G0-phase cells. Since M-phase cells also communicate poorly and contain few assembled gap junctions, phosphorylation at S368 appears to be negatively correlated with gap junction assembly. Thus, both gap junctional communication and S368 phosphorylation change during S phase and G2/M, implying that phosphorylation at S368 might play a role in key cell-cycle events.
Content may be subject to copyright.
Introduction
Gap junctions are specialized membrane domains composed of
collections of channels that directly connect neighboring cells
(Willecke et al., 2002). These pathways provide for the cell-to-
cell diffusion of small molecules, including ions, amino acids,
nucleotides and second messengers (e.g. Ca2+, cAMP, cGMP,
IP3). Recent studies on the targeted disruption of connexin
genes, which encode vertebrate gap junction channel proteins,
provide strong support for roles in cell growth control and
embryonic development, as well as the transmission of
metabolites and electrical signals between cells (Willecke et
al., 2002).
Transient changes in gap junctional communication,
probably regulated by signaling cascades, have been observed
and appear necessary for normal cell cycling. For example, gap
junctional communication was reported to be moderate during
G1/S, increased through S and decreased in G2/M (Bittman and
LoTurco, 1999; Stein et al., 1992). The downregulation of
junctional communication during G2/M has been correlated
with increased p34cdc2 kinase-dependent phosphorylation of
Cx43 (Kanemitsu et al., 1998; Lampe et al., 1998a) and
redistribution of Cx43 from gap junctions to the cytoplasm
(Lampe et al., 1998a; Xie et al., 1997). Gap junctional
structures reassemble and communication is gradually restored
as cells proceed through G1(Stein et al., 1992; Xie et al., 1997).
Cx43 is phosphorylated at multiple serine residues in vivo
(Berthoud et al., 1992; Brissette et al., 1991; Crow et al., 1990;
Kadle et al., 1991; Laird et al., 1991; Musil et al., 1990), and
upon phosphorylation, Cx43 migration in polyacrylamide gel
electrophoresis (SDS-PAGE) is reduced. Although apparently
not required for the formation of functional channels (Dunham
et al., 1992; Fishman et al., 1991), phosphorylation of gap
junction proteins appears to regulate channel function (gating)
and the rates of channel assembly and turnover (Brissette et al.,
1991; Kwak et al., 1995a; Kwak et al., 1995b; Kwak et al.,
1995c; Lampe, 1994; Lampe et al., 2000).
In the sustained absence of connexin expression,
tumorigenesis is enhanced (Laird et al., 1999; Moennikes et
al., 1999). The correlation between neoplastic transformation
and reduced gap junctional communication (Atkinson et al.,
1981; Azarnia and Loewenstein, 1984; de Feijter et al., 1990)
has led to the hypothesis that reduced cell-cell communication
is a critical step in multistage carcinogenesis (Fitzgerald and
Yamasaki, 1990; Trosko et al., 1990). PKC has received
considerable attention because PKC activators (e.g. TPA),
which promote tumorigenesis, both increase Cx43
2203
Phorbol esters such as 12-O-tetradeconylphorbol-13-
acetate (TPA) activate protein kinase C, increase
Connexin43 (Cx43) phosphorylation, and decrease cell-cell
communication via gap junctions in many cell types.
Previous work has implicated protein kinase C (PKC) in
the direct phosphorylation of Cx43 at S368, which results
in a change in single channel behavior that contributes to
a decrease in intercellular communication. We have
examined Cx43 phosphorylation in several cell lines with
an antibody specific for phosphorylated S368. We show that
this antibody detects Cx43 only when it is phosphorylated
at S368 and, consistent with previous results, TPA
treatment causes a dramatic increase in phosphorylation
at S368. However, in some cell types, the increased
phosphorylation at S368 did not cause a detectable shift
in migration as compared with the nonphosphorylated
Cx43. Immunofluorescence showed increased S368
immunolabeling in cytoplasmic and plasma membrane
structures in response to TPA. Immunoblot analysis of
synchronized cells showed increased phosphorylation at
S368 during S and G2/M phases of the cell cycle. S-phase
cells contained more total Cx43 but assembled fewer
functional gap junctional channels than G0-phase cells.
Since M-phase cells also communicate poorly and contain
few assembled gap junctions, phosphorylation at S368
appears to be negatively correlated with gap junction
assembly. Thus, both gap junctional communication and
S368 phosphorylation change during S phase and G2/M,
implying that phosphorylation at S368 might play a role in
key cell-cycle events.
Key words: Gap junctions, Connexins, Tumor promoter,
Phosphorylation, Carcinogenesis
Summary
Connexin43 phosphorylation at S368 is acute during S
and G2/M and in response to protein kinase C
activation
Joell L. Solan1, Matthew D. Fry2, Erica M. TenBroek3and Paul D. Lampe1,*
1Fred Hutchinson Cancer Research Center, Seattle, WA 98109, USA, and Department of Pathobiology, University of Washington, WA 98195, USA
2Cell Signaling Technology, Beverly, MA 01915, USA
3Genetics, Cell Biology and Development, University of Minnesota, St Paul, MN 55108, USA
*Author for correspondence (e-mail: plampe@fhcrc.org)
Accepted 12 February 2003
Journal of Cell Science 116, 2203-2211 © 2003 The Company of Biologists Ltd
doi:10.1242/jcs.00428
Research Article
2204
phosphorylation and decrease gap junction communication in
several different cell types (Berthoud et al., 1992; Berthoud et
al., 1993; Brissette et al., 1991; Lampe, 1994; Reynhout et al.,
1992). PKC has been shown to phosphorylate Cx43 at S368,
and this site has been shown to underlie a TPA-induced
reduction in intercellular communication and alteration of
single channel behavior (Lampe et al., 2000). However, in
some cell types, TPA treatment did not lead to a shift in Cx43
mobility in SDS-PAGE (which is thought to indicate increased
Cx43 phosphorylation) but did change gap junctional
communication (e.g. Rivedal and Opsahl, 2001), leading to
confusion as to the role of Cx43 phosphorylation in this
process.
Here, we report that Cx43 phosphorylation at S368 was
indeed increased by TPA treatment in all cell types tested, but
that Cx43 mobility was not significantly affected in some.
Furthermore, S368 phosphorylation was increased during key
stages of the cell cycle where gap junctional assembly is
reduced. Thus, in addition to its role in the regulation of
gap junction channel gating, phosphorylation at S368 was
negatively correlated with gap junction assembly.
Materials and Methods
Cell line maintenance and transfection
Normal rat kidney (NRK) epithelial cells (NRK-E51, American Type
Culture Collection, ATCC, Rockville, MD), Chinese Hamster Ovary
(CHO) and HeLa cells were cultured in DMEM (Mediatech,
Pittsburgh, PA) supplemented with 5% fetal calf serum and antibiotics
in a humidified 5% CO2environment. A serine-to-alanine site 368
mutant Cx43 cDNA was generated using the Chameleon double-
stranded, site-directed mutagenesis kit (Stratagene, La Jolla, CA) and
was then subcloned into the bicistronic expression vector pIREShyg
(Clontech Laboratoroes, Palo Alto, CA) and transfected into the HeLa
cell lines. Stably transfected clones were isolated by repeated dilution
subcloning in the presence of the selective antibiotic hygromycin (200
µg/ml).
Metabolic labeling and Cx43 immunoprecipitation
NRK cells were cultured, metabolically labeled with
[32P]orthophosphate (ICN, 64014L) or 35S-Trans label (ICN,
5100607), and immunoprecipitated essentially as previously
described (TenBroek et al., 2001). Briefly, cells were labeled with
[32P]orthophosphate at 1.0 mCi/ml for 3 hours in phosphate-deficient
medium (Gibco-Invitrogen, Grand Island, NY) and, where indicated,
were treated with 50 ng/ml TPA during the final 30 minutes.
Alternatively, cells were washed three times and labeled with 35S-
Trans label at 0.1 mCi/ml for 3 hours in methionine-free media
(Gibco-Invitrogen) and, where indicated, were treated with 50 ng/ml
TPA during the final 30 minutes. The cells were rinsed in PBS, lysed
in RIPA buffer [25 mM Tris-HCl, 100 mM NaCl, 10 mM EDTA,
50 mM NaF, 500 µM Na3VO4, 0.25% Triton X-100, 2 mM
phenylmethylsulphonyl fluoride (PMSF) and 1×Roche Complete
protease inhibitors], clarified with protein A beads, and
immunoprecipitated with p368 antibody [a rabbit anti-Phospho-Cx43
(Ser368) antibody #3511; Cell Signaling Technology, Beverly, MA],
rabbit antibody C6219 from Sigma (St Louis, MO) and/or monoclonal
Cx43CT1 antibody. Cx43CT1 antibody is an antibody prepared to a
peptide representing the last 23 amino acids of Cx43 (described in
Cooper and Lampe, 2002). Cx43CT1 behaves like antibody 13-8300
from Zymed, which was prepared to the same region of Cx43, in that
it immunoprecipitates primarily the ‘NP’ form of Connexin unless
cells are treated with TPA, when slower migrating forms were
detected (Cruciani and Mikalsen, 1999). After four washes in RIPA
buffer, the immunoprecipitates were treated with Laemmli sample
buffer and run via SDS-PAGE (10% polyacrylamide, Tris-glycine
gels).
Immunoblotting
Cells were lysed in sample buffer containing 50 mM NaF, 500 µM
Na3VO4, 2 mM PMSF and 1×Complete protease inhibitors (Roche
Diagnostics, Indianapolis, IN) and cellular proteins were separated by
SDS-PAGE on 10% Tris-glycine gels. For alkaline phosphatase
treatment, cells were lysed in 0.2% SDS, 2 mM PMSF and 1×protease
inhibitors, and briefly sonicated followed by addition of one-tenth
volume of 10×phosphatase buffer (M183A; Promega, Madison, WI)
and incubation with 10 units of calf intestinal alkaline phosphatase
(M182A; Promega) for 1 hour at 37°C. After electrophoresis, protein
was transferred to nitrocellulose, the membrane was blocked, and
antibodies were incubated as previously indicated (Lampe et al.,
1998a). Primary and secondary antibodies utilized were p368
antibody, mouse anti-Cx43 (Cx43NT1 described in Goldberg et al.,
2002), mouse anti-vinculin (Sigma), peroxidase-conjugated donkey
anti-mouse or mouse anti-rabbit secondary antibodies (Jackson
Immunoresearch Laboratories, West Grove, PA). Where indicated, the
blots were ‘stripped’ for 30 minutes at 50°C in 62.5 mM Tris pH 6.8,
1% SDS and 5% β-mercaptoethanol buffer followed by washing for
2 hours with at least six changes of PBS. Signal was visualized with
SuperSignal West Pico or Femto Chemiluminescent Substrate (Pierce
Chemicals, Rockford, IL) followed by exposure to Kodak Biomax
MR film. Densitometry of autoradiographs was performed on a
Macintosh G3 using a Sharp JX-325 scanner to collect the image and
the public domain NIH Image program (developed at the US National
Institutes of Health and available at http://rsb.info.nih.gov/nih-image).
Immunofluorescence
NRK cells were untreated or treated with TPA for 30 minutes at 37°C,
washed twice in PBS, and fixed in cold methanol/acetone (50:50) for
1 minute followed by blocking for 1 hour in 1% bovine serum albumin
in PBS. Cells were incubated with anti-Cx43 antibody p368 and/or
Cx43IF1 (see Cooper and Lampe, 2002; TenBroek et al., 2001) in
blocking solution for 1 hour. Following several PBS washes, the
cultures were incubated with Alexa594-conjugated goat anti-rabbit
antibody (Molecular Probes, Eugene, OR) and/or fluorescein
isothiocyanate-conjugated donkey anti-mouse antibody (Jackson
Immunoresearch Laboratories) for 30-60 minutes and counterstained
with DAPI (Molecular Probes), followed by several washes in PBS.
The coverslips were mounted onto slides with DABCO antifade
medium [25 mg/ml of 1,4-diazobicyclo-(2,2,2)octane (Sigma) diluted
in Spectroglycerol (Kodak) and 10% PBS, pH 8.6] and viewed with
a Nikon Diaphot TE300 fluorescence microscope, equipped with a
40×(1.3 n.a.) oil objective and a Princeton Instruments cooled digital
camera driven by an attached PC and Metamorph imaging software.
Cell synchronization
G0cells were prepared by contact-inhibiting NRK cells at least 3 days
past confluency without addition of fresh media. To obtain G1 cells,
confluent cells were trypsinized, then diluted to 60-80% confluency
and allowed to progress 8-10 hours for early G1and 14-16 hours for
late G1. Cell-cycle analysis showed that by 18 hours these cells begin
to enter S phase. For preparation of G1/S, S, G2and G2/M cells,
confluent cells were trypsinized then diluted to 60-80% confluency in
media containing 1 mM thymidine for 16 hours to induce a G1/S
block. Cells were released from G1/S by washing and replacement of
37°C complete media. Cell-cycle analysis showed that S phase lasts
4-6 hours in these cells and that cells cycle through G2/M to G1by 9-
11 hours after washout. We typically observed 70-90% synchrony as
Journal of Cell Science 116 (11)
2205Phosphorylation at S368 of Cx43
cells progress through S to G1. Cell-cycle analysis was performed by
fluorescence activated cell sorting. Specifically, cells were trypsinized,
then pelleted in PBS with 2% fetal bovine serum and fixed in 70%
EtOH. Cells were pelleted, washed and incubated with 5 µg/ml RNase
at 37°C for 30 minutes, and then stained with 50 µg/ml propidium
iodide on ice for 1 hour. DNA content was assessed on a Becton
Dickinson FACScalibur and data analyzed using CellQuest software.
Gap junctional communication/assembly
Gap junctional communication was assayed via dye transfer according
to published methods using either an assembly-preloading assay with
calcein-AM (Lampe et al., 1998b) or by microinjection of fluorescent
dyes. Briefly, for the preloading assay, one 10 cm plate of NRK cells
was labeled with 0.5 µM calcein-AM (Molecular Probes), the cell-
permeant ester of calcein that is cleaved to membrane-impermeant
calcein by cellular esterases. Three other culture plates were labeled
with 0.25 µM DiI (Molecular Probes). After washing twice with PBS,
the two populations of cells were each trypsin/EDTA suspended,
treated with trypsin inhibitor and pelleted. The cells were suspended
in the appropriate media, mixed, plated on culture dishes and placed
in a 37°C incubator. Cells were allowed to adhere for 2 hours then
digital images of calcein and DiI were captured. The assignment of a
cell as an acceptor of dye via transfer rather than a poorly loaded or
leaking donor is checked by digitally overlaying images of DiI and
calcein fluorescence. If a cell adjacent to a calcein-loaded, DiI-
negative cell contains both punctate DiI and more-diffuse calcein
fluorescence, gap junction assembly and dye transfer occurred. If a
DiI-labeled cell adjacent to a calcein-loaded cell does not contain
calcein, then dye transfer did not occur at that interface. A more-
complete description of this assay is published elsewhere (Lampe,
1994; Lampe et al., 1998b). The fraction of cells that transferred dye
were determined by dividing the number of DiI-labeled cells that
contained calcein (i.e. transfers) by the number of cell interfaces
between calcein-loaded and DiI-labeled cells (i.e. total).
Dye transfer in established cultures was analyzed by microinjection
of a 10 mM solution of each of the gap junction permeable dyes,
Alexa hydrazide 488 and 594 (Mr=570.5 and 758.8, respectively;
Molecular Probes) in 0.2 M KCl. The dyes were microinjected using
a 5 millisecond pulse of air at 10 psi from a General Valve Picospritzer
II, and the number of cells receiving dye was analyzed after 10
minutes using the imaging system described above.
Results
Phospho-Cx43-Ser368 (p368) antibody is specific for
phosphorylation of S368
We have developed an antibody that reacts with Cx43 when it
is phosphorylated at S368. To test the specificity of the p368
antibody, HeLa cells that did not express any Cx43 were stably
transfected with wild-type (wt) Cx43 or Cx43 containing a
serine-to-alanine substitution at position 368 (S368A) and
were examined by immunoblot analysis. Previously, we have
shown that cells treated with TPA showed increased
phosphorylation on Cx43, especially at S368 (Lampe et al.,
2000). HeLa cells, either treated with TPA for 30 minutes or
untreated, were washed and directly lysed in sample buffer, and
whole-cell lysates were immunoblotted and probed for Cx43
content using the p368 antibody (rabbit) followed by
stripping/reprobing of the blot using the mouse monoclonal
Cx43 antibody Cx43NT1 (Fig. 1). The anti-Cx43 probing of
HeLa cells expressing wt Cx43 or S368A Cx43 (Fig. 1, α-
Cx43, CON lanes) showed typical migration patterns of Cx43
in SDS-PAGE, with multiple bands representing different
phosphorylation states of Cx43. The predominant
nonphosphorylated (NP) form migrates fastest, followed by
slower-migrating phosphorylated forms, often referred to as
P1, P2, etc., which can be converted to the faster-migrating
form via alkaline phosphatase treatment (Berthoud et al., 1992;
Brissette et al., 1991; Kadle et al., 1991; Laird et al., 1991;
Lampe, 1994; Musil et al., 1990). Upon TPA treatment,
essentially all of the Cx43 migration in HeLa cells containing
wt Cx43 was shifted to slower-migrating species as has been
observed previously in many different cell types (reviewed by
Lampe and Lau, 2000). HeLa cells containing Cx43 with a
S368A site-directed mutation responded much less extensively
to TPA treatment (Fig. 1, α-Cx43), indicating that
modification/phosphorylation on S368 affects the mobility
shift, at least in these HeLa cells. Probing this same blot
with the p368 antibody (Fig. 1, α-p368) showed that cells
containing wt Cx43 had a low level of Cx43 phosphorylated at
S368 (CON) that appeared to migrate similarly to NP Cx43.
Upon TPA treatment, both a tenfold increased signal and a shift
in migration of wt Cx43 was observed with the p368 antibody.
Lysates from cells containing Cx43 with the S368A mutation
showed no p368 antibody reactivity regardless of TPA
treatment or long exposure of blots.
To verify further that this antibody recognized a
phosphorylated species of Cx43, NRK cell lysates from control
cells and TPA-treated cells were incubated with alkaline
phosphatase and analyzed by immunoblot. As above,
immunoblots were first processed with the p368 antibody, then
stripped and reprobed with the anti-Cx43 antibody, allowing
precise alignment and determination of the extent of migration
of the bands. Immunoblots incubated with anti-Cx43 showed
little change in response to TPA treatment (Fig. 2, α-Cx43
panel; compare CON and TPA lanes). When these cell lysates
were incubated with alkaline phosphatase, all of the Cx43
migrated as the NP form (Fig. 2, α-Cx43 panel, AP lanes),
which is consistent both with effective alkaline phosphatase
treatment and with what has been shown by other investigators,
as noted above. Processing of the blot with anti-p368 antibody
Fig. 1. The p368 antibody reacts with Cx43 only when S368 is
present. Shown is an immunoblot of whole cell lysates from HeLa
cells transfected with wild-type (wt) Cx43 or Cx43 containing a
S368A mutation. Cells were either incubated in the presence (TPA)
or absence (CON) of 50 ng/ml TPA for 30 minutes. The immunoblot
was probed with either an antibody to the N-terminal region of Cx43
(α-Cx43) or the anti-p368 antibody (α-p368). Positions of the
molecular weight markers are shown on the left.
2206
showed a sixfold increase in signal upon TPA treatment (Fig.
2, α-p368 panel; compare CON and TPA) and this signal was
completely lost upon alkaline phosphatase treatment (Fig. 2,
α-p368 panel, AP lanes). These data show that the p368
antibody reactivity appears to be specific for S368 only when
it is phosphorylated (i.e. it is phosphorylation-state specific)
and that a dramatic increase in phosphorylation at S368 is
generated in response to TPA.
The ‘NP’ form of Cx43 can be phosphorylated on S368
Figs 1 and 2 indicate that an isoform of Cx43 that migrated
similarly to NP Cx43 reacted with the p368 antibody. To
determine more directly whether a phosphorylated species
migrated to the same extent as the NP form of Cx43, we
performed metabolic labeling on NRK cells with
[32P]orthophosphate or [35S]methionine. Cx43 from
[32P]orthophosphate-labeled cells was immunoprecipitated,
run on SDS-PAGE and blotted to nitrocellulose. These samples
were analyzed first by autoradiography (Fig. 3, 32P panel)
and then immunoblot analysis using p368 (α-p368 panel)
and Cx43NT1 (α-Cx43) monoclonal antibodies. In the
autoradiograph, Cx43 immunoprecipitated from untreated
cells showed two band, indicated as P1 and P2, whereas cells
treated with TPA showed a more-broad phosphorylation
pattern some of which appeared to migrate at the same position
as the NP form. The α-p368 panel, which represents the
chemiluminescent signal obtained from the same blot probed
with p368 antibody, shows a dramatic TPA-dependent increase
in signal co-migrating with the NP form, whereas probing the
same blot with the α-Cx43 antibody showed minor differences
in the typical pattern for Cx43 with or without TPA treatment.
Thus, the TPA-dependent increase in Cx43 phosphorylation
levels found by autoradiography was not nearly as extensive as
that observed with the p368 antibody immunoreactivity. This
result confirms that S368 phosphorylation, in particular, is
increased dramatically via TPA treatment, whereas
phosphorylation at many other residues was not as TPA
responsive (Lampe et al., 2000), essentially diluting the p368
signal. Furthermore, the total Cx43 signal and the ratio of
‘phosphorylated’ (i.e. P1 + P2) to nonphosphorylated (Fig. 2,
α-368 panel) were quite similar regardless of TPA treatment,
in spite of the fact that dramatic changes in S368
phosphorylation occurred.
NRK cells were also labeled with [35S]methionine, and
immunoprecipitations were carried out using either the p368
antibody or an anti-Cx43 (Cx43CT1) antibody that shows a
strong preference for the NP migratory isoform. These
samples were run on SDS-PAGE and analyzed by
autoradiography. In NRK cells, the α-Cx43 antibody
immunoprecipitated a single band that did not change
significantly in intensity upon TPA treatment (Fig. 4, NRK
panel, α-Cx43). As expected, this band migrates at the same
position as NP. The p368 antibody also immunoprecipitated a
single band that migrated exactly with the band
immunoprecipitated with the α-Cx43 antibody and showed
increased signal intensity upon TPA treatment (Fig. 4, NRK
panel, α-p368). Thus, the p368 antibody was able to
immunoprecipitate a Cx43 isoform that migrates the same as
the NP form in our typical Laemmli gel system, and TPA-
treated cells contained more of this isoform than control cells.
Taken together, these metabolic labeling data show that a
phosphorylated species of Cx43 essentially co-migrates with
the NP form and that this phosphoform can be clearly detected
using the p368 antibody. Nonphosphorylated Cx43 and Cx43
phosphorylated at S368 probably could be separated given the
appropriate separation technique since these species vary in net
charge. It is noteworthy that standard isoelectric focusing and
two-dimensional analysis of Cx43 has been shown to be
Journal of Cell Science 116 (11)
Fig. 2. The p368 antibody reacts with Cx43 only when it is
phosphorylated at S368. Untreated (CON) or TPA-treated cells were
lysed in sample buffer or treated with alkaline phosphatase (+AP
lanes) prior to SDS-PAGE and immunoblotting. The blot was probed
with the p368 antibody (α-p368 panel) followed by stripping and
reprobing with the Cx43 antibody (α-Cx43 panel).
Fig. 3. TPA-treated NRK cells show increased phosphorylation at
S368 with no apparent shift in migration in SDS-PAGE. NRK cells
were metabolically labeled with [32P]orthophosphate and incubated
in the presence (TPA) or absence (CON) of 50 ng/ml TPA for 30
minutes followed by immunoprecipitation of Cx43, blotting to
nitrocellulose, and probing the blot first by autoradiography (32P) and
then using the anti-p368 antibody (α-p368) and ultimately the N-
terminal Cx43 antibody (α-Cx43).
Fig. 4. The p368 antibody immunoprecipitates more Cx43 from
either NRK or CHO cells after TPA treatment. NRK and CHO cells
were metabolically labeled with [35S]methionine and either treated
with TPA or left untreated (CON) and then lysed and
immunoprecipitated with either the Cx43 antibody (α-Cx43 lanes) or
the p368 antibody (α-p368 lanes).
2207Phosphorylation at S368 of Cx43
difficult (Stockert et al., 1999). Nevertheless, when analyzed
by a standard Tris/glycine SDS-PAGE system that has been
used by most investigators, the migration of Cx43
phosphorylated on S368 often coincided with
nonphosphorylated Cx43. For this reason, we believe that it is
probably most accurate to refer to the fastest-migrating form
as P0 rather than NP when discussing standard SDS-PAGE
separation of Cx43 from cells that have been treated with
kinase effectors or growth factors, and we do so below. This is
an interim solution since different cell types and slightly
modified gel systems appear to produce Cx43 with varying
migratory properties. A better definition of terms will probably
require a thorough understanding of the molecular events that
underlie the shift in migration.
The TPA-dependent shift in Cx43 migration, but not
phosphorylation of S368, is cell-type specific
Although the migration of Cx43 derived from NRK cells does
not shift significantly in the presence of TPA, many other cell
types can show a dramatic shift, essentially leaving little faster-
migrating species as shown for HeLa cells in Fig. 1. We found
that CHO cells also show a dramatic shift in response to TPA,
as is shown via immunoprecipitation in Fig. 4 and western
immunoblot in Fig. 5 (CHO panels). Fig. 4 shows
[35S]methionine-labeled CHO cell lysates immunoprecipitated
with α-Cx43 or α-p368 antibodies. In TPA-treated CHO cells,
α-Cx43 (Cx43CT1) immunoprecipitated both the NP/P0
migratory isoform and the slower-migrating isoforms (Fig. 4,
CHO panel, α-Cx43; see Materials and Methods for antibody
description). Immunoprecipitation of TPA-treated CHO cell
lysates with α-p368 antibody shows primarily the slower-
migrating isoforms (Fig. 4, CHO panel, α-p368) indicating
that, in this cell line, phosphorylation on S368 was coincident
with a shift in migration.
Similarly, Fig. 5 shows an immunoblot of NRK and CHO
whole cell lysates that was probed with the antibody for p368
(α-p368) and stripped/reprobed for Cx43 (α-Cx43). Consistent
with the immunoprecipitation results, Cx43 from NRK cells
did not shift its migration in response to TPA while the protein
extensively shifts to slower-migrating phosphoforms in CHO
cells. Both cell types show large TPA-dependent increases in
reactivity to the p368 antibody, but the p368 signal in NRK
cells primarily migrated at the P0 and P1 positions whereas the
p368 signal was highly shifted in the CHO cells. HeLa cells
containing wt Cx43 (Fig. 1) were intermediate between the
two, as p368 is found in the P1 and P2 forms. Notably, in all
cell lines examined, a low level of p368 was present in
untreated cells and often co-migrated with the NP/P0 isoform,
which indicates that phosphorylation of S368 is part of the
normal lifecycle of Cx43 in these cells. To examine whether
phosphorylation at S368 might be consistent with the early
phosphorylation event found in the presence of Brefeldin A
(BFA) (Laird et al., 1995), we treated NRK and CHO cell
lysates with BFA and found decreased p368 antibody labeling.
However, α-p368 binding was not eliminated, so no firm
conclusions can be drawn with respect to this event. Thus, we
have used the p368 antibody to examine TPA-induced
phosphorylation of Cx43 in NRK, CHO and Cx43-transfected
HeLa cells and found that all cell types examined show
increased phosphorylation on S368, but the degree to which
this resulted in a shift in the migration of Cx43 varied between
cell types.
TPA-induced phosphorylation of S368 occurs on both
intracellular and plasma membrane Cx43
To determine whether a specific pool of Cx43 is
phosphorylated in response to TPA, immunofluorescence was
performed on NRK cells with an antibody specific for Cx43
(Cx43IF1) and the p368 antibody. NRK cells show extensive
immunofluorescence for Cx43 at cell-cell interfaces (Fig. 6,
upper left). Upon TPA treatment, Cx43 immunofluorescence
showed no apparent change although the cells adopted a
slightly more fibroblastic appearance (Fig. 6, lower left). The
p368 antibody also showed some cell-cell interface labeling
and a light reticulate pattern throughout the cytoplasm (upper
center panel). The apparent cytoplasmic pool of p368 staining
does appear to be at least partly associated with the
endoplasmic reticulum as there was co-localization of p368
with an endoplasmic reticulum-specific dye, R6 (data not
shown). After TPA treatment, the p368 signal was greatly
increased in both cytoplasmic and interface membranes (lower
center panel). The plasma membrane pool of p368 shows co-
localization with the Cx43IF1 antibody, whereas less-distinct
co-localization of this antibody with the intracellular pool was
observed. The increase in intracellular fluorescence does
appear to be specific to the p368 epitope as co-incubation of
p368 antibody with the peptide antigen used to generate the
antibody blocked antibody binding, while co-incubation of the
antibody with a nonphosphorylated peptide representing 360-
382 of Cx43 did not block binding (data not shown). We have
observed that there is competition between Cx43IF1 and p368
antibody binding at cell-cell contacts. This was manifest by a
decrease in p368 signal when p368 and Cx43IF1 antibodies
were added together, but was reversed by inclusion of the
nonphosphorylated peptide, which removed the Cx43IF1
signal.
Fig. 5. NRK and CHO cells both show large increases in
phosphorylation at S368 upon TPA treatment but the resulting Cx43
mobilities are very different. Cells were either incubated in the
presence of no drugs (CON), 50 ng/ml TPA for 30 minutes (TPA) or
5 µg/ml brefeldin A for 4 hours (BFA), and processed for
immunoblot and separately probed with the anti-p368 antibody (α-
p368) and the N-terminal Cx43 (α-Cx43).
2208
Phosphorylation on S368 is regulated as cells progress
through the cell cycle
Given that phosphorylation on S368 appeared to be part of the
normal lifecycle of Cx43, we wanted to determine
circumstances under which this event was regulated. As it has
previously been shown that Cx43 phosphorylation increases as
cells progress through the cell cycle (Kanemitsu et al., 1998),
we looked at Cx43 and phosphorylation of S368 in cells
synchronized at different stages of the cell cycle in NRK cells.
Fig. 7 shows an immunoblot probed first for p368 (α-p368)
and then stripped/reprobed for Cx43 (α-Cx43). Vinculin was
also detected for a loading control. Densitometry was
performed for Cx43 and p368 antibody binding, and the ratio
of p368/Cx43 densitometry is shown at the bottom of the
figure. Cx43 phosphorylated at S368 was most abundant
relative to total Cx43 during S and G2/M. This result is
consistent with previous reports where gap junctional
communication was shut-down during mitosis (Stein et al.,
1992; Xie et al., 1997) and phosphorylation at S368 had been
shown to reduce communication (Lampe et al., 2000). Here,
we found that G0 cells contain very little p368 and that S368
is increasingly phosphorylated as cells approach and progress
through S phase.
Given the 7×increase in phosphorylation at S368 when G0
and S phase cells were compared (Fig. 7), we wanted to
examine Cx43 distribution and intercellular communication in
these two cell populations. G0cells showed strong plasma
membrane staining for Cx43 at cell-cell interfaces consistent
with gap junctions (Fig. 8A), while S-phase cells showed both
typical gap junctional labeling and also extensive perinuclear
staining (Fig. 8B). Immunofluorescent labeling with the p368
antibody showed both cytoplasmic and gap junctional staining
for both G0- and S-phase cells (data not shown).
Since TPA treatment of cells has been reported to decrease
intercellular communication via changes in channel gating
(e.g., Kwak et al., 1995c; Lampe et al., 2000; Moreno et al.,
1994) and gap junction assembly (Lampe, 1994), we assessed
both the ability to transfer dye and the ability to assemble
junctions in G0- and S-phase cells. When we microinjected G0-
and S-phase cells with two fluorescent dyes of the Alexa series
(A488, Mr=570.5; A594, Mr=758.8), we found that S-phase
cells transferred both dyes approximately twice as well as G0
cells (Fig. 8C). However, G0-phase cells were approximately
twice as likely to transfer dye to their neighbors than S-phase
Journal of Cell Science 116 (11)
Fig. 6. The p368 antibody binds to both junctional and cytoplasmic membranes. NRK cells that had been incubated in the presence (TPA) of 50
ng/ml TPA or absence (CON) were processed for immunofluorescence with the anti-p368 (α-pS368) and the Cx43IF1 (α-Cx43) antibodies (left
and center panels) or with the anti-p368 antibody plus the immunizing peptide (right panel, α-pS368 + peptide).
Fig. 7. The extent of p368 phosphorylation is increased through the
cell and is maximal during S and G2/M. Synchronized cells collected
at the indicated cell-cycle stage were processed for immunoblotting
and probed with antibodies to p368 (α-p368), Cx43NT1 (α-Cx43),
or vinculin (for a loading control). The molecular weight or
migration position of the Cx43 is indicated on the right, and the ratio
of the extent of p368 to Cx43NT1 antibody labeling is shown on the
bottom line.
2209Phosphorylation at S368 of Cx43
cells when the calcein/DiI assay, which requires nascent gap
junction assembly, was performed (Fig. 8D).
Discussion
Previously, we have demonstrated that phosphorylation of
Cx43 on S368 is stimulated by TPA in vivo and mediated by
PKC in vitro (Lampe et al., 2000). Furthermore,
electrophysiological studies of Cx43 and the Cx43-S368A
mutant revealed that phosphorylation at S368 is necessary for
a TPA-induced alteration of Cx43 channel behavior that
contributes to decreased gap junctional communication. Here,
we report that phosphorylation levels at S368 are high in S and
G2/M, and that cells at quiescence show only very low levels.
In addition, S-phase cells assembled gap junctions poorly
compared with G0-phase cells, implying a role for
phosphorylation at S368 in the regulation of assembly.
Cx43 phosphorylation at S368 appears to occur normally in
dividing cells. The seven- to eightfold increase in the level of
phosphorylation on S368 at S and G2/M, respectively, correlates
well with increased cytoplasmic localization of Cx43 during S
(Fig. 7) and G2/M (Lampe et al., 1998a; Xie et al., 1997),
consistent with a role for S368 phosphorylation in regulating
Cx43 trafficking/assembly into gap junctional structures.
Interestingly, we also occasionally observed a unique
and apparently nuclear envelope/endoplasmic reticulum
localization of the p368 antibody at the early stages of G2/M
(data not shown). This immunolocalization was highly
transitory because it was lost as the nuclear envelope broke
down as the cells entered mitosis. Although this localization
appeared specific for the antibody based on antigen competition
studies, Cx43IF1 antibody immunolabeling of the nuclear
envelope region of G2/M cells was not nearly as striking as the
p368 antibody. Thus, we cannot rule out the possibility that an
alternative non-connexin epitope that specifically reacts with
the p368 antibody is expressed in early mitosis.
Much of the work examining TPA-mediated downregulation
of Cx43 has been motivated by the role of PKC activators as
tumor promoters and the potential role of gap junctional
communication as a tumor suppressor. Although details are
still poorly understood, there is a wealth of data showing that
Cx43 phosphorylation is increased and gap junctional
communication is reduced upon activation of PKC (reviewed
by Lampe and Lau, 2000). However, the use of different assays
for communication, several methods for assaying Cx43
phosphorylation and various cellular systems expressing
different isoforms of PKC (Cruciani et al., 2001; Munster and
Weingart, 1993) have confused the interpretation of the role
PKC plays as a modulator of gap junctional communication.
For example, several reports have fueled the controversy as to
whether mitogen-activated protein kinase (MAPK) or PKC is
the actual kinase that phosphorylates Cx43 and reduces gap
junctional communication after growth factor or phorbol ester
treatment, or whether Cx43 phosphorylation even plays a direct
role (e.g., Hossain et al., 1999; Kanemitsu and Lau, 1993;
Rivedal and Opsahl, 2001; Vikhamar et al., 1998). One
presumption found in many of these reports that might cloud
interpretation of the data is that a shift in Cx43 migration has
been equated with increased phosphorylation. We know that
Cx43 can be phosphorylated at many (>5) sites in untreated
cells and at many more sites in growth factor-treated cells
(Lampe and Lau, 2000). At this time, we have no
understanding of the molecular events responsible for the shift
in migration, or of any of the serines involved. By comparing
two cell types where TPA led to a shift in Cx43 migration in
one but no change in another, the logical but potentially
erroneous conclusion could be that Cx43 phosphorylation
levels only changed in one of the cell types. For example, from
the α-Cx43 panel of Fig. 5, one could conclude that there was
a large change in Cx43 phosphorylation in CHO cells upon
TPA treatment, whereas NRK cells showed little change and
thus appeared unresponsive to TPA treatment; by contrast, the
α-p368 panel shows that phosphorylation was dramatically
increased at this site in NRK cells. In fact, there probably is
some correlation with the extent of shift and the overall level
of Cx43 phosphorylation. However, specific phosphorylation
events and not the overall level of phosphorylation probably
elicit a specific regulatory event such as assembly, disassembly
or gating changes. We believe re-evaluation of many of these
seemingly conflicting results might be resolved by assaying for
TPA and growth factor effects with the p368 and other
phosphorylation-site-specific antibodies.
Fig. 8. G0- and S-phase cells show different Cx43 cellular
distributions and different abilities to transfer fluorescent dyes in
established and junctional assembly assays. (A) G0cells show
extensive junctional Cx43 immunostaining. (B) S-phase cells show
extensive junctional and cytoplasmic membrane staining. (C) The
number of cells that receive either Alexa488 (A488) or Alexa594
(A594) from the injected cell is quantitated for cells in established
G0- or S-phase cultures (mean±s.d.). (D) Cells in G0or S were
assayed for the ability to assemble gap junctions and transfer calcein
(quantitated as the number of transfers per the total number of
interfaces between a calcein-loaded and a recipient cell, mean±s.d.).
2210
Phosphorylation of Cx43 appears to regulate the trafficking
of Cx43 to the plasma membrane, assembly of Cx43 into gap
junctional structures, single channel behavior and Cx43
degradation. The latter three events have been reported to be
sensitive to TPA and, therefore, could be regulated by PKC
(Kanemitsu and Lau, 1993; Kwak et al., 1995a; Kwak et al.,
1995c; Lampe, 1994). Our immunofluorescence data with the
p368 antibody and comparison of the kinetics of the mobility
shift in SDS-PAGE with decreases in gap junctional
communication (Kanemitsu and Lau, 1993) indicate that S368
phosphorylation and potentially other PKC-mediated events
can occur prior to export to the plasma membrane
(Lampe, 1994). Therefore, at least some TPA-dependent
phosphorylation at S368 occurs prior to gap junction assembly.
Intercellular communication was reduced by TPA in
quiescent but not proliferating NRK cells (Paulson et al.,
1994). Data presented here indicates that, in addition to S368
being a TPA-responsive site, there is regulation of S368
phosphorylation during the normal lifespan of Cx43 in
untreated, cycling cells. Although S-phase cells transferred dye
more rapidly than G0cells in established cultures, S-phase
cultures were less able to form new functional gap junctions in
an assembly assay (Fig. 8D). Clearly, cell-cycle regulation
plays a key role during tumorigenesis. The cell-cycle-mediated
regulation shown here might indicate a more-subtle and
physiological role for gap junctional communication through
S368-mediated effects on assembly. Cell-cycle-mediated
regulation of Cx43 has been shown during mitosis, when there
is a dramatic change in phosphorylation and Cx43 is localized
predominately to cytoplasmic membranes. A model in which
assembly is most efficient during G0/G1, and then decreases as
cells progress towards mitosis, this being partially due to
phosphorylation at S368, fits our data. During tumorigenesis
or faulty regulation of the cell cycle, this decrease in assembly
could have dramatic effects on gap junctional communication.
This work was supported by grants GM55632 (P.D.L.) and
GM46277 (R.G.J.) from the National Institutes of Health.
References
Atkinson, M. M., Menko, A. S., Johnson, R. G., Sheppard, J. R. and
Sheridan, J. D. (1981). Rapid and reversible reduction of junctional
permeability in cells infected with a temperature-sensitive mutant of avian
sarcoma virus. J. Cell Biol. 91, 573-578.
Azarnia, R. and Loewenstein, W. R. (1984). Intercellular communication and
control of cell growth. X. Alteration of junctional permeability by the src
gene. J. Membr. Biol. 82, 191-205.
Berthoud, V. M., Ledbetter, M. L. S., Hertzberg, E. L. and Saez, J. C.
(1992). Connexin43 in MDCK cells: Regulation by a tumor-promoting
phorbol ester and calcium. Eur. J. Cell Biol. 57, 40-50.
Berthoud, V. M., Rook, M., Hertzberg, E. L. and Saez, J. C. (1993). On the
mechanism of cell uncoupling induced by a tumor promoter phorbol ester
in clone 9 cells, a rat liver epithelial cell line. Eur. J. Cell Biol. 62, 384-396.
Bittman, K. S. and LoTurco, J. J. (1999). Differential regulation of connexin
26 and 43 in murine neocortical precursors. Cereb. Cortex 9, 188-195.
Brissette, J. L., Kumar, N. M., Gilula, N. B. and Dotto, G. P. (1991). The
tumor promoter 12-O-tetradecanoylphorbol-13-acetate and the ras
oncogene modulate expression and phosphorylation of gap junction
proteins. Mol. Cell. Biol. 11, 5364-5371.
Cooper, C. D. and Lampe, P. D. (2002). Casein kinase 1 regulates connexin43
gap junction assembly. J. Biol. Chem. 277, 44962-44968.
Crow, D. S., Beyer, E. C., Paul, D. L., Kobe, S. S. and Lau, A. F. (1990).
Phosphorylation of connexin43 gap junction protein in uninfected and Rous
sarcoma virus-transformed mammalian fibroblasts. Mol. Cell. Biol. 10,
1754-1763.
Cruciani, V. and Mikalsen, S. O. (1999). Stimulated phosphorylation of
intracellular connexin43. Exp. Cell Res. 251, 285-298.
Cruciani, V., Sanner, T. and Mikalsen, S. O. (2001). Pharmacological
evidence for system-dependent involvement of protein kinase C isoenzymes
in phorbol ester-suppressed gap junctional communication. Carcinogenesis
22, 221-231.
de Feijter, A. W., Ray, J. S., Weghorst, C. M., Klaunig, J. E., Goodman,
J. I., Chang, C. C., Ruch, R. J. and Trosko, J. E. (1990). Infection of rat
liver epithelial cells with c-Ha-ras: Correlation between oncogene
expression, gap junctional communication and tumorigenicity. Mol.
Carcinog. 3, 54-67.
Dunham, B., Liu, S., Taffet, S., Trabka-Janik, E., Delmar, M., Petryshyn,
R., Zheng, S. and Vallano, M. (1992). Immunolocalization and expression
of functional and nonfunctional cell-to-cell channels from wild-type and
mutant rat heart connexin43 cDNA. Circ. Res. 70, 1233-1243.
Fishman, G. I., Moreno, A. P., Spray, D. C. and Leinwand, L. A. (1991).
Functional analysis of human cardiac gap junction channel mutants. Proc.
Natl. Acad. Sci. USA 88, 3525-3529.
Fitzgerald, D. J. and Yamasaki, H. (1990). Tumor promotion: Models and
assay systems. Tetragen. Carcinog. Mutagen. 10, 89-102.
Goldberg, G. S., Moreno, A. P. and Lampe, P. D. (2002). Heterotypic and
homotypic gap junction channels mediate the selective transfer of
endogenous molecules between cells. J. Biol. Chem. 277, 36725-36730.
Hossain, M. Z., Jagdale, A. B., Ao, P. and Boynton, A. L. (1999). Mitogen-
activated protein kinase and phosphorylation of connexin43 are not
sufficient for the disruption of gap junctional communication by platelet-
derived growth factor and tetradecanoylphorbol acetate. J. Cell. Physiol.
179, 87-96.
Kadle, R., Zhang, J. T. and Nicholson, B. J. (1991). Tissue-specific
distribution of differentially phosphorylated forms of Cx43. Mol. Cell. Biol.
11, 363-369.
Kanemitsu, M. Y. and Lau, A. F. (1993). Epidermal growth factor stimulates
the disruption of gap junctional communication and connexin43
phosphorylation independent of 12-O-tetradecanoyl 13-acetate-sensitive
protein kinase C: The possible involvement of mitogen-activated protein
kinase. Mol. Biol. Cell 4, 837-848.
Kanemitsu, M. Y., Jiang, W. and Eckhart, W. (1998). Cdc2-mediated
phosphorylation of the gap junction protein, connexin43, during mitosis.
Cell Growth Differ. 9, 13-21.
Kwak, B. R., Hermans, M. M. P., de Jonge, H. R., Lohmann, S. M.,
Jongsma, H. J. and Chanson, M. (1995a). Differential regulation of
distinct types of gap junction channels by similar phosphorylating
conditions. Mol. Biol. Cell 6, 1707-1719.
Kwak, B. R., Saez, J. C., Wilders, R., Chanson, M., Fishman, G. I.,
Hertzberg, E. L., Spray, D. C. and Jongsma, H. J. (1995b). Effects of
cGMP-dependent phosphorylation on rat and human connexin43 gap
junction channels. Pflügers Arch 430, 770-778.
Kwak, B. R., van Veen, T. A. B., Analbers, L. J. S. and Jongsma, H. J.
(1995c). TPA increases conductance but decreases permeability in neonatal
rat cardiomyocyte gap junction channels. Exp. Cell Res. 220, 456-463.
Laird, D. L., 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., Puranam, K. L. and Revel, J. P. (1991). Turnover and
phosphorylation dynamics of connexin43 gap junction protein in cultured
cardiac myocytes. Biochem. J. 273, 67-72.
Laird, D. W., Fistouris, P., Batist, G., Alpert, L., Huynh, H. T., Carystinos,
G. D. and Alaoui-Jamali, M. A. (1999). Deficiency of connexin43 gap
junctions is an independent marker for breast tumors. Cancer Res. 59, 4104-
4110.
Lampe, P. D. (1994). Analyzing phorbol ester effects on gap junction
communication: A dramatic inhibition of assembly. J. Cell Biol. 127, 1895-
1905.
Lampe, P. D. and Lau, A. F. (2000). Regulation of gap junctions by
phosphorylation of connexins. Arch. Biochem. Biophys. 384, 205-215.
Lampe, P. D., Kurata, W. E., Warn-Cramer, B. and Lau, A. F. (1998a).
Formation of a distinct connexin43 phosphoisoform in mitotic cells is
dependent upon p34cdc2 kinase. J. Cell Sci. 111, 833-841.
Lampe, P. D., Nguyen, B. P., Gil, S., Usui, M., Olerud, J., Takada, Y.
and Carter, W. G. (1998b). Cellular interaction of integrin α3β1 with
laminin 5 promotes gap junctional communication. J. Cell Biol. 143, 1735-
1747.
Lampe, P. D., TenBroek, E. M., Burt, J. M., Kurata, W. E., Johnson, R.
G. and Lau, A. F. (2000). Phosphorylation of connexin43 on serine368 by
Journal of Cell Science 116 (11)
2211Phosphorylation at S368 of Cx43
protein kinase C regulates gap junctional communication. J. Cell Biol. 126,
1503-1512.
Moennikes, O., Buchmann, A., Ott, T., Willecke, K. and Schwarz, M.
(1999). The effect of connexin32 null mutation on hepatocarcinogenesis in
different mouse strains. Carcinogenesis 20, 1379-1382.
Moreno, A. P., Saez, J. C., Fishman, G. I. and Spray, D. C. (1994). Human
connexin43 gap junction channels. Regulation of unitary conductances by
phosphorylation. Circ. Res. 74, 1050-1057.
Munster, P. N. and Weingart, R. (1993). Effects of phorbol ester on gap
junctions of neonatal rat heart cells. Pflügers Arch. 423, 181-188.
Musil, L. S., Beyer, E. C. and Goodenough, D. A. (1990). Expression of the
gap junction protein connexin43 in embryonic chick lens: Molecular
cloning, ultrastructural localization, and post-translational phosphorylation.
J. Membr. Biol. 116, 163-175.
Paulson, A. F., Johnson, R. G. and Atkinson, M. M. (1994). Intercellular
communication is reduced by TPA and Ki-ras p21 in quiescent, but not
proliferating, NRK cells. Exp. Cell Res. 213, 64-70.
Reynhout, J. K., Lampe, P. D. and Johnson, R. G. (1992). An activator of
protein kinase C inhibits gap junction communication between cultured
bovine lens cells. Exp. Cell Res. 198, 337-342.
Rivedal, E. and Opsahl, H. (2001). Role of PKC and MAP kinase in EGF-
and TPA-induced connexin43 phosphorylation and inhibition of gap
junction intercellular communication in rat liver epithelial cells.
Carcinogenesis 22, 1543-1550.
Stein, L. S., Boonstra, J. and Burghardt, R. C. (1992). Reduced cell-cell
communication between mitotic and nonmitotic cells. Exp. Cell Res. 198,
1-7.
Stockert, R. J., Spray, D. C., Gao, Y., Suadicani, S. O., Ripley, C. R.,
Novikoff, P. M., Wolkoff, A. W. and Hertzberg, E. L. (1999). Deficient
assembly and function of gap junctions in Trf1, a trafficking mutant of the
human liver-derived cell line HuH-7. Hepatology 30, 740-747.
TenBroek, E. M., Lampe, P. D., Solan, J. L., Reynhout, J. K. and Johnson,
R. G. (2001). Ser364 of connexin43 and the upregulation of gap junction
assembly by cAMP. J. Cell Biol. 155, 1307-1318.
Trosko, J. E., Chang, C. C., Madhukar, B. V. and Klaunig, J. E. (1990).
Chemical, oncogene, and growth factor inhibition of gap junctional
intercellular communication: an integrative hypothesis of carcinogenesis.
Pathobiology 58, 265-278.
Vikhamar, G., Rivedal, E., Mollerup, S. and Sanner, T. (1998). Role of
Cx43 phosphorylation and MAP kinase activation in EGF induced
enhancement of cell communication in human kidney epithelial cells. Cell
Adhes. Commun. 5, 451-460.
Willecke, K., Eiberger, J., Degen, J., Eckardt, D., Romualdi, A.,
Guldenagel, M., Deutsch, U. and Sohl, G. (2002). Structural and
functional diversity of connexin genes in the mouse and human genome.
Biol. Chem. 383, 725-737.
Xie, H., Laird, D. W., Chang, T.-H. and Hu, V. W. (1997). A mitosis-specific
phosphorylation of the gap junction protein connexin43 in human vascular
cells: biochemical characterization and localization. J. Cell Biol. 137, 203-
210.
... De plus, l'assemblage des jonctions communicantes est également dépendant du cycle cellulaire. Il apparaît 50% plus efficace en phase G0 qu'en phase S (Solan, 2003). Enfin, la communication par les jonctions gap est également modulée au cours du cycle cellulaire. ...
... D'autre part, le traitement au palbociclib ne semble pas altérer les capacités agrégatives des cellules tumorales en condition ancrage-indépendant bien qu'il ait été montré comme responsable d'une diminution de la formation de métastases (Liu, 2017). Étant donnée l'impact de l'engagement des cellules au point R sur la mise en place des complexes jonctionnels telles que les jonctions adhérentes (Day, 1999), les jonctions communicantes (Solan, 2003) et les jonctions serrées (Tapia, 2009), l'arrêt des cellules en phase G1 pré-R induit par le traitement au palbociclib pourrait s'accompagner d'une augmentation de l'adhérence intercellulaire responsable de l'absence de modulation des capacités agrégatives des cellules traitées en comparaison avec les cellules non traitées observée dans ces travaux de thèse. Ces hypothèses pourraient faire l'objet de travaux supplémentaires qui permettraient de définir l'effet du traitement des cellules tumorales par les inhibiteurs de CDK4/6 sur la régulation des mécanismes impliqués dans formation de clusters de CTC par agrégation des cellules en condition ancrage-indépendant. ...
... On constate que les cellules MCF-7 traitées au palbociclib forment des clusters de manière plus efficace dans le temps en comparaison avec les cellules traitées au RO3306. Ces résultats appuient les résultats précédents et soulignent l'impact de l'engagement des cellules au point R sur leur capacité à former clusters.De plus, différents travaux mettant en lien l'arrêt des cellules en G0/G1 et la mise en place des jonctions intercellulaires, ont révélé que l'expression membranaire des jonctions serrées(Tapia, 2009) et de E-cadhérine(Day, 1999) ou encore l'assemblage des jonctions gap(Solan, 2003) étaient des évènements prédominants dans cette état spécifique du cycle cellulaire. Il a également été montré l'utilisation des inhibiteurs de CDK4/6, tels que le palbociclib, permettait d'augmenter l'expression de E-cadhérine (Rencuzogullari, 2019 ; Chen, 2020). ...
Thesis
La mortalité associée au cancer est principalement liée à la maladie métastatique. Le mécanisme de formation des métastases comporte plusieurs étapes depuis le détachement de cellules de la tumeur primaire, leur intravasion dans les systèmes circulatoires où elles survivent sous forme de cellules tumorales circulantes (CTCs), puis leur extravasion dans les organes distants où elles établissent des lésions malignes. Initialement décrit comme provenant de l'échappement de cellules individuelles depuis la masse tumorale primaire, depuis plusieurs années, le détachement et la migration de groupes de cellules ont également été observés et mis en lien avec la formation de métastases. On les retrouve au niveau de la circulation sanguine sous forme de clusters de CTC dont le potentiel métastatique a été révélé comme supérieur à celui de cellules tumorales circulantes isolées. De plus, leur présence chez les patients est associée à un pronostic péjoratif d'évolution de la maladie. Ainsi, la compréhension des mécanismes responsables de la formation de ces structures agrégatives de cellules tumorales présente un enjeu majeur dans le traitement de la progression du cancer. Le cycle cellulaire est un processus physiologique contrôlé par différents régulateurs tels que les kinases dépendantes des cyclines (CDK), et assure le bon déroulement du phénomène de division cellulaire d'une cellule mère en deux cellules filles identiques. Des altérations de sa régulation sont des caractéristiques associées à la prolifération incontrôlée des cellules tumorales. Dans ce contexte, l'objectif de mes travaux de thèse a été d'étudier les relations entre la capacité des cellules tumorales à former des clusters et la dynamique de progression dans le cycle cellulaire. Nous avons mis en œuvre différentes technologies semi-automatisées, précédemment développées au laboratoire, associant l'imagerie en temps réel et le traitement d'image. Ces approches nous ont permis de réaliser une analyse quantitative de la dynamique l'agrégation des cellules tumorales mammaires de la lignée MCF-7 in vitro en condition ancrage-indépendant. Dans une première partie de nos travaux, nous avons cherché à déterminer le rôle l'agrégation des cellules tumorales sur la progression des cellules dans le cycle cellulaire. A l'échelle de la population cellulaire, nous n'avons pas mis en évidence de modification de la répartition des cellules dans le cycle cellulaire au cours de ce processus. À l'échelle cellulaire, via l'utilisation d'une lignée de cellules MCF-7 exprimant un rapporteur fluorescent de l'activité de la kinase CDK2, nous avons montré que les cellules ayant franchi le point de restriction (R) formaient des clusters de manière moins efficace, suggérant que la position des cellules dans le cycle modulait leur capacité à agréger. Dans une seconde partie, j'ai étudié l'impact de l'altération de la progression des cellules tumorales mammaires dans le cycle cellulaire en réponse à des traitements pharmacologiques sur leur capacité à former des clusters. J'ai ainsi montré que des cellules synchronisées au point de contrôle intra-mitotique, présentaient une altération de leur capacité à agréger, altération qui semble être associée à une moindre capacité à former des extensions membranaires. De manière similaire, le traitement des cellules MCF-7 avec du taxol ou avec de la vinorelbine, deux poisons des microtubules utilisés en thérapie anti-tumorale, altère leur capacité à agréger. Dans une moindre mesure, des résultats préliminaires ont également montré qu'un traitement avec les inhibiteurs des kinases dépendantes des cyclines CDK4/6 altérait la capacité à former des clusters. Ensemble, ces résultats apportent des éléments nouveaux et complémentaires aux études préalables réalisées sur les régulateurs de la formation de clusters impliqués dans la formation de métastases et pourraient permettre de mieux comprendre les effets, parfois adverses, de certains traitements.
... In contrast to the S368 site of Cx43, cyclic adenosine monophosphate-and protein kinase A-dependent phosphorylation of the S365 site of Cx43 enhances the expression of GJs and opens the channel gate. Studies have also shown that phosphorylated-S365 may have a role because it prevents the phosphorylation of Cx43 at S368 [13,14]. In addition, phosphorylation at S368 may have a role in key events in the cell cycle. ...
... In addition, phosphorylation at S368 may have a role in key events in the cell cycle. Studies have shown that the phosphorylation level of S368 in the S phase and G2/M phase of the cell cycle increases significantly, whereas the number of functional connection channels and intercellular communication decrease, that is, the phosphorylation level at the S368 site is negatively correlated with GJ assembly [13]. ...
Article
Full-text available
Gap junctions are the main form of interaction between cardiomyocytes, through which the electrochemical activities between cardiomyocytes can be synchronized to maintain the normal function of the heart. Connexins are the basis of gap junctions. Changes in the expression, structural changes (e.g., phosphorylation and dephosphorylation), and distribution of connexins can affect the normal electrophysiological activities of the heart. Myocardial infarction (MI) and concurrent arrhythmia, shock, or heart failure can endanger life. The structural and functional damage of connexin (Cx) 43 in cardiomyocytes is a central part of the pathological progression of MI and is one of the main pathological mechanisms of arrhythmia after MI. Therefore, increasing Cx43 expression has become one of the main measures to prevent MI. Also, intervention in Cx43 expression can improve the structural and electrical remodeling of the myocardium to improve MI prognosis. Here, research progress of Cx43 in MI and its prevention and treatment using Traditional Chinese Medicine formulations is reviewed.
... However, in some cell types, TPA treatment led to an increase in Cx43 site-specific phosphorylation, i.e. changed Cx43 phosphorylation profile, but without a shift in Cx43 electrophoresis mobility (Kubincova et al., 2019;Rivedal and Opsahl, 2001;Solan et al., 2003). Specifically, for example, the increased phosphorylation on Ser368 did not result in a detectable shift in the migration of Cx43 bands (Solan et al., 2003). ...
... However, in some cell types, TPA treatment led to an increase in Cx43 site-specific phosphorylation, i.e. changed Cx43 phosphorylation profile, but without a shift in Cx43 electrophoresis mobility (Kubincova et al., 2019;Rivedal and Opsahl, 2001;Solan et al., 2003). Specifically, for example, the increased phosphorylation on Ser368 did not result in a detectable shift in the migration of Cx43 bands (Solan et al., 2003). Therefore, EDCs and TPA could dysregulate testicular GJIC in Leydig TM3 cells via site-specific phosphorylation(s) of Cx43 protein even without a shift in Cx43 electrophoresis mobility. ...
Article
A decline in male fertility possibly caused by environmental contaminants, namely endocrine-disrupting chemicals (EDCs), is a topic of public concern and scientific interest. This study addresses a specific role of testicular gap junctional intercellular communication (GJIC) between adjacent prepubertal Leydig cells in endocrine disruption and male reproductive toxicity. Organochlorine pesticides (lindane, methoxychlor, DDT), industrial chemicals (PCB153, bisphenol A, nonylphenol and octylphenol) as well as personal care product components (triclosan, triclocarban) rapidly dysregulated GJIC in murine Leydig TM3 cells. The selected GJIC-inhibiting EDCs (methoxychlor, triclosan, triclocarban, lindane, DDT) caused the immediate GJIC disruption by the relocation of gap junctional protein connexin 43 (Cx43) from the plasma membrane and the alternation of Cx43 phosphorylation pattern (Ser368, Ser279, Ser282) of its full-length and two N-truncated isoforms. After more prolonged exposure (24 h), EDCs decreased steady-state levels of full-length Cx43 protein and its two N-truncated isoforms, and eventually (triclosan, triclocarban) also tight junction protein ZO-1. The disturbance of GJIC was accompanied by altered activity of mitogen-activated protein kinases MAPK-Erk1/2 and MAPK-p38, and a decrease in stimulated progesterone production. Our results indicate that EDCs might disrupt testicular homeostasis and development via disruption of testicular GJIC, a dysregulation of junctional and non-junctional functions of Cx43, activation of MAPKs, and disruption of an early stage of steroidogenesis in prepubertal Leydig cells. These critical disturbances of Leydig cell development and functions during a prepubertal period might be contributing to impaired male reproduction health later on.
... Existen muchos tipos de conexinas, pero una de las más abundantes, ubicuas y estudiadas es conexina 43, presente también en el ovario (27) . Conexina 43 no solo es una proteína que permite el anclaje célula-célula, sino que también es capaz de regular otros importantes aspectos del funcionamiento de la célula, tales como sobrevida, apoptosis y ciclo celular (28)(29)(30) , contribuyendo a la regulación del tamaño de las masas celulares en los tejidos. El control del funcionamiento de esta proteína se realiza mediante modificaciones químicas, tales como fosforilación y desfosforilación de algunos aminoácidos de su estructura proteica (28) . ...
Article
Ovarian cancer is one of the most aggressive and poor prognosis cancer, which appears predominantly as Epithelial Ovarian Cancer (EOC). Many studies have been conducted to establish connections between neurotrophin receptors and the development, progression and response of cancer therapy. The tyrosine kinases receptors (TRK) have been considered as important target in several cancers, including EOC. Metastasic process and resistance to cancer therapies have been associated with the TRK neurotrophin receptor B (TRKB), whose main ligand is brain derivated neurotrophic ligand (BDNF). Another important aspect in tumor development is the expression of adhesion molecules such as connexin 43. This protein is present in many tissues included the ovary and cancer cells. When TRK receptors are activated by theirs ligands, connexin 43 is phosphorilated and promotes several processes in tissues, like remodeling gap junctions and cellular permeability. All these features are very important in tumor cells, and probably would be involved in the process of metastasis, tumor growth and arrival of nutrients and therapy drugs to the tumor mass. The aim of this revew is to expose current knowledge about connexin 43 and TRKB receptor in ovarian cancer.
... Phosphorylation of Cx43 at Ser368 is important for the cell cycle. Phosphorylation at this site is generally elevated in the S and G2/M phases of mitosis [102]. MAPK mainly modifies the Ser255, Ser262, Ser279, and Ser282 sites of Cx43, and GJIC is inhibited after phosphorylation at these sites [81]. ...
Article
Full-text available
Gap junctions (GJs), which are composed of connexins (Cxs), provide channels for direct information exchange between cells. Cx expression has a strong spatial specificity; however, its influence on cell behavior and information exchange between cells cannot be ignored. A variety of factors in organisms can modulate Cxs and subsequently trigger a series of responses that have important effects on cellular behavior. The expression and function of Cxs and the number and function of GJs are in dynamic change. Cxs have been characterized as tumor suppressors in the past, but recent studies have highlighted the critical roles of Cxs and GJs in cancer pathogenesis. The complex mechanism underlying Cx and GJ involvement in cancer development is a major obstacle to the evolution of therapy targeting Cxs. In this paper, we review the post-translational modifications of Cxs, the interactions of Cxs with several chaperone proteins, and the effects of Cxs and GJs on cancer.
... Biopsy samples from heterozygous p.R451G patients displayed increased fibro-fatty infiltration, decreased DSP expression at the ID, and mislocalization of Cx43 from the ID [11]. Additionally, iPSC-derived cardiomyocytes identified a reduced expression of Cx43 and increased pCx43-S368, a marker associated with reduced channel opening, altered stability, and has been associated with potential downstream protein degradation via the ubiquitin proteolytic system [11,[36][37][38][39]. Decreased expression of DSP that was not connected to a loss of DSP mRNA was also identified. ...
Article
Full-text available
Arrhythmogenic cardiomyopathy (ACM) is an inherited disorder characterized by fibro-fatty infiltration with an increased propensity for ventricular arrhythmias and sudden death. Genetic variants in desmosomal genes are associated with ACM. Incomplete penetrance is a common feature in ACM families, complicating the understanding of how external stressors contribute towards disease development. To analyze the dual role of genetics and external stressors on ACM progression, we developed one of the first mouse models of ACM that recapitulates a human variant by introducing the murine equivalent of the human R451G variant into endogenous desmoplakin (DspR451G/+). Mice homozygous for this variant displayed embryonic lethality. While DspR451G/+ mice were viable with reduced expression of DSP, no presentable arrhythmogenic or structural phenotypes were identified at baseline. However, increased afterload resulted in reduced cardiac performance, increased chamber dilation, and accelerated progression to heart failure. In addition, following catecholaminergic challenge, DspR451G/+ mice displayed frequent and prolonged arrhythmic events. Finally, aberrant localization of connexin-43 was noted in the DspR451G/+ mice at baseline, becoming more apparent following cardiac stress via pressure overload. In summary, cardiovascular stress is a key trigger for unmasking both electrical and structural phenotypes in one of the first humanized ACM mouse models.
... PKC activation was found to phosphorylate Ser 368 and induce closure and internalisation of gap junctions 298 . However, dephosphorylation of the same residue was also found to decrease gap junctional activity 299 "truncated" isoforms originated from AUG starting codons located into the Cx43 gene. ...
Thesis
The Atypical Chemokine Receptor 3 (ACKR3) and CXCR4 are two G protein-coupled receptors (GPCR) belonging to the CXC chemokine receptor family. Both receptors are activated upon CXCL12 binding and are over-expressed in various tumours, including glioma, where they have been found to promote proliferation and invasive behaviours. Upon CXCL12 binding, CXCR4 activates canonical GPCR signalling pathways involving Gαi protein and β-arrestins. In addition, CXCR4 was found to interact with several proteins able to modify its signalling, trafficking and localization. In contrast, the cellular pathways underlying ACKR3-dependent effects remain poorly characterized. Several reports show that ACKR3 engages β-arrestin-dependent signalling pathways, but its coupling to G proteins is restricted to either specific cellular populations, including astrocytes, or occurs indirectly via its interaction with CXCR4. ACKR3 also associates with the epidermal growth factor receptor to promote proliferation of tumour cells in an agonist-independent manner. These examples suggest that the extensive characterization of ACKR3 and CXCR4 interactomes might be a key step in understanding or clarifying their roles in physiological and pathological contexts. This thesis addressed this issue employing an affinity purification coupled to high-resolution mass spectrometry proteomic strategy that identified 19 and 151 potential protein partners of CXCR4 and ACKR3 transiently expressed in HEK-293T cells, respectively. Amongst ACKR3 interacting proteins identified, we paid particular attention on the gap junction protein Connexin-43 (Cx43), in line with its overlapping roles with the receptor in the control of leukocyte entry into the brain, interneuron migration and glioma progression. Western blotting and BRET confirmed the specific association of Cx43 with ACKR3 compared to CXCR4. Likewise, Cx43 is co-localized with ACKR3 but not CXCR4 in glioma initiating cell lines, and ACKR3 and Cx43 are co-expressed in astrocytes of the sub-ventricular zone and surrounding blood vessels in adult mouse brain, suggesting that both proteins form a complex in authentic cell or tissue contexts. Further functional studies showed that ACKR3 influences Cx43 trafficking and functionality at multiple levels. Transient expression of ACKR3 in HEK-293T cells to mimic ACKR3 overexpression detected in several cancer types, induces Gap Junctional Intercellular Communication (GJIC) inhibition in an agonist-independent manner. In addition, agonist stimulation of endogenously expressed ACKR3 in primary cultured astrocytes inhibits Cx43-mediated GJIC through a mechanism that requires activation of Gαi protein, and dynamin- and β-arrestin2-dependent Cx43 internalisation. Therefore, this thesis work provides the first functional link between the CXCL11/CXCL12/ACKR3 axis and gap junctions that might underlie their critical role in glioma progression.
Article
Background: Scn5a heterozygous null (Scn5a+/-) mice have historically been used to investigate arrhythmogenic mechanisms of diseases such as Brugada Syndrome and Lev's disease. Previously, we demonstrated that reducing ephaptic coupling (EpC) in ex vivo hearts exacerbates pharmacological Nav1.5 loss of function (LOF). Whether this effect is consistent in a genetic Nav1.5 LOF model is yet to be determined. We hypothesized that loss of EpC would result in greater reduction in conduction velocity (CV) for the Scn5a+/- mouse relative to wild type (WT). Methods: In vivo ECGs and ex vivo optical maps were recorded from Langendorff-perfused Scn5a+/- and WT mouse hearts. EpC was reduced with perfusion of a hyponatremic solution, the clinically relevant osmotic agent mannitol, or a combination of the two. Results: Neither in vivo QRS duration nor ex vivo CV during normonatremia was significantly different between the two genotypes. In agreement with our hypothesis, we found that hyponatremia severely slowed CV and disrupted conduction for 4/5 Scn5a+/- mice, but 0/6 WT mice. Additionally, treatment with mannitol slowed CV to a greater extent in Scn5a+/- relative to WT hearts. Unexpectedly, treatment with mannitol during hyponatremia did not further slow CV in either genotype, but resolved the disrupted conduction observed in Scn5a+/- hearts. Similar results in guinea pig hearts suggest the effects of mannitol and hyponatremia are not species specific. Conclusion: Loss of EpC through either hyponatremia or mannitol alone results in slowed or disrupted conduction in a genetic model of Nav1.5 LOF. However, combination of these interventions attenuates conduction slowing.
Article
Background: Cardiac conduction is understood to occur through gap junctions. Recent evidence supports ephaptic coupling as another mechanism of electrical communication in the heart. Conduction via gap junctions predicts a direct relationship between conduction velocity (CV) and bulk extracellular resistance. By contrast, ephaptic theory is premised on the existence of a biphasic relationship between CV and the volume of specialized extracellular clefts within intercalated discs such as the perinexus. Our objective was to determine the relationship between ventricular CV and structural changes to micro- and nanoscale extracellular spaces. Methods: Conduction and Cx43 (connexin43) protein expression were quantified from optically mapped guinea pig whole-heart preparations perfused with the osmotic agents albumin, mannitol, dextran 70 kDa, or dextran 2 MDa. Peak sodium current was quantified in isolated guinea pig ventricular myocytes. Extracellular resistance was quantified by impedance spectroscopy. Intercellular communication was assessed in a heterologous expression system with fluorescence recovery after photobleaching. Perinexal width was quantified from transmission electron micrographs. Results: CV primarily in the transverse direction of propagation was significantly reduced by mannitol and increased by albumin and both dextrans. The combination of albumin and dextran 70 kDa decreased CV relative to albumin alone. Extracellular resistance was reduced by mannitol, unchanged by albumin, and increased by both dextrans. Cx43 expression and conductance and peak sodium currents were not significantly altered by the osmotic agents. In response to osmotic agents, perinexal width, in order of narrowest to widest, was albumin with dextran, 70 kDa; albumin or dextran, 2 MDa; dextran, 70 kDa or no osmotic agent, and mannitol. When compared in the same order, CV was biphasically related to perinexal width. Conclusions: Cardiac conduction does not correlate with extracellular resistance but is biphasically related to perinexal separation, providing evidence that the relationship between CV and extracellular volume is determined by ephaptic mechanisms under conditions of normal gap junctional coupling.
Thesis
Gap junctions in the normal and regenerating inner ear of chick hatchlings have been immunohistochemically and functionally examined. Additionally, the effect of blocked gap junction channels on the proliferation of supporting cells in response to hair cell death has been assessed. Studies have been conducted primarily in organotypic cultures of the auditory and vestibular sensory epithelium of chicken hatchlings. For regeneration studies, explants of basilar papilla and utricle were exposed to gentamicin, an ototoxic antibiotic, to induce hair cell loss. Following hair cell death, proliferation of supporting cells was upregulated and after 5 days immature, new hair cells were apparent. Blocking of gap junctions by carbenoxolone led to a significant reduction in the proliferation of supporting cells, suggesting an involvement of intercellular communication in the regeneration of hair cells. Expression of connexin 43 (Cx43) and the chicken-specific connexin 31 (cCx31) was examined by immunohistochemistry and confocal microscopy. cCx31 was strongly expressed in the normal basilar papilla and utricle. Cx43 was confined to the supporting cells of the auditory sensory epithelium, where its immunolabelling co-localised with cCx31. In response to hair cell loss, Cx43 was transiently downregulated. This finding, together with the absence of Cx43 in the sensory epithelium of the utricle, which has a constant turnover of hair cells, might point to an inhibitory effect of Cx43 on supporting cell proliferation. A dye-coupling assay, based on fluorescence recovery after photobleaching (FRAP), has been developed to examine the diffusion of the fluorescent tracer, calcein, between supporting cells in the intact tissue. Recovery of fluorescence occurred in supporting cells, but not in hair cells and was inhibited by the presence of carbenoxolone. Most notably, an asymmetric dye transfer across the basilar papilla was observed. The absence of directional permeability in the utricular macula, and in the drug-damaged basilar papilla, strongly suggests that the co-expression of cCx31 and Cx43 results in chemically rectifying gap junctions.
Article
Full-text available
Wounding of skin activates epidermal cell migration over exposed dermal collagen and fibronectin and over laminin 5 secreted into the provisional basement membrane. Gap junctional intercellular communication (GJIC) has been proposed to integrate the individual motile cells into a synchronized colony. We found that outgrowths of human keratinocytes in wounds or epibole cultures display parallel changes in the expression of laminin 5, integrin α3β1, E-cadherin, and the gap junctional protein connexin 43. Adhesion of keratinocytes on laminin 5, collagen, and fibronectin was found to differentially regulate GJIC. When keratinocytes were adhered on laminin 5, both structural (assembly of connexin 43 in gap junctions) and functional (dye transfer) assays showed a two- to threefold increase compared with collagen and five- to eightfold over fibronectin. Based on studies with immobilized integrin antibody and integrin-transfected Chinese hamster ovary cells, the interaction of integrin α3β1 with laminin 5 was sufficient to promote GJIC. Mapping of intermediate steps in the pathway linking α3β1–laminin 5 interactions to GJIC indicated that protein trafficking and Rho signaling were both required. We suggest that adhesion of epithelial cells to laminin 5 in the basement membrane via α3β1 promotes GJIC that integrates individual cells into synchronized epiboles.
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
The transformed or normal phenotype of cultured normal rat kidney cells infected with a temperature-sensitive mutant of avian sarcoma virus is conditional on the temperature at which the cells are grown. Using dye injection techniques, we show that junction-mediated dye transfer is also temperature-sensitive. The extent and rate of transfer between infected cells grown at the transformation-permissive temperature (35 degrees C) is significantly reduced when compared to infected cells grown at the nonpermissive temperature (40.5 degrees C) or uninfected cells grown at either temperature. Infected cells subjected to reciprocal temperature shifts express rapid and reversible alterations of dye transfer capacities, with responses evident by 15 min and completed by 60 min for temperature shifts in either direction. These results suggest that altered junctional capacities may be fundamental to the expression of the ASV-induced, transformed phenotype.
Article
Connexin32 (Cx32) is the major gap junctional protein in mouse liver. We have shown recently that the formation of liver tumours in Cx32-deficient mice is strongly increased in comparison with control wild-type mice, demonstrating that the deficiency in gap junctional communication has an enhancing effect on hepatocarcinogenesis. We have now compared the effect of Cx32 deficiency on liver carcinogenesis in two strains of mice with differing susceptibility to hepatocarcinogenesis. Heterozygous Cx32 +/- females were crossed with male Cx32 wild-type C57BL/6J (low susceptibility) or C3H/He (high susceptibility) mice. Since the Cx32 gene is located on the X-chromosome, the resulting F1 males segregated to the genotypes Cx32 Y/+ and Cx32 Y/- . Genotyping was performed by PCR-analysis using tail-tip DNA. Weanling male mice were i.p. injected with a single dose of N-nitrosodiethylamine and were killed 16, 21 or 26 weeks later. The number, volume fraction and size distribution of precancerous liver lesions characterized by a deficiency in the marker enzyme glucose-6-phosphatase were quantitated. The results demonstrate that Cx32 deficiency only slightly affects the number of enzyme-altered lesions, but strongly enhances their growth, both in the resistant and the susceptible mouse strain, suggesting that decreased intercellular communication results in tumour promoting activity irrespective of the genetic background of the mouse strain used. Since Cx32-deficient C3H/ He hybrids were -5-10 times more sensitive than C3H/He hybrids with an intact Cx32 gene, this mouse strain may prove very useful for toxicological screening purposes.
Article
The effects of mitosis on gap junctional intercellular communication (GJIC) were quantified in a clonal cell line of spontaneously immortalized rat granulosa cells (SIGC) using a fluorescence recovery after photobleaching assay. Reduction of GJIC was associated with the process of mitosis and was first apparent at the onset of prophase. Resumption of GJIC between newly divided cells and surrounding cells occurred slowly, requiring several hours following cytokinesis before reestablishment of maximal rates. Mitotic rates of GJIC in SIGC were comparable to values obtained in interphase cells partially uncoupled by 0.5 m M octanol. Limited studies of other cell lines generalized the mitotic-associated reduction of GJIC observed in SIGC. The data suggest that mitosis is one process which alters GJIC. This could be of significance when there is a change in the rate of proliferation, such as in the acquisition of immortalization, an early stage of transformation.
Article
The role of the v-Ha-ras oncogene in tumorigenesis in an in vitro/in vivo model system was studied by investigating the expression of the Ha-ras gene, gap junctional intercellular communication, and tumorigenicity as endpoints. Infection of a Fischer 344 rat liver epithelial cell line (WB 344) with a retrovirus containing the v-Ha-ras oncogene resulted in altered cell morphology and decreased contact sensitivity. Gap junctional intercellular communication in v-Ha-ras infected WB cells (WBHa-ras), assessed by fluorescence redistribution after photobleaching (FRAP), microinjection/dye transfer, and scrape-loading/dye transfer techniques, was markedly decreased compared with the level in control WB cells. Injection of 107 WBHa-ras cells into the portal vein of male F344 rats caused multiple focal hepatic lesions within 1 and 2 wk, merging to large invading tumors after 3 and 4 wk. Examination of the methylation pattern of the Ha-ras gene in WBHa-ras and control WB cells showed that the infected Ha-ras gene was relatively hypomethylated in comparison to the normal cellular Ha-ras gene, indicating a greater potential for expression. There was an increased level of Ha-ras mRNA in hepatomas as compared with both adjacent nontumor liver tissue and liver tissue obtained from normal animals. Three cell lines derived from three different primary hepatic tumors induced by an injection of WBHa-ras cells in a F344 rat displayed similar growth characteristics, levels of gap junctional communication, and methylation patterns as the original WBHa-ras cells. The results of these studies have established a strong positive correlation between expression of the Ha-ras oncogene, reduced gap junctional intercellular communication, decreased contact sensitivity, and tumorigenicity of the v-Ha-ras-infected rat liver epithelial cells.
Article
Myocytes were isolated from the ventricles of neonatal rat hearts and cultured for 1–3 days. Newly formed cell pairs were used to examine the conductance of gap junctions, g j. Measurements were performed using a dual voltage-clamp method in conjunction with a whole-cell, tight-seal recording. Exposure to the phorbol ester 12-O-tetradecanoylphorbol-13-acetate (TPA, 100–160 nM) led to a decrease in g j. Single-channel events recorded immediately before complete uncoupling yielded a single-channel conductance, j, of 40.5 pS, implying that TPA affects the channel kinetics rather than j. TPA-induced uncoupling was observed at subphysiological levels of cytosolic Ca2+ (pipette solution=18 nM), not at physiological levels (pipette solution=170 nM). The effects of TPA could not be mimicked by 250 M 1-oleoyl-2-acetyl-glycerol (OAG). Preincubation with TPA (up to 24 h) revealed no changes in g j attributable to down-regulation of protein kinase C, PKC. Pretreatment with PKC inhibitors, staurosporine or PKCI, prevented the TPA-dependent decrease in g j. TPA-dependent uncoupling was not impaired by 4-bromophenacyl bromide, an inhibitor of phospholipase A2, PLA2; conversely, an arachidonic acid-dependent decrease in g j was not prevented by PKCI. This suggests that g j regulation does not involve an interaction between PLA2 and PKC.
Article
Currently little is known about the regulation of gap junction communication in the lens. We report here on the effects of the protein kinase C activator, 12-O-tetradecanoylphorbol-13-acetate (TPA), on cultured bovine lens cells which appeared to be epithelial in nature. Dramatically reduced intercellular transfer of the fluorescent dye Lucifer yellow was observed when the cultured lens cells were treated with octanol, a known inhibitor of gap junction communication. TPA (4 beta isomer) was also shown to reduce intercellular permeability within these cultures. In contrast, an inactive form of TPA, 4 alpha-TPA, did not decrease dye transfer. Permeability was evaluated in terms of both the number of cells receiving dye and the rate of decrease in fluorescence intensity in the injected cell. The maximum decreases in dye transfer occurred at 2 h of TPA treatment and dye transfer gradually increased to control levels over a time course of many hours. Incubation of cultures with 32Pi and immunoprecipitation using antibodies to the N- and C-terminal regions of connexin43 demonstrated a gap junction phosphoprotein of 43,000 Da. Phosphorylation of connexin43 increased during the first 2 h of TPA treatment. These results suggest that protein kinase C has a direct or indirect effect on gap junction communication in cultured lens cells.
Article
The carboxyl terminal cytoplasmic domain of distinct gap junction proteins may play an important role in assembly of functional channels as well as differential responsiveness to pH, voltage, and intracellular second messengers. Oligonucleotide-directed site-specific mutagenesis in a paired Xenopus laevis oocyte expression system was used to examine the expression of mRNAs encoding wild-type and carboxyl terminal mutant connexin43 (Cx43) proteins. Oocytes were stripped, injected with mRNA or distilled water (dH2O), preincubated for 16-20 hours, and then paired for 5-10 hours; this process was followed by electrophysiological recording using the dual voltage-clamp technique. Initial experiments compared the relative junctional conductances (Gjs) in oocyte pairs expressing Cx43 (382 amino acid residues) and two truncated mutants lacking most or a portion of the cytoplasmic carboxyl terminal. The shortest mutant (M241) contained 240 amino acid residues and was devoid of all phosphorylatable serine residues in the cytoplasmic tail; its length approximated the length of liver connexin26. The longest mutant (M257) tested contained 256 amino acid residues, including two serine residues. Oocyte pairs expressing M241 yielded a Gj similar to that of oocytes injected with dH2O, whereas M257 yielded a Gj similar to that of oocytes injected with Cx43. Immunoprecipitation studies showed that Cx43, M257, and M241 proteins were readily detectable in oocytes injected with their respective mRNAs, indicating that the lack of Gj observed with the M241 mRNA was not due to reduced translation. Immunocytochemical studies revealed that wild-type and both truncated mutants were localized to the area of cell-to-cell contact between the paired oocytes, indicating that protein targeting to the membrane was not inhibited in oocytes injected with M241 mRNA. Oocyte pairs expressing mutants in which serine residues were replaced with nonphosphorylatable amino acids (serine codon No. 255 AGC was converted to GCC, alanine, designated as M255S----A, and serine codon No. 244 AGC was converted to GGC, glycine, designated as M244S----G) showed Gjs similar to M257, indicating that these serine residues and, by inference, their phosphorylation state are not critical for expression of functional channels. The importance of the length of the carboxyl terminus was assessed by comparing the Gjs in a series of mutants that were intermediate in length between M257 and M241. Gradual shortening of the carboxyl terminus produced a gradual reduction of Gj relative to M257. However, simple deletion of amino acid residues 241-257 from the wild-type Cx43 did not affect Gj relative to M257.(ABSTRACT TRUNCATED AT 400 WORDS)
Article
Prior to confluence, cultures of Madin Darby canine kidney (MDCK) cells expressed gap junctional communication, as assessed by fluorescent dye transfer, as well as relatively high levels of an anti-connexin43 immunoreactive component referred to as connexin43 (Cx43). After confluence, dye coupling and levels of Cx43 were dramatically reduced. Immunofluorescence analysis of the distribution of Cx43 in subconfluent cultures showed punctate labeling on the plasma membrane at regions of cell apposition and a more diffuse labeling in perinuclear regions. Western blots of total cell homogenates showed that the dephosphorylated form of Cx43 was more abundant than the phosphorylated forms. Phosphorylation of Cx43 was not significantly affected by 8-Bromo-cAMP or 8-Bromo-cGMP. However, 12-O-tetradecanoylphorbol-13-acetate (TPA) inhibited dye coupling and induced an increase in the amount of phosphorylated forms of Cx43 at the expense of the dephosphorylated form. This effect occurred as rapidly as 5 min after TPA treatment without apparent changes in distribution of Cx43 or cell morphology. These results suggest that second messenger pathways involving protein kinase C, but not cAMP- or cGMP-dependent protein kinase, led to changes in electrophoretic mobility of Cx43, revealed by Western blot, consistent with an alteration in the state of phosphorylation of the gap junction protein. Treatments with staurosporine, a protein kinase inhibitor, or okadaic acid, a protein phosphatase inhibitor, either alone or in combination with TPA, indicated that the abundance of the dephosphorylated form of Cx43 in MDCK cells was due to low kinase activity. It was also found that lowering the concentration of extracellular Ca2+, which reduced cell contact, did not affect the abundance, the state of phosphorylation, or the TPA-induced phosphorylation of Cx43. These results suggest that neither extracellular Ca2+ nor cell contact is required for basal or TPA-induced phosphorylation of Cx43.