ArticlePDF Available

Pannexin1 and Pannexin3 Delivery, Cell Surface Dynamics, and Cytoskeletal Interactions

Authors:

Abstract and Figures

Pannexins (Panx) are a class of integral membrane proteins that have been proposed to exhibit characteristics similar to those of connexin family members. In this study, we utilized Cx43-positive BICR-M1R(k) cells to stably express Panx1, Panx3, or Panx1-green fluorescent protein (GFP) to assess their trafficking, cell surface dynamics, and interplay with the cytoskeletal network. Expression of a Sar1 dominant negative mutant revealed that endoplasmic reticulum to Golgi transport of Panx1 and Panx3 was mediated via COPII-dependent vesicles. Distinct from Cx43-GFP, fluorescence recovery after photobleaching studies revealed that both Panx1-GFP and Panx3-GFP remained highly mobile at the cell surface. Unlike Cx43, Panx1-GFP exhibited no detectable interrelationship with microtubules. Conversely, cytochalasin B-induced disruption of microfilaments caused a severe loss of cell surface Panx1-GFP, a reduction in the recoverable fraction of Panx1-GFP that remained at the cell surface, and a decrease in Panx1-GFP vesicular transport. Furthermore, co-immunoprecipitation and co-sedimentation assays revealed actin as a novel binding partner of Panx1. Collectively, we conclude that although Panx1 and Panx3 share a common endoplasmic reticulum to Golgi secretory pathway to Cx43, their ultimate cell surface residency appears to be independent of cell contacts and the need for intact microtubules. Importantly, Panx1 has an interaction with actin microfilaments that regulates its cell surface localization and mobility.
Content may be subject to copyright.
Pannexin1 and Pannexin3 Delivery, Cell Surface Dynamics,
and Cytoskeletal Interactions
*
S
Received for publication, November 4, 2009, and in revised form, January 6, 2010 Published, JBC Papers in Press, January 19, 2010, DOI 10.1074/jbc.M109.082008
Ruchi Bhalla-Gehi, Silvia Penuela, Jared M. Churko, Qing Shao, and Dale W. Laird
1
From the Department of Anatomy and Cell Biology, University of Western Ontario, London, Ontario N6A 5C1, Canada
Pannexins (Panx) are a class of integral membrane proteins
that have been proposed to exhibit characteristics similar to
those of connexin family members. In this study, we utilized
Cx43-positive BICR-M1R
k
cells to stably express Panx1, Panx3,
or Panx1-green fluorescent protein (GFP) to assess their traf-
ficking, cell surface dynamics, and interplay with the cytoskel-
etal network. Expression of a Sar1 dominant negative mutant
revealed that endoplasmic reticulum to Golgi transport of
Panx1 and Panx3 was mediated via COPII-dependent vesicles.
Distinct from Cx43-GFP, fluorescence recovery after photo-
bleaching studies revealed that both Panx1-GFP and Panx3-
GFP remained highly mobile at the cell surface. Unlike Cx43,
Panx1-GFP exhibited no detectable interrelationship with
microtubules. Conversely, cytochalasin B-induced disruption of
microfilaments caused a severe loss of cell surface Panx1-GFP, a
reduction in the recoverable fraction of Panx1-GFP that
remained at the cell surface, and a decrease in Panx1-GFP vesic-
ular transport. Furthermore, co-immunoprecipitation and co-
sedimentation assays revealed actin as a novel binding partner
of Panx1. Collectively, we conclude that although Panx1 and
Panx3 share a common endoplasmic reticulum to Golgi secre-
tory pathway to Cx43, their ultimate cell surface residency
appears to be independent of cell contacts and the need for
intact microtubules. Importantly, Panx1 has an interaction with
actin microfilaments that regulates its cell surface localization
and mobility.
The pannexin family is a new class of integral membrane
glycoproteins that have been identified to share sequence
homology with the invertebrate gap junction proteins, innexins
(1). Unlike the connexin family that is composed of 21 members
(2), both the human and mouse genomes contain only three
pannexin-encoding genes (Panx1, Panx2, and Panx3) (1).
Although connexins and pannexins exhibit no sequence ho-
mology, pannexins are predicted to share similar topology with
connexins, which includes four transmembrane domains, two
extracellular loops, a cytoplasmic loop, and intracellular amino
and carboxyl termini (3–5). Our previous study has shown that
ectopically expressed Panx1 and Panx3 are capable of traffick-
ing to normal rat kidney (NRK)
2
cell surfaces; however, their
distribution profile at cell-cell interfaces is not typically clus-
tered or punctate as seen for Cx43 (5). Consistently, electron
micrographs of Panx1 overexpressing Madin-Darby canine
kidney cells also revealed dispersed Panx1 localization at the
plasma membrane with no evidence of gap junction plaques (4).
Cx43 has a relatively short half-life of 1–3 h (6). As a result,
Cx43 subunits assembled into connexons within the trans-
Golgi network (7) are constantly being transported to and re-
moved from the plasma membrane (8). Recent reports
provide evidence that although Panx1 appears to form similar
hexameric channel units defined as pannexons (4), they exhibit
slower turnover dynamics in comparison with Cx43, as as-
sessed by the use of pharmacological blockers of protein syn-
thesis and protein trafficking (4, 5). Functionally, most studies
support the premise that pannexins, in particular Panx1, form
single membrane channels (4, 5, 9–12), whereas far less evi-
dence supports the notion that they also can form intercellular
channels (13, 14).
The delivery of Cx43 for the assembly of functional channels
at the cell surface has been extensively studied using fluores-
cent protein and epitope tags (15, 16). Cx43-GFP has been
shown to exhibit similar distribution and functional charac-
teristics as its untagged counterpart, when assessed by dye
permeability and electrical conductance (8, 17). Likewise, GFP
tagging of the carboxyl-terminal tail of Panx1 did not signifi-
cantly alter the localization profile of Panx1 in NRK cells (5).
However, a recent electrical conductance assessment provided
evidence that mouse Panx1-GFP exhibits reduced channel
function when expressed in human embryonic kidney (HEK)
293 cells (18), suggesting that its functional state is somewhat
impaired but not completely eliminated by the GFP tag. In
other studies, untagged and tetracysteine-tagged Panx1 exhib-
ited comparable capabilities in rescuing the trafficking of the
glycosylation-defective mutant of Panx1 in Madin-Darby
canine kidney cells (19). Collectively, these studies would sug-
gest that the trafficking and life cycle properties of Panx1 can be
assessed by GFP tagging approaches in conjunction with real-
time dynamic imaging.
It has been previously established through a number of stud-
ies using both untagged and GFP-tagged Cx43 that the delivery
and regeneration of Cx43 gap junction plaques is facilitated by
*This work was supported by a Canadian Institutes of Health Research oper-
ating grant (to D. W. L.)and a National Sciences and Engineering Council of
Canada studentship (to R. B.).
S
The on-line version of this article (available at http://www.jbc.org) contains
supplemental Figs. 1–5 and Movies 1–5.
1
To whom correspondence should be addressed: Dept. of Anatomy and Cell
Biology, Dental Science Bldg., University of Western Ontario, London,
Ontario N6A 5C1, Canada. E-mail: Dale.Laird@schulich.uwo.ca.
2
The abbreviations used are: NRK, normal rat kidney; HEK, human embryonic
kidney; GFP, green fluorescent protein; BFA, brefeldin A; FRAP, fluores-
cence recovery after photobleaching; BSA, bovine serum albumin; PBS,
phosphate-buffered saline; ROI, region of interest; ER, endoplasmic reticu-
lum; C-tail, carboxyl-terminal tail.
THE JOURNAL OF BIOLOGICAL CHEMISTRY VOL. 285, NO. 12, pp. 9147–9160, March 19, 2010
© 2010 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in the U.S.A.
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9147
by guest on September 18, 2015http://www.jbc.org/Downloaded from
microtubules and requires a fully functional secretory pathway
(8, 20, 21). Other studies using rapid time lapse imaging of GFP-
and tetracysteine-tagged Cx43 have shown that Cx43 is deliv-
ered in 100–150-nm vesicles that coalesce laterally into the
preexisting gap junction plaques (15, 22). However, the delivery
of pannexins to the cell surface, their dynamic organization at
specific cell surface microdomains, and their dependence on an
intact cytoskeletal network have yet to be investigated.
The life cycle of Cx43 is governed, in part, by many direct and
indirect binding partners (23–25), whereas pannexin binding
partners are only beginning to be identified with some evidence
tosupportPanx1interactionwithaproteinsubunitofthevoltage-
dependent potassium channel (26) and regulatory cross-talk
with P2X7 receptors (9, 27). In the present study, we further
expand on a very limited number of pannexin-binding proteins
by identifying actin as a specific Panx1 binding protein.
EXPERIMENTAL PROCEDURES
Cell Culture and Reagents—BICR-M1R
k
cells originally de-
rived from a rat mammary tumor and HEK 293T, NRK, and
REK (rat keratinocyte) cells were cultured in high glucose Dul-
becco’s modified Eagle’s medium (Invitrogen), supplemented
with 10% fetal bovine serum, 100 units/ml penicillin, 100
g/ml
streptomycin, and 2 mML-glutamine (all from Invitrogen). B16-
BL6 murine melanoma cells were kindly provided by Dr. Mou-
lay A. Alaoui-Jamali-(Department of Medicine and Oncology,
McGill University (Montreal, Canada)) and cultured as de-
scribed previously (28). Trypsin (0.25%, 1 mMEDTA), Opti-
MEM I medium, and Lipofectamine 2000 were purchased from
Invitrogen. Brefeldin A (BFA), nocodazole, and cytochalasin B
were purchased from Sigma.
Expression Constructs of Mouse Panx1, Panx2, and Panx3
and Engineering of the GST-Panx1 Carboxyl Domain—Un-
tagged and GFP-tagged expression constructs encoding full-
length mouse Panx1 and Panx3 were previously described (5),
and correspond to current NCBI reference sequence encoding
426 amino acids for Panx1 (NP_062355) and 392 amino acids
for Panx3 (NP_766042).
As previously described, Panx2 was originally cloned from
mouse brain, and the purified PCR product was inserted into
the EcoRI-SalI site of pEGFP-N1 vector (Clontech) to generate
the untagged Panx2 (29). The sequence was confirmed to
encode 677 amino acids according to RefSeq. Furthermore,
Panx2 cDNA was amplified using forward primer 5-CCCAA-
GCTTATGCACCACCTC to create a HindIII site and 5-GGCG-
ACCGGTCCAAACTCCACA to create an AgeI site at the 5-
and 3-ends of Panx2, respectively. PCR products and the
vector pEGFP-N1 (Clontech) were digested with HindIII and
AgeI and ligated, and clones were selected. GFP was fused in
frame to the carboxyl terminus of Panx2 with the addition of a
five-amino acid polylinker (GGACCGGTCGCCACC) and val-
idated by sequencing.
Panx1 carboxyl tail primers (forward, 5-CTAGGATTCCG-
GCAGAAAACGGAC; reverse, 5-CGAGTCGACTTAGCA-
GGACGGATT), with flanking sites for BamHI and SalI ampli-
fying the Panx1 carboxyl tail sequence (corresponding to amino
acids 299426), were created. This sequence was ligated into
the pGEX-6P-3 GST vector and transformed into BL21 bacte-
ria. Batch purification using Sepharose 4B was performed as
described in the GST Gene Fusion System manual with some
modifications (GE Healthcare). A single BL21 clone trans-
formed with either GST alone or GST-Panx1 carboxyl tail was
grown to an A
600 nm
of 1 and was induced by the addition of 0.5
mMisopropyl
-D-1-thiogalactopyranoside. A total of 500 ml of
either GST or GST-Panx1 carboxyl tail bacterial culture was
shaken overnight at room temperature. The next day, purifica-
tion was performed using 400
l of 50% Sepharose slurry,
washed in PBS seven times, and stored at 4 °C for the co-sedi-
mentation assays.
Transfection and Engineering of Stable Cell Lines—DsRed-
tagged Sar1
WT
and Sar1
H79G
cDNA constructs were described
previously (8) and used for transfection into BICR-M1R
k
cells
engineered to express Panx1 or Panx3. Briefly, BICR-M1R
k
cells were grown overnight to 50 –70% confluence on glass cov-
erslips and transfected in Opti-MEM I medium containing 4
l
of Lipofectamine 2000 and 1
g of Panx1 or 3
g of Panx3
plasmid DNA together with 2
g of Sar1
WT
or Sar1
H79G
expres-
sion constructs.
The Panx1-GFP cDNA construct was described previously
(5) and used to transfect BICR-M1R
k
cells stably expressing
untagged Panx1 and B16-BL6 cells. Cells were grown overnight
to 50–70% confluence on 35-mm glass bottom dishes and
transfected in Opti-MEM I medium containing 1.5
l of Lipo-
fectamine 2000 and 1.5
g of Panx1-GFP plasmid DNA for 4 h
at 37 °C.
For Panx3 and Panx3-GFP co-transfections, BICR-M1R
k
cells were grown overnight to 20–30% confluence on 35-mm
glass bottom dishes and transfected in Opti-MEM I medium
containing 4
l of Lipofectamine 2000 and 3
g of Panx3
together with 1.5
g of Panx3-GFP plasmid DNA. Opti-MEM I
medium was replaced with complete culture medium 4 h after
transfection at 37 °C.
For Panx2 and Panx2-GFP expression, BICR-M1R
k
cells
grown overnight to 50–70% confluence on glass coverslips
were transfected in Opti-MEM I medium containing 3
lof
Lipofectamine 2000 and 1.5
g of Panx2 or 1.5
g of Panx2-
GFP plasmid DNA. For co-transfection of Panx2 with Panx2-
GFP, 0.75
g of each plasmid DNA was used in Opti-MEM I
medium containing 3
l of Lipofectamine 2000.
Full-length constructs encoding mouse Panx1, Panx1-GFP,
and Panx3 were inserted into the AP-2 retroviral vector and
transfected into the 293GPG packaging cell line as described
earlier by Qin et al. (30). Following transfection, replication-
defective retroviral supernatants were collected and filtered
through a 0.45-
m filter (Pall Gelman Laboratories, Ann
Arbor, MI). BICR-M1R
k
cells expressing Cx43-GFP were engi-
neered to stably express Panx1 and Panx3 by following a previ-
ously described protocol (30). Retrovirus encoding Panx1-GFP
was also used to stably express GFP-tagged Panx1 in BICR-
M1R
k
cells.
Treatments with Pharmacological Reagents—Panx1- and
Panx3-overexpressing BICR-M1R
k
cells were treated with 5
g/ml BFA for 19 h at 37 °C, and cell lysates were collected and
subjected to immunoblotting. For elucidating the role of
cytoskeletal elements in pannexin trafficking, Panx1-GFP-ex-
pressing cells were treated with 10
Mnocodazole or 2.5
g/ml
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9148 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
cytochalasin B for 90 min at 37 °C and fixed for immunocyto-
chemistry. For fluorescence recovery after photobleaching
(FRAP) studies, Panx1-GFP expressing cells were pretreated
with cytoskeletal inhibitors for 90 min prior to imaging up to
3–4 h in presence of these same inhibitors.
Immunocytochemistry—Cells were immunolabeled as previ-
ously described (5). Briefly, cells grown on glass coverslips were
fixed using ice-cold 80% methanol and 20% acetone for 20 min
at 4 °C. Cytochalasin B-treated cells were fixed using 3.7%
formaldehyde for 30 min at room temperature and permeabi-
lized for 45 min in a 1% blocking solution (bovine serum albu-
min (BSA) (Sigma), containing 0.1% Triton X-100). Cells were
incubated with a 500-fold dilution of polyclonal anti-Cx43 anti-
body (Sigma), a 100-fold dilution of polyclonal anti-GPP130
antibody (Convance), polyclonal anti-Panx2 antibody (Zymed
Laboratories Inc., San Francisco, CA) or monoclonal anti-tubu-
lin antibody (Convance) for1hatroom temperature. Affinity-
purified polyclonal Panx1 and Panx3 antibodies were used at a
concentration of 2
g/ml. F-actin was localized using a 200-fold
dilution of rhodamine phalloidin (Invitrogen). Cells were incu-
bated in goat anti-rabbit antibody conjugated to Texas Red or
fluorescein isothiocyanate (1:100; Jackson Laboratories, West
Grove, PA) or a goat anti-mouse antibody conjugated to Texas
Red (1:100; Jackson Laboratory). Cells were rinsed with PBS,
and nuclei were stained with Hoechst 33342 and mounted.
Immunolabeled cells were imaged using a 63 oil objective lens
mounted on a Zeiss LSM 510 META system (Zeiss, Toronto,
Canada).
Immunoblotting and Co-immunoprecipitation—Cell lysates
from BICR-M1R
k
cells transiently co-transfected with Sar1 and
Panx1 or Panx3 cDNA constructs, and BFA-treated cells were
collected using a lysis buffer containing 1% Triton X-100, 10
mMTris, 150 mMNaCl, 1 mMEDTA, 1 mMEGTA, 0.5% Non-
idet P-40, 100 mMsodium fluoride, and 100 mMsodium
orthovanadate and a protease inhibitor tablet (one tablet/10 ml
of buffer; Roche Applied Science), pH 7.4. Protein concentra-
tions were measured using a BCA protein determination kit
(Pierce). In total, 20–30
g of protein was resolved using 10%
SDS-PAGE and transferred to nitrocellulose membrane (Pall
Life Sciences, NY). Nitrocellulose membranes were blocked in
Licor blocking solution (Lincoln, NE) or 3% BSA solution and
probed overnight with polyclonal affinity-purified anti-Panx1
or anti-Panx3 antibodies (0.2
g/ml) at 4 °C. Monoclonal anti-
-actin antibody (1:5000; Sigma) was used to assess gel loading.
Detection of primary antibody binding was performed by using
mouse IgG IR dye 800 (Rockland Immunochemicals) and rabbit
IgG Alexa 680 (Invitrogen) with the Odyssey infrared imaging
system (Licor).
For co-immunoprecipitation experiments, 1 mg of protein
lysates from WT-, Panx1-, and Panx1-GFP-overexpressing
BICR-M1R
k
cells was incubated overnight at 4 °C in the lysis
buffer (1% Triton X-100, 10 mMTris, 150 mMNaCl, 1 mM
EDTA, 1 mMEGTA, 0.5% Nonidet P-40, 1 mMsodium fluo-
ride, and 1 mMsodium orthovanadate) containing 10
g/ml
anti-Panx1 antibody. The antibody complex was pulled down
with 30
l of precleaned protein A-Sepharose beads (in PBS) for
2 h on the rocker at 4 °C. The antibody-bead complex was cen-
trifuged at 4500 rpm at 4 °C for 2 min, and supernatant was
aspirated. Unbound nonspecific protein was separated from
bound proteins by washing three times with 500
l of lysis
buffer, and the bound complex was detached by boiling for 5
min in 30
lof2Laemmli loading sample buffer containing
–mercaptoethanol. Samples were resolved by 10% SDS-PAGE
and transferred to nitrocellulose membranes, which were
probed with anti-Panx1 and anti-
-actin antibodies.
FRAP Analysis—To assess Panx1 or Panx3 dynamics at the
cell surface, BICR-M1R
k
cells expressing Panx1-GFP, alone
or in combination with the untagged Panx1, or Panx3-GFP
together with Panx3 were cultured on 35-mm glass bottom
dishes and subjected to FRAP. B16-BL6 cells expressing Panx1-
GFP were also subject to FRAP analysis. Rapid time lapse imag-
ing was performed on a Zeiss LSM 510 META system to quan-
tify the movement of Panx1-GFP into the bleached region of
interest (ROI), as described previously (16). Briefly, glass bot-
tom dishes were placed in an environmentally controlled cham-
ber, and ROIs representing the various plasma membrane
domains were selected and photobleached using scan iterations
at 488 nm with 100% laser strength. Images were acquired
2–5 s apart for up to 60 s with 0.9% laser strength to avoid
further photobleaching. Fluorescence intensities within ROIs
were quantified as described previously (16). Briefly, fluores-
cence recovery was recorded before photobleaching, immedi-
ately upon completion of photobleaching (twas set to 0 s), and
postbleaching at the following time intervals: 15, 25, 50, and
60 s. Postbleach intensities were corrected and normalized for
any residual fluorescence, and the recoverable fraction of
Panx1-GFP or Panx3-GFP was calculated using the equation,
F
Nt
(F
t
F
0
)/(F
i
F
0
) as previously described (31), where F
Nt
represents normalized fluorescence at a time point t;F
t
is fluo-
rescence intensity within ROI at tseconds postphotobleach; F
0
is fluorescence intensity upon photobleaching at t0; and F
i
is
fluorescence intensity prior to photobleaching. All FRAP
experiments were repeated three times for each experimental
set, with each set containing multiple ROIs at t15, 25, 50, and
60 s. F
Nt
values from replicates of each experimental set were
combined, and a non-linear regression analysis was per-
formed to obtain a curve of best fit using GraphPad Prism
software (San Diego, CA). To compare the mobility dynam-
ics of Panx1-GFP or Panx3-GFP in various plasma mem-
brane domains, a one-way analysis of variance followed by
Tukey’s multiple comparison tests were performed. For
comparisons between untreated and nocodazole- and
cytochalasin B-treated experimental sets, ttests were per-
formed using GraphPad Prism software.
Vesicle Movement—Panx1-GFP containing vesicles were
analyzed by measuring the total distance traveled of vesicles
that remained in confocal plane of focus for the duration of the
analysis. Vesicles of 0.5–0.8
m in diameter were monitored
by 1.8-s interval image scans for a total period of 8.8 s (n
18–20 over four independent repeats).
Actin Co-sedimentation Assays—Muscle actin (Cytoskeleton
Inc., Denver, CO) was resuspended in buffer (5 mMTris-HCl,
pH 8.0, 0.2 mMCaCl
2
) to a concentration of 1 mg/ml in 250
l
on ice for 30 min. 25
lof10actin polymerization buffer (500
mMKCl, 20 mMMgCl
2
,10mMATP, 100 mMTris, pH 7.5) was
added to the monomeric actin and incubated at room temper-
Delivery and Cell Surface Dynamics of Panx1 and Panx3
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9149
by guest on September 18, 2015http://www.jbc.org/Downloaded from
ature for 1 h (F-actin stock at 23
M). Protein samples of BSA,
GST, Panx1 carboxyl terminal tail (C-tail), and GST-tagged
Panx1 C-tail were centrifuged at 150,000 gfor1hat4°C.
Supernatants were collected and placed on ice. For co-sedi-
mentation assays, a GST fusion protein containing the C-tail of
mouse Panx1 was used, either as a fusion protein after elution
from the Sepharose beads or after cleavage of the Panx1 C-tail
from the fusion protein using PreScission Protease (GE Health-
care) for 16 h at 4 °C, as per the manufacturer’s instructions.
Following the Cytoskeleton Inc. protocol for actin binding pro-
tein assays, 50-
l samples were prepared with F-actin alone,
GST protein with or without F-actin, BSA (negative control)
plus F-actin, and Panx1 C-tail or GST fusion protein with or
without F-actin. Samples were incubated at room temperature
for 30 min, followed by centrifugation at 150,000 gfor 1.5 h at
24 °C. Supernatants were carefully removed and mixed with 10
lof4Laemmli reducing sample buffer. The pellets were
resuspended in 30
l of double-distilled water and mixed with
30
lof2Laemmli buffer. Equal volumes of resuspended
pellets and supernatants were run in duplicate on 10% SDS-
PAGE. Gels containing samples were either stained overnight
with Sypro-Ruby Protein gel stain (Invitrogen) or transferred
onto nitrocellulose membranes using an iBlot apparatus for dry
transfer (Invitrogen). Sypro-Ruby-stained gels were destained
and exposed to UV for visualization of the major bands. Nitro-
cellulose membranes were incubated overnight with primary
anti-Panx1 antibodies, washed, and probed with Alexa-680
anti-rabbit secondary for detection of Panx1 bands with a LiCor
scanner as described previously (5).
Biotinylation Assays—BICR-M1R
k
cells grown on 100-mm
dishes were transiently transfected with 5–10
g of Panx3-
GFP-encoding cDNA constructs, and 48 h post-transfection,
cells expressing Panx3-GFP were subjected to biotinylation
treatment on ice as described previously (5). Cells were incu-
bated in PBS or with cold PBS containing EZ-link Sulfo-NHS-
LC-biotin (0.5 mg/ml; Pierce) for 20 min at 4 °C. Control and
biotin-treated cells were washed and incubated in 100 mMgly-
cine buffer for 15 min at 4 °C to quench the biotin. Cells were
then lysed with SDS lysis buffer (1% Triton X-100 and 0.1% SDS
in PBS), and protein concentrations were measured using a
BCA protein determination kit (Pierce). In total, 1000
gof
protein from control and biotin-treated cell lysates were rocked
overnight at 4 °C in the presence of 50
l of neutravidin-agarose
beads (Pierce). Beads were washed three times with immuno-
precipitation lysis buffer (150 mMNaCl, 10 mMTris-HCl, pH
7.4, 1 mMEDTA, 0.5% Nonidet P-40, and 1% Triton X-100)
containing 1 mMNaF and 1 mMNa
3
VO
4
and once with PBS
containing 1 mMNaF and 1 mMNa
3
VO
4
. The beads were air-
dried and resuspended in 50
lof2Laemmli loading sam-
ple buffer containing
-mercaptoethanol before boiling for 5
min. As a lysate control, 40
g of total protein from control
and biotin samples was also resolved by SDS-PAGE and
transferred to nitrocellulose membranes for immunoblot-
ting with anti-Panx3 antibody. Glyceraldehyde-3-phosphate
dehydrogenase was used as a control to detect any unex-
pected biotin internalization.
RESULTS
We have previously shown that the Cx43-positive BICR-
M1R
k
(rat mammary tumor) cell line is an excellent reference
model for investigating dynamic delivery events and assembly
and turnover mechanisms of Cx43 (16, 32); therefore, we
designed our experimental approach to compare Panx1 and
Panx3 trafficking dynamics in Cx43-positive BICR-M1R
k
cells.
We engineered stable cells lines to ectopically express Panx1 or
Panx3. Western blots analysis revealed that wild-type BICR-
M1R
k
cells are negative for Panx1 and Panx3 (Fig. 1A), but
when engineered to express pannexins, Panx1 resolved as mul-
tiple species ranging from 41 to 48 kDa, whereas Panx3 was
detected as a doublet at 43 kDa (Fig. 1A). These multiple
pannexin species have previously been shown to be the result of
glycosylation (4, 5). Immunolabeling of both Panx1 and Panx3
revealed that both of these pannexins were capable of traffick-
ing and localizing in a relatively uniform pattern at the apposing
cell surface (Fig. 1B). In contrast, detectable Panx2 and Panx2-
GFP were mainly localized in the intracellular compartments of
not only BICR-M1R
k
cells but also HEK 293T, NRK, and rat
keratinocytes (supplemental Fig. 1). Not unexpected, the co-
expression of Panx2 with Panx2-GFP did not alter the distribu-
tion pattern of untagged or tagged Panx2. These studies indi-
cate that Panx2 has a unique distribution when compared with
Panx1 and Panx3 even when expressed in the same reference
cells.
Trafficking of Panx1 and Panx3 Is Mediated through Sar1-de-
pendent COPII Vesicles—It has been widely documented that
Sar1 (secretion-associated and Ras-related) GTPase is critical
for COPII (coat protein II) assembly and vesicular transport of
newly synthesized proteins from the endoplasmic reticulum
(ER) compartment (33, 34). Dominant negative GTP-bound
mutant Sar1
H79G
has previously been shown to block the ER-
FIGURE 1. Panx1 and Panx3 are capable of trafficking to the cell surface in
BICR-M1R
k
cells. BICR-M1R
k
cells were engineered to stably express Panx1 or
Panx3. Immunoblotting with affinity-purified antibodies (anti-Panx1 and
anti-Panx3) revealed multiple banding profiles of Panx1 (41– 48 kDa) and
Panx3 (41– 43 kDa) (A).
-Actin was used as a protein loading control (A).
Immunolabeling of Panx1 and Panx3 (B) identified that both pannexins traf-
ficked and localized to the cell surface (arrows). Nuclei are stained with
Hoechst 33342 (blue). Bars,10
m. Results shown are representative of three
independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9150 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
Golgi transport of newly synthesized Cx43 in BICR-M1R
k
cells
(8). In order to determine if Panx1 and Panx3 follow the classi-
cal ER-Golgi secretory pathway mediated through a COPII-de-
pendent mechanism, we engineered BICR-M1R
k
cells to co-
express Panx1 or Panx3 along with Sar1
WT
or Sar1
H79G
.
Co-expression of Panx1 with Sar1
WT
revealed a typical uniform
distribution of Panx1 at the cell surface (Fig. 2A,arrows),
whereas Sar1
WT
was localized to the paranuclear region in an
intact Golgi-like compartment (Fig. 2, Aand B). In comparison,
expression of Sar1
H79G
resulted in fragmentation of a Golgi-like
compartment in the paranuclear region (Fig. 2, Aand B) and
retention of Panx1 and Panx3 in ER-like patterns with little
evidence of cell surface localization (Fig. 2, Aand B,arrow-
heads). In the same culture environments, BICR-M1R
k
cells
expressing Panx1 or Panx3 but not Sar1
H79G
revealed a uniform
cell surface labeling of Panx1 and Panx3 (Fig. 2, Aand B,insets,
arrows). Furthermore, as we previously reported (8), the distri-
bution of Cx43 gap junction plaque-like structures was not evi-
dent in the presence of Sar1
H79G
(data not shown).
Because Panx1 and Panx3 resolve as multiple bands (Figs. 1A
and 2C), we wanted to determine the effect of dominant-nega-
tive Sar1 on the different molecular species. When compared
with Panx1 alone or with Sar1
WT
, the presence of Sar1
H79G
revealed an accumulation of the intermediate species (previ-
ously demonstrated to be a high mannose glycosylation species
(4)), with a noticeable reduction in the most extensively glyco-
sylated species of Panx1 (Fig. 2C). Similarly, an accumulation of
lower molecular weight species (previously reported to be a
high mannose glycosylation species (29)) of the Panx3 doublet
was also revealed in the presence of Sar1
H79G
(Fig. 2C), suggest-
ing that the higher molecular weight species of Panx1 and
Panx3 is the consequence of additional post-translational proc-
essing that occurs upon exiting the ER. Consistently, long term
BFA treatment (19 h) (a pharmacological blocker known to
inhibit anterograde transport of proteins between the ER and
Golgi apparatus (35)) of Panx1- and Panx3-expressing BICR-
M1R
k
cells caused a detectable increase in the high mannose
species of Panx1 and Panx3, with a noticeable reduction in the
higher molecular weight species of both Panx1 and Panx3 (Fig.
2D). Collectively, these results suggest that both Panx1 and
Panx3 are co-translationally inserted into the ER and trans-
ported in a COPII-dependent mechanism to the Golgi appara-
tus, where they are substrates for further glycosylation and
editing.
GFP-tagged Panx1 Mimics the Distribution Profile of Un-
tagged Panx1 and Is Suitable to Investigate the Dynamic Distri-
bution of Panx1—In order to assess the dynamic properties of
Panx1, we first stably expressed Panx1-GFP in BICR-M1R
k
cells and evaluated its distribution profile with respect to
untagged Panx1. Immunofluorescent labeling revealed that
Panx1-GFP exhibited a cell surface distribution pattern (Fig.
3A,arrows) similar to that observed for untagged Panx1 (Fig.
3F,arrows), with a notable increase in intracellular fluorescent
signal (Fig. 3A). Clearly, the cell surface pattern for Panx1-GFP
was distinct from that observed for Cx43-GFP when expressed
in the same cell type (Fig. 3A,inset).
To assess the dynamic activity of Panx1 at the cell surface and
within intracellular compartments, we performed rapid time
lapse imaging on cells that expressed Panx1-GFP. Panx1-GFP
was not only visualized in a relatively uniform pattern at the cell
surface (Fig. 3A) but also as mobile bright fluorescent clusters
(Fig. 3, Band C). Rapid time lapse imaging revealed that these
clusters (suggestive of Panx1-GFP aggregates) were mobile at
the plasma membrane and in regions devoid of cell-cell con-
tacts (Fig. 3, Band C,filled arrows; see supplemental Movie 1).
In addition, Panx1-GFP was also found in distinct intracellular
vesicle-like structures that were highly mobile (Fig. 3D,unfilled
arrows, and supplemental Movie 1). Surprisingly, and distinct
from that observed for functional Cx43, Panx1-GFP was found
in dynamic finger-like projections indicative of membrane pro-
FIGURE 2. Trafficking of Panx1 and Panx3 was disrupted in the presence
of a dominant-negative Sar1 mutant. BICR-M1R
k
cells expressing Panx1 or
Panx3 together with Sar1
WT
or Sar1
H79G
were immunolabeled for Panx1 (A)or
Panx3 (B). Both Panx1 and Panx3 were capable of trafficking and localizing to
the cell surface in the presence of Sar1
WT
(Aand B,filled arrows). Expression of
Sar1
H79G
resulted in Panx1 and Panx3 being retained in an ER-like compart-
ment (Aand B,arrowheads); however, when cells expressed Panx1 or Panx3
without expressing Sar1
H79G
in the same cellular environment, both Panx1
and Panx3 trafficked to the plasma membrane (Aand B,insets,arrows). West-
ern blotting of Panx1 and Panx3 in the presence of Sar1
H79G
(C) or after long
term BFA treatment (19 h) (D) revealed an accumulation of the high mannose
species of Panx1 and Panx3, with a noticeable reduction in the higher molec-
ular weight glycosylation species (Cand D). Nuclei are stained with Hoechst
33342 (blue). Bars,10
m. Results shown are representative of three indepen-
dent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9151
by guest on September 18, 2015http://www.jbc.org/Downloaded from
trusion (Fig. 3E,arrowheads, and supplemental Movie 1). This
localization to membrane protrusion was also frequently evi-
dent in cells expressing untagged Panx1, suggesting that this
localization profile is not a consequence of the GFP tag (Fig. 3F,
arrowheads).
Panx1-GFP Is Highly Mobile at All Plasma Membrane Do-
mains as Revealed by FRAP Analysis—To analyze the dynamic
mobility characteristics of Panx1-GFP, we first assessed the
ability of Panx1-GFP to traffic and localize at three distinct
plasma membrane domains where there was no neighboring
cell (Fig. 4A,red arrow), the neighboring cell expressed Panx1-
GFP (Fig. 4A,blue arrow), or the neighboring cell was devoid of
Panx1-GFP (Fig. 4A,purple arrow). Once it was determined
that Panx1-GFP was not differentially distributed to any of
these cell surface domains (Fig. 4A), regions of Panx1-GFP
located at these distinct domains were selected and subjected to
FRAP analysis (Fig. 4, B–E). After initial photobleaching, within
2 s, there was a rapid movement of fluorescent molecules from
the outer edges into the photobleached area of all bleached cell
surface domains (Fig. 4, B–D,insets). FRAP curve analysis
revealed the mobile fluorescent fractions to be 45–60% at all
microdomains over a time course of only 60 s (Fig. 4E). More-
over, there was no significant difference in the total recovered
fraction of Panx1-GFP within any of the plasma membrane
domains, an observation that was similar to when we ana-
lyzed the percentage recovery of transiently transfected
Panx1-GFP (40–50%) in BICR-M1R
K
cells stably expressing
untagged Panx1 (supplemental Fig. 2), or B16-BL6 cells,
identified for the first time to express endogenous Panx1
(supplemental Fig. 3, Aand B).
A Subpopulation of GFP-tagged Panx3 Is Evident at the Cell
Surface When Expressed Alone or Co-expressed with Panx3 and
Reveals Dynamic Localization to the Membrane Protrusions
We have previously reported that Panx3-GFP has a substantial
trafficking defect, causing it to be retained within the endoplas-
mic reticulum of NRK cells (5). However, when overexpressed
in BICR-M1R
k
cells, some evidence of Panx3-GFP localization
to the cell surface was detected (Fig. 5A,filled arrows), along
with the intracellular ER-like distribution of Panx3-GFP (Fig.
5A,unfilled arrows). To further confirm that a subpopulation of
Panx3-GFP can traffic and localize to the cell surface, we per-
formed biotinylation assays in live BICR-M1R
k
cells transiently
expressing Panx3-GFP. Incubation of Panx3-GFP-expressing
cells with biotin followed by pull-downs with neutravidin beads
revealed that the 70-kDa Panx3-GFP indeed traffics to the cell
surface (supplemental Fig. 4). Glyceraldehyde-3-phosphate
dehydrogenase was used as a control to confirm that biotin was
not internalized (supplemental Fig. 4). It was interesting to note
that when Panx3-GFP was co-expressed with Panx3 in BICR-
M1R
k
cells, there was an apparent increase in the cell surface
population of Panx3-GFP (Fig. 5B,filled arrows), with some
expected intracellular distribution (Fig. 5B,unfilled arrows).
This finding suggests that Panx3-GFP may interact with Panx3
to facilitate its traffic to the plasma membrane.
To assess the dynamic activity of Panx3-GFP at the cell sur-
face and within intracellular compartments, we performed
rapid time-lapse imaging on cells co-expressing Panx3 and
Panx3-GFP. Panx3-GFP was localized at the cell surface in a
relatively uniform pattern and as bright clusters either
approaching or at the cell surface (Fig. 5C,filled arrows). In
addition, Panx3-GFP was also visualized in distinct vesicle-like
structures (Fig. 5C,unfilled arrows) and observed in cell surface
protrusions (Fig. 5D,arrows) that were also occasionally iden-
tified in cells expressing only Panx3 (Fig. 5E,arrows).
Panx3-GFP Is Highly Dynamic at All Cell Surface Domains
We co-expressed Panx3 with Panx3-GFP and used GFP-tagged
Panx3 as a tracer to assess and quantify the mobile fraction in
various plasma membrane domains where the neighboring cell
surface was devoid of Panx3-GFP (Fig. 6A), the neighboring cell
expressed Panx3-GFP (Fig. 6Band supplemental Movie 2), or
there was no neighboring cell (Fig. 6C). FRAP analysis of
FIGURE 3. Panx1 was localized to multiple sites, compartments, and
microdomains. Immunolabeling of Panx1-GFP with an anti-Panx1 antibody
revealed its localization at the cell surface (A,filled arrows) in a pattern that
was distinct from Cx43-GFP (A,inset). Regions of interest from live BICR-M1R
k
cells expressing Panx1-GFP (B,red and blue rectangles) were chosen for live
imaging and imaged at t0, 10, 50, 70, and 110 s (Cand D). Rapid time lapse
imaging revealed that Panx1-GFP is distributed primarily in a uniform pattern,
whereas mobile bright fluorescent clusters could be identified at the cell sur-
face (Band C,filled arrows) and within the cell (D,unfilled arrows). Panx1-GFP
was clearly localized to dynamic plasma membrane protrusions (E,arrow-
heads) that were evident in BICR-M1R
k
cells expressing untagged Panx1
(F,arrowheads). Nuclei in Aand Fare stained with Hoechst 33342 (blue). Bars,
10
m. Representative of three independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9152 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
selected regions expressing Panx3-GFP revealed that within
60 s after photobleaching, there was a rapid reentry of GFP-
tagged Panx3 molecules into the photobleached area at all
microdomains (Fig. 6, A–C,insets),
and the mobile fraction was calcu-
lated to be 3040%. However, no
significant difference was noticed in
the total recovered fraction of Panx3-
GFP within any of the plasma mem-
brane domains (Fig. 6D).
Cell Surface Population of Panx1-
GFP and Panx3 Is Insensitive to
Nocodazole Treatment—It has been
documented that nocodazole dis-
ruption of microtubules impairs the
continuous trafficking and regener-
ation of Cx43-GFP at the cell sur-
face (8). To investigate if trafficking
and recovery of Panx1-GFP into the
photobleached area is dependent on
microtubules, Panx1-GFP-express-
ing BICR-M1R
k
cells were exposed
to nocodazole for 90 min. Nocoda-
zole treatment resulted in charac-
teristic morphological changes in
the cells from spindle-shaped to
more cuboidal with a notable dis-
ruption of the microtubule architec-
ture (Fig. 7, A–C). However, the dis-
tribution of Panx1-GFP (Fig. 7B)
and Panx3 (supplemental Fig. 5B)at
the cell surface remained relatively
unaffected by the nocodazole treat-
ment. Furthermore, FRAP studies
revealed that Panx1-GFP migrated
into the photobleached area simi-
larly in both untreated and nocoda-
zole-treated cells (Fig. 7, Cand D).
Interestingly, the inward progres-
sion of fluorescent recovery from
the edges of the photobleached area
toward the center was not detected
until 10 s postbleaching (Fig. 7C,
inset). Furthermore, FRAP analysis
of Panx1-GFP in the presence of
nocodazole treatment revealed no
significant difference in the recov-
ered fraction when compared with
the untreated cells (Fig. 7D). Thus,
nocodazole treatment does not
visually affect the cell surface local-
ization of Panx1-GFP or the dynam-
ics of recovery into the photo-
bleached area at the cell surface.
The Cell Surface Stability of
Panx1-GFP and Panx3 Is Sensi-
tive to Cytochalasin B Treatment,
whereas the Mobility of Panx1-GFP
Transport Vesicles Is Perturbed in the Absence of Intact
Microfilaments—We next wanted to assess the role of actin
microfilaments in stabilizing Panx1 and Panx3 at the cell sur-
FIGURE 4. Panx1-GFP is highly mobile at all plasma membrane locations. Panx1-GFP was localized to three
distinct plasma membrane domains of BICR-M1R
k
cells, as depicted by the schematic diagram (A). Fluorescent
images of Panx1-GFP were superimposed with DIC images to highlight the microenvironment surrounding the
cell being analyzed (A–D). Selected cell regions where Panx1-GFP was localized at the three distinct plasma
membrane domains were photobleached, and fluorescence recovery back into the photobleached areas was
assessed and normalized over 60 s (B–D). Panx1-GFP recovery within the photobleached area was not signifi-
cantly different (p0.05) among all three domains (E). Bars,10
m. n6 –9 per plasma membrane domain
collected from three independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9153
by guest on September 18, 2015http://www.jbc.org/Downloaded from
face as well as evaluating their cell surface dynamic properties
and transport of Panx1-GFP. Treatment of cytochalasin B (90
min) caused a redistribution of F-actin from the cell periphery
(Fig. 8Aand supplemental Fig. 5A) to the paranuclear region
(Fig. 8Band supplemental Fig. 5A), with a subsequent change in
cell morphology. In cytochalasin B-treated cells, Panx1-GFP
and Panx3 were mainly localized in intracellular compartments
(Fig. 8Band supplemental Fig. 5A,arrowheads), whereas a
small population remained evident at the cell surface (Fig. 8B
and supplemental Fig. 5A, arrows). These findings suggest that
actin may play a crucial role in the cell surface stability of
Panx1-GFP and Panx3. FRAP analysis of the remaining and
detectable cell surface population of Panx1-GFP after cytocha-
lasin B treatment (supplemental Movie 3) revealed that Panx1-
GFP was mobile, but the recovery rate was slower and the
amount of the recovered fraction was significantly less (15–
20%) when compared with the untreated cells (Fig. 8C). Fur-
thermore, rapid time lapse imaging on the same field of cells
visualized before (supplemental Movie 4) and after
(supplemental Movie 5) cytochalasin B treatment revealed that
Panx1-GFP-carrying vesicles were freely able to move and
travel in the untreated cells; however, this movement was
greatly perturbed when cells were treated with cytochalasin B
with a 60% decrease in vesicle movement over a fixed interval
of time (8.8 s) (Fig. 8D). Collectively, these studies suggest that
microfilaments play a multifaceted role in pannexin stabiliza-
tion at the cell surface and vesicular transport.
F-actin Directly Binds Panx1 at the Carboxyl Terminus—To
further examine the possible interaction of Panx1 with actin,
lysates of wild-type BICR-M1R
k
cells engineered to express
Panx1 or Panx1-GFP were subjected to immunoprecipitation
of Panx1 prior to immunoblotting for Panx1 or
-actin. As
expected, multiple glycosylated species of Panx1 (resolved
below the IgG band) and Panx1-GFP were detected in the
immunoprecipitates and cell lysates of Panx1-overexpressing
cells but not wild-type BICR-M1R
k
cells (Fig. 9A,top). Interest-
ingly,
-actin was found to co-immunoprecipitate with Panx1
from both Panx1- and Panx1-GFP-expressing BICR-M1R
k
cells
(Fig. 9A,bottom). To further validate the interaction of actin
with Panx1, we conducted co-sedimentation assays where poly-
merized actin was mixed with GST fusion protein containing
the C-tail of Panx1. As observed by Sypro-stained gel, once
polymerized, F-actin typically sediments in the pellet fraction.
In the absence of polymerized actin, GST-Panx1 C-tail was
found in both the soluble and pellet fractions (Fig. 9B). How-
ever, when combined with F-actin, GST-Panx1 C-tail sedi-
ments preferentially in the pellet fraction (Fig. 9B). As a control,
GST alone did not sediment in the pellet fraction with or with-
out F-actin (Fig. 9B). Because the Panx1 C-tail fusion protein
detection at 43 kDa was partially masked by actin, we cleaved
the C-tail of Panx1 from the GST and performed a similar co-
sedimentation assay. Further Panx1 immunoblots revealed the
Panx1 C-tail at 15 kDa in the pellet fraction (Fig. 9C). These
results suggest that F-actin binds directly to the carboxyl termi-
nus of Panx1.
DISCUSSION
With the recent discovery of pannexins as a new family of
channel- or conduit-forming proteins, there has been a
growing interest in elucidating their biochemical properties,
life cycle, and cellular roles. Previous reports have linked
Panx1 channels to the release of ATP in neurons (36) and
astrocytes (37) and in cellular response to pathological
insults, such as initiation of inflammatory action (10, 38),
ischemia-induced death of neurons (39), and tumor suppres-
sion (13). Our understanding of Panx3 is even more rudi-
mentary. Although Panx3 has been shown to be a cell surface
glycoprotein that forms conduits capable of dye uptake (5,
40), its cellular function remains largely unknown. It was our
hypothesis that, being integral membrane proteins with
sequence relationships to invertebrate innexin gap junction
proteins, Panx1 and Panx3 would exhibit characteristics similar
to those of the well studied Cx43 gap junction protein in terms
of the secretory pathway governing their trafficking, dynamic
FIGURE 5. Delivery of Panx3-GFP to the cell surface. Wild-type BICR-M1R
k
cells engineered to express Panx3-GFP (A) or both Panx3 and Panx3-GFP (B)
were immunolabeled with anti-Panx3 antibody. Panx3-GFP was retained
mainly in an ER-like pattern (Aand B,unfilled arrows) with some evidence of a
cell surface distribution (Aand B,filled arrows), whereas co-expression of
Panx3 appeared to increase the cell surface population of Panx3-GFP (B).
Rapid time lapse imaging of live BICR-M1R
k
cells co-expressing Panx3-GFP
and Panx3 revealed that Panx3-GFP was distributed primarily in a uniform
pattern with notable mobile fluorescent clusters at the cell surface (C,filled
arrows) and within the cell (C,unfilled arrow). Localization of Panx3-GFP to
plasma membrane protrusions (D,arrows) was similar to that found in cells
expressing only Panx3 (E,arrows). Nuclei in A,B, and Eare stained with
Hoechst 33342 (blue). Bars,10
m. Results shown are representative of three
independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9154 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
properties within the plasma membrane, and interplay with the
cytoskeletal network.
Trafficking of Panx1 and Panx3 to the Cell Surface—It has
been well established that Cx43 is co-translationally inserted
into the ER and oligomerizes into connexons in the trans-Golgi
network before trafficking to the plasma membrane (25). In the
case of Panx1, treatment with endoglycosidase H and peptide:
N-glycosidase F enzymes have revealed that the lower molecu-
lar weight species constitutes the core protein that gets glyco-
sylated to the high mannose form in the ER before further
glycosylation and editing in the Golgi and delivery to the cell
surface (4, 5). Although recent studies have extensively focused
on correlating glycosylation status with the delivery of Panx1 to
the cell surface (4, 19), molecular mechanisms underlying the
secretory pathway taken by Panx1 and Panx3 were largely
unknown. To address if the transport of Panx1 and Panx3
from the ER to the Golgi apparatus is mediated through
COPII vesicles, we transiently co-expressed Panx1 or Panx3
with a dominant-negative GTP-bound mutant Sar1
H79F
. Stabi-
lization of Sar1 in the dominant negative GTP-bound state has
been previously shown to efficiently
block COPII-mediated ER transport
of proteins to the cell surface (41).
Our data demonstrated that
Sar1
H79F
expression severely inhib-
ited the cell surface localization of
Panx1 and Panx3, thus suggesting
that COPII vesicular trafficking of
these pannexin family members is
required prior to their eventual
delivery to the cell surface. These
data also indicate that efficient GTP
hydrolysis of Sar1 is crucial for reg-
ulating their transport from the ER
compartment because restriction of
GTP hydrolysis of Sar1
H79G
to GDP
has been previously shown to arrest
the cargo-containing vesicles at the
ER exit sites (34). Consistent with
our findings, we previously showed
that Sar1 function was necessary for
delivery of Cx43 to the cell surface
(8). Our data support the premise
that the post-ER pool of Panx1 and
Panx3 is correlated with pannexin
processing to the highly complex
glycosylation species, whereas the
accumulation of the high mannose
intermediate species of Panx1 and
Panx3 is consistent with retention
of Panx1 and Panx3 in the ER. Sim-
ilar Sar1 GTPase dependence was
observed for K
ATP
channels, where
expression of dominant negative
mutants Sar1
H79G
(mutant incapa-
ble of hydrolyzing GTP) and
Sar1
T39N
(mutant restricted in its
ability to exchange GDP for GTP)
prevented proper channel processing and the cell surface
expression of the channel (42).
ER to Golgi transport of Panx1 and Panx3, as well as their
state of glycosylation, was further confirmed by the use of BFA,
which is known to inactivate Arf1, thereby inhibiting ER to
Golgi transport (43). Consistent with a previous study (19), we
noticed a dramatic increase in the high mannose form of Panx1
in BFA-treated cells. Interestingly, we also observed an accu-
mulation of the high mannose species of Panx3 in response to
BFA treatment, consistent with the inhibition of both Panx1
and Panx3 being delivered to the Golgi apparatus for further
processing. Vesicular trafficking of Panx1 and Panx3 would also
strongly suggest that, like Cx43, both proteins are integral
transmembrane proteins that get transported from ER mem-
branes in COPII vesicles.
Mobility Dynamics of Panx1 and Panx3—In our study, we
used Panx1-GFP and Panx3-GFP as tracer probes to elucidate
the distribution profile and cell surface dynamics of Panx1 and
Panx3. The distribution of Panx1-GFP at the plasma membrane
appeared to be uniform with occasional Panx1-enriched
FIGURE 6. Panx3-GFP is highly mobile at all plasma membrane domains. Panx3-GFP was localized to three
distinct plasma membrane domains of BICR-M1R
k
cells co-expressing Panx3 (A–C). Selected cell surface regions
containing Panx3-GFP were photobleached, and fluorescence recovery back into the photobleached areas
was assessed and normalized over the time course of 60 s. The percentage of Panx3-GFP recoverable fraction
was not found to be significantly different among all three plasma membrane domains examined (p0.05)
(D). Bars,10
m. n12–25 per plasma membrane domain collected from three independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9155
by guest on September 18, 2015http://www.jbc.org/Downloaded from
domains, consistent with previous findings using the untagged
Panx1 (5). In contrast, Panx3-GFP alone revealed an increased
intracellular profile, as reported earlier (5), with some clear evi-
dence of its localization to the cell surface that was confirmed
by cell surface biotinylation assays. Since adequate delivery of
Panx3-GFP to the cell surface was required to investigate its
dynamics at the plasma membrane, we co-expressed it with
untagged Panx3 and found an increased cell surface expression
of Panx3-GFP. It is possible that Panx3 co-oligomerized with
Panx3-GFP to facilitate its delivery to the plasma membrane; if
this holds true, it would also suggest that having a GFP tag at the
carboxyl-terminal tail of Panx3 does not interfere with the
intermixing of Panx3 subunits. A similar mechanism has been
proposed for the cell surface rescue of a trafficking-defective,
glycosylation-deficient mutant of Panx1 when co-expressed
with either wild-type Panx1 or tet-
racysteine-tagged Panx1, further
indicating that tagging of pannexins
does not impair its ability to assem-
ble together with its untagged coun-
terpart (19).
In our study, delivery of both
Panx1-GFP and Panx3-GFP to the
cell surface appeared to occur at
multiple plasma membrane do-
mains via intracellular vesicle-like
structures that formed bright clus-
ters upon apparent fusion with the
plasma membrane. These clusters
were found to be quite mobile and
displaced laterally within the cell
surface membrane, thus supporting
a model of untargeted delivery of
Panx1 and Panx3 to all cell surface
microdomains. In contrast, Cx43-
GFP, typically known to localize in
punctate-like structures at the cell
surface, has been documented to
have both an arbitrary delivery to all
plasma membrane domains (8, 16)
as well as a preferred microtubule-
dependent delivery to adherens
junctions that reside in close prox-
imity to the preexisting gap junc-
tions (44). Interestingly, our data
support the premise that Panx1 and
Panx3 are enriched in membrane
protrusions at areas that are devoid
of contacting cells, a situation not
typically observed for Cx43 unless
non-functional Cx43 mutant stud-
ies are performed (16). Localization
of pannexins in the finger-like
membrane protrusions could sug-
gest that it may play a role in cell
migration because the process of
cell motility typically involves actin
polymerization and entails forma-
tion of fanlike or pointed projections (lamellipodium and filo-
podia, respectively) at the leading edge (45). Previously, the
absence of Panx1 from the leading edge of a corneal epithelium
wound in P
2
X
7
/
mice was correlated with delayed corneal
re-epithelialization and compromised wound healing (46).
Other channel-forming proteins, such as aquaporin-1, have
been implicated in increased cell migration by localizing to the
lamellipodia (47), whereas migration of lymphocytes (48) and
embryonic nerve cells (49) has been correlated with voltage-de-
pendent K
channels, thus supporting the role of channel-
forming proteins in regulating cell motility.
Our FRAP assessments of Panx1 and Panx3 mobility identi-
fied that, similar to Cx43 (16), lateral movement of Panx1-GFP
and Panx3-GFP occurred from the outer edge to the center of
the photobleached areas. It is notable that although the recov-
FIGURE 7. The cell surface population of Panx1-GFP is insensitive to nocodazole treatment. Untreated (A)
or nocodazole-treated (B) Panx1-GFP-expressing BICR-M1R
k
cells were immunolabeled for tubulin. As
expected, nocodazole treatment collapsed tubulin into paranuclear regions (Aand B); however, the distribu-
tion profile of Panx1-GFP at the cell surface and in the intracellular compartments (Aand B) remained relatively
unchanged with collapsed tubulin (B). FRAP analysis in presence of nocodazole revealed that Panx1-GFP was
able to recover into the photobleached area, and the percentage of recoverable fraction was not significantly
different from the untreated cells (Cand D). Bars,10
m. n5–10 per plasma membrane domains, data
collected over four independent repeats.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9156 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
ery of fluorescence into the photobleached area is probably the
result of lateral movement of pannexins due to the time course
being examined, there is also probably a contribution from
newly delivery fluorescent protein-tagged pannexins to the cell
surface. The relative rate of Panx1-GFP and Panx3-GFP recov-
ery into the photobleached area was quite comparable in any of
the examined plasma membrane microdomains, thus suggest-
ing that the assembly state of these pannexin family members
remains relatively unchanged with respect to its subcellular
location within the plasma membrane. In our study, Panx1-
GFP (and to a slightly lesser extent Panx3-GFP) exhibited over
twice the mobility of Cx43-GFP molecules, which are typically
arranged in gap junction-like clus-
ters (16). Slow recovery of Cx43-
GFP was also reported in HeLa cells
(50). Given the relatively uniform
cell surface distribution of Panx1
and Panx3 that appear to be untar-
geted to specific microdomains, we
speculate that these pannexins are
not likely to be packaged into dense
crystalline-like structures as re-
ported for Cx43 (51). Thus, the dis-
tribution and mobility of these
pannexins are more in line with
other channels and receptors, such
as Na
channels (52) and acetylcho-
line receptors (53).
The percentage of fluorescence
recovery after photobleaching for
both Panx1-GFP and Panx3-GFP
(representing the mobile fraction)
reached a plateau between 40 and
60%, with the remaining compo-
nent representing the immobile
fraction. Typically, the size of the
immobile fraction is dependent on
the nature of the protein and the
membrane microenvironment being
assessed. For instance, the immobile
fraction of sodium channels ranges
from 10% in the cell body to 40%
in the neurite terminals (52). Like-
wise, the mobile fraction of glycine
receptors ranges from 50% in the
neuronal cell body to 70% in the
processes (54). In addition, the vis-
cosity of the membrane microdo-
main (55), tethering of proteins with
scaffolds/binding partners, and
interaction with the cytoskeleton
can all contribute to the size of the
immobile fraction (56).
Cytoskeletal Dependence of
Pannexin Trafficking and Mobility
In order to assess the role of the
cytoskeleton in pannexin trafficking
and cell surface mobility dynamics,
we used nocodazole and cytochalasin B to disrupt microtubules
and microfilaments, respectively. Nocodazole-induced disrup-
tion of microtubules did not significantly alter the cell surface
distribution of either Panx3 or Panx1-GFP, which may not be
totally unexpected given the predicted long half-life of Panx1 (4,
5). This finding is quite distinct from Cx43, where enhanced
growth of gap junctions (20) and Cx43 molecular movement
into the photobleached gap junctions was minimal in nocoda-
zole-treated cells (8), suggesting that Cx43 is much more de-
pendent on microtubules than Panx1. In contrast, disruption of
microfilaments revealed concomitant accumulation of paranu-
clear Panx1-GFP and Panx3 with collapsed actin microfila-
FIGURE 8. Effect of cytochalasin B on Panx1-GFP. Untreated (A) or cytochalasin B-treated (B) Panx1-GFP-
expressing BICR-M1R
k
cells were labeled with phalloidin for F-actin localization. As expected, cytochalasin B
caused the redistribution of F-actin from the cell surface (A) to the paranuclear region (B). The collapse of F-actin
microfilaments coincided with the intracellular accumulation of Panx1-GFP (B,arrowheads), whereas a small
population of Panx1-GFP remained evident at the cell surface (B,arrows). FRAP analysis in the presence of
cytochalasin B treatment revealed that the cell surface population of Panx1-GFP was significantly impaired
from entering the photobleached area (p0.05) (C)n3. Quantification of the total distance traveled by
Panx1-GFP carrying vesicles within the same field of cells analyzed before and after the cytochalasin B treat-
ment indicated a significant (p0.05) reduction in vesicle mobility in cytochalasin B-treated cells (D). Bars,10
m. Results shown are representative of five independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9157
by guest on September 18, 2015http://www.jbc.org/Downloaded from
ments, whereas only a subpopulation of Panx1-GFP or Panx3
remained at the cell surface. The longer turnover dynamics of
pannexins and the rapid intracellular accumulation upon short
term cytochalasin B treatment support the premise that actin
provides stability to the cell surface population of Panx3 and
Panx1. On the other hand, Cx43 gap junctions have been doc-
umented to remain considerably more independent to the
assembly state of microfilaments (57). Mobility assessment of
FIGURE 9. F-actin binds Panx1 at the carboxyl terminus. Wild type (WT) or Panx1- or Panx1-GFP-expressing BICR-M1R
k
cells were lysed and subjected to
immunoprecipitation (IP) for Panx1 prior to immunoblotting (IB) the immunoprecipitates and cell lysates for Panx1 or
-actin.
-Actin co-immunoprecipitated
with Panx1 and Panx1-GFP (A). Monomeric actin was polymerized into F-actin, incubated with either GST fusion protein containing the carboxyl-terminal tail
of Panx1 (B) or the carboxyl-terminal tail of Panx1 alone (C) and separated into supernatant or pellet fractions (denoted by Sand P, respectively) prior to
immunoblotting for Panx1. Panx1 was found to co-sediment with F-actin in the pellet fractions (Band C). Parallel gels were stained with Sypro gel stain, and BSA
and GST were used as controls in the co-sedimentation assays. Results shown are representative of three independent experiments.
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9158 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
the remaining cytochalasin B-insensitive subpopulation of
Panx1-GFP at the cell surface revealed a smaller mobile fraction
of 35% in the absence of intact microfilaments. This surpris-
ing finding was somewhat distinct from Na,K-ATPase and reg-
gie-1/flotillin-2, where the disruption of microfilaments caused
an increase in the mobile fraction (58, 59). The smaller mobile
fraction of Panx1-GFP noticed in our study may be explained by
the fact that transport vesicles carrying Panx1-GFP either to or
from the plasma membrane were less mobile, and this decrease
may mechanistically account for the reduced fluorescent recov-
ery into the photobleached area. The relative speeds of Panx1-
GFP-carrying vesicles in untreated cells (0.2
m/s) and
cytochalasin B-treated cells (0.07
m/s) were also quite dif-
ferent. Although credited to trafficking on microtubules tracks,
vesicles containing Cx43-GFP showed comparable average
speeds of 0.5
m/s in untreated HeLa cells (21).
Interaction of Panx1 with Actin—Given the finding that actin
microfilaments regulated the distribution and cell surface
mobility of Panx1, we speculated that a direct interaction
between Panx1 and actin may exist. Because we were best
equipped to address this question for Panx1, we first demon-
strated that actin does in fact co-immunoprecipitate with
Panx1. To further assess 1) if Panx1 binds to monomeric or
filamentous actin, 2) if Panx1 interaction to actin is direct, and
3) which domain of Panx1 might be responsible for interaction,
we conducted a co-sedimentation assay with a GST-Panx1
C-tail fusion protein. Here we show that Panx1 sediments pref-
erentially in the F-actin fraction, the interaction appears to be
direct, and it is the carboxyl-terminal tail of Panx1 that seems
responsible for interacting with actin. Comparatively, actin
binding to Cx43 is thought to be facilitated via its interaction
with zonula occludens-1, which is a known Cx43 binding part-
ner (60). Future studies will be needed to elucidate the Panx1
motif responsible for actin binding at the C-tail and whether
actin also binds to Panx3.
In summary, trafficking and assembly of pannexins is a pre-
cisely regulated process. Our study is the first to report that
Panx1 and Panx3 transport is dependent on Sar1-mediated
COPII vesicles, cell surface Panx1 and Panx3 have dynamic
mobility properties, and Panx1 directly interacts with F-actin.
Acknowledgment—We acknowledge Jamie Simek for providing exper-
tise in optimizing the FRAP technique.
REFERENCES
1. Panchin, Y., Kelmanson, I., Matz, M., Lukyanov, K., Usman, N., and Luky-
anov, S. (2000) Curr. Biol. 10, R473–R474
2. Willecke, K., Eiberger, J., Degen, J., Eckardt, D., Romualdi, A., Gu¨ldenagel,
M., Deutsch, U., and So¨hl, G. (2002) Biol. Chem. 383, 725–737
3. Baranova, A., Ivanov, D., Petrash, N., Pestova, A., Skoblov, M., Kelmanson,
I., Shagin, D., Nazarenko, S., Geraymovych, E., Litvin, O., Tiunova, A.,
Born, T. L., Usman, N., Staroverov, D., Lukyanov, S., and Panchin, Y.
(2004) Genomics 83, 706–716
4. Boassa, D., Ambrosi, C., Qiu, F., Dahl, G., Gaietta, G., and Sosinsky, G.
(2007) J. Biol. Chem. 282, 31733–31743
5. Penuela, S., Bhalla, R., Gong, X. Q., Cowan, K. N., Celetti, S. J., Cowan, B. J.,
Bai, D., Shao, Q., and Laird, D. W. (2007) J. Cell Sci. 120, 3772–3783
6. Laird, D. W., Puranam, K. L., and Revel, J. P. (1991) Biochem. J. 273,
67–72
7. Musil, L. S., and Goodenough, D. A. (1993) Cell 74, 1065–1077
8. Thomas, T., Jordan, K., Simek, J., Shao, Q., Jedeszko, C., Walton, P., and
Laird, D. W. (2005) J. Cell Sci. 118, 4451–4462
9. Iglesias, R., Locovei, S., Roque, A., Alberto, A. P., Dahl, G., Spray, D. C., and
Scemes, E. (2008) Am. J. Physiol. Cell Physiol. 295, C752–C760
10. Pelegrin, P., and Surprenant, A. (2007) J. Biol. Chem. 282, 2386–2394
11. Dahl, G., and Locovei, S. (2006) IUBMB Life 58, 409–419
12. Barbe, M. T., Monyer, H., and Bruzzone, R. (2006) Physiology 21,
103–114
13. Lai, C. P., Bechberger, J. F., Thompson, R. J., MacVicar, B. A., Bruzzone, R.,
and Naus, C. C. (2007) Cancer Res. 67, 1545–1554
14. Bruzzone, R., Hormuzdi, S. G., Barbe, M. T., Herb, A., and Monyer, H.
(2003) Proc. Natl. Acad. Sci. U.S.A. 100, 13644–13649
15. Gaietta, G., Deerinck, T. J., Adams, S. R., Bouwer, J., Tour, O., Laird, D. W.,
Sosinsky, G. E., Tsien, R. Y., and Ellisman, M. H. (2002) Science 296,
503–507
16. Simek, J., Churko, J., Shao, Q., and Laird, D. W. (2009) J. Cell Sci. 122,
554–562
17. Bukauskas, F. F., Jordan, K., Bukauskiene, A., Bennett, M. V., Lampe, P. D.,
Laird, D. W., and Verselis, V. K. (2000) Proc. Natl. Acad. Sci. U.S.A. 97,
2556–2561
18. Ma, W., Hui, H., Pelegrin, P., and Surprenant, A. (2009) J. Pharmacol. Exp.
Ther. 328, 409–418
19. Boassa, D., Qiu, F., Dahl, G., and Sosinsky, G. (2008) Cell. Commun. Adhes.
15, 119–132
20. Johnson, R. G., Meyer, R. A., Li, X. R., Preus, D. M., Tan, L., Grunenwald,
H., Paulson, A. F., Laird, D. W., and Sheridan, J. D. (2002) Exp. Cell Res.
275, 67–80
21. Lauf, U., Giepmans, B. N., Lopez, P., Braconnot, S., Chen, S. C., and Falk,
M. M. (2002) Proc. Natl. Acad. Sci. U.S.A. 99, 10446–10451
22. Jordan, K., Solan, J. L., Dominguez, M., Sia, M., Hand, A., Lampe, P. D., and
Laird, D. W. (1999) Mol. Biol. Cell 10, 2033–2050
23. Giepmans, B. N., Verlaan, I., Hengeveld, T., Janssen, H., Calafat, J., Falk,
M. M., and Moolenaar, W. H. (2001) Curr. Biol. 11, 1364–1368
24. Herve´, J. C., Bourmeyster, N., Sarrouilhe, D., and Duffy, H. S. (2007) Prog.
Biophys. Mol. Biol. 94, 29–65
25. Laird, D. W. (2006) Biochem. J. 394, 527–543
26. Bunse, S., Haghika, A., Zoidl, G., and Dermietzel, R. (2005) Cell. Commun.
Adhes. 12, 231–236
27. Pelegrin, P., and Surprenant, A. (2006) EMBO J. 25, 5071–5082
28. Fan, D., Liaw, A., Denkins, Y. M., Collins, J. H., Van Arsdall, M., Chang,
J. L., Chakrabarty, S., Nguyen, D., Kruzel, E., and Fidler, I. J. (2002) J. Exp.
Ther. Oncol. 2, 286–297
29. Penuela, S., Bhalla, R., Nag, K., and Laird, D. W. (2009) Mol. Biol. Cell 20,
4313–4323
30. Qin, H., Shao, Q., Curtis, H., Galipeau, J., Belliveau, D. J., Wang, T., Alaoui-
Jamali, M. A., and Laird, D. W. (2002) J. Biol. Chem. 277, 29132–29138
31. Lippincott-Schwartz, J., Snapp, E., and Kenworthy, A. (2001) Nat. Rev.
Mol. Cell Biol. 2, 444–456
32. Laird, D. W., Castillo, M., and Kasprzak, L. (1995) J. Cell Biol. 131,
1193–1203
33. Lee, M. C., and Miller, E. A. (2007) Semin. Cell Dev. Biol. 18, 424–434
34. Dong, C., Zhou, F., Fugetta, E. K., Filipeanu, C. M., and Wu, G. (2008) Cell.
Signal. 20, 1035–1043
35. Lippincott-Schwartz, J., Yuan, L. C., Bonifacino, J. S., and Klausner, R. D.
(1989) Cell 56, 801–813
36. Thompson, R. J., Jackson, M. F., Olah, M. E., Rungta, R. L., Hines, D. J.,
Beazely, M. A., MacDonald, J. F., and MacVicar, B. A. (2008) Science 322,
1555–1559
37. Scemes, E., Suadicani, S. O., Dahl, G., and Spray, D. C. (2007) Neuron. Glia
Biol. 3, 199–208
38. Kanneganti, T. D., Lamkanfi, M., Kim, Y. G., Chen, G., Park, J. H., Franchi,
L., Vandenabeele, P., and Nu´n˜ez, G. (2007) Immunity 26, 433– 443
39. Thompson, R. J., Zhou, N., and MacVicar, B. A. (2006) Science 312,
924–927
40. Penuela, S., Celetti, S. J., Bhalla, R., Shao, Q., and Laird, D. W. (2008) Cell.
Commun. Adhes. 15, 133–142
41. Sato, K. (2004) J. Biochem. 136, 755–760
Delivery and Cell Surface Dynamics of Panx1 and Panx3
MARCH 19, 2010VOLUME 285• NUMBER 12 JOURNAL OF BIOLOGICAL CHEMISTRY 9159
by guest on September 18, 2015http://www.jbc.org/Downloaded from
42. Taneja, T. K., Mankouri, J., Karnik, R., Kannan, S., Smith, A. J., Munsey, T.,
Christesen, H. B., Beech, D. J., and Sivaprasadarao, A. (2009) Hum. Mol.
Genet. 18, 2400–2413
43. Donaldson, J. G., Finazzi, D., and Klausner, R. D. (1992) Nature 360,
350–352
44. Shaw, R. M., Fay, A. J., Puthenveedu, M. A., von Zastrow, M., Jan, Y. N.,
and Jan, L. Y. (2007) Cell 128, 547–560
45. Rorth, P. (2009) Annu. Rev. Cell Dev. Biol. 25, 407–429
46. Mayo, C., Ren, R., Rich, C., Stepp, M. A., and Trinkaus-Randall, V. (2008)
Invest. Ophthalmol. Vis. Sci. 49, 4384–4391
47. Saadoun, S., Papadopoulos, M. C., Hara-Chikuma, M., and Verkman, A. S.
(2005) Nature 434, 786–792
48. Levite, M., Cahalon, L., Peretz, A., Hershkoviz, R., Sobko, A., Ariel, A.,
Desai, R., Attali, B., and Lider, O. (2000) J. Exp. Med. 191, 1167–1176
49. Hendriks, R., Morest, D. K., and Kaczmarek, L. K. (1999) J. Neurosci. Res.
58, 805–814
50. Falk, M. M., Baker, S. M., Gumpert, A. M., Segretain, D., and Buckheit,
R. W., 3rd (2009) Mol. Biol. Cell 20, 3342–3352
51. Unger, V. M., Kumar, N. M., Gilula, N. B., and Yeager, M. (1999) Novartis
Found. Symp. 219, 22–30; discussion 31–43
52. Angelides, K. J., Elmer, L. W., Loftus, D., and Elson, E. (1988) J. Cell Biol.
106, 1911–1925
53. Poo, M. (1982) Nature 295, 332–334
54. Srinivasan, Y., Guzikowski, A. P., Haugland, R. P., and Angelides, K. J.
(1990) J. Neurosci. 10, 985–995
55. Klausner, R. D., Kleinfeld, A. M., Hoover, R. L., and Karnovsky, M. J. (1980)
J. Biol. Chem. 255, 1286–1295
56. Nigg, E. A., and Cherry, R. J. (1980) Proc. Natl. Acad. Sci. U.S.A. 77,
4702–4706
57. Wang, Y., and Rose, B. (1995) J. Cell Sci. 108, 3501–3508
58. Langhorst, M. F., Solis, G. P., Hannbeck, S., Plattner, H., and Stuermer,
C. A. (2007) FEBS Lett. 581, 4697–4703
59. Paller, M. S. (1994) J. Membr. Biol. 142, 127–135
60. Giepmans, B. N., and Moolenaar, W. H. (1998) Curr. Biol. 8, 931–934
Delivery and Cell Surface Dynamics of Panx1 and Panx3
9160 JOURNAL OF BIOLOGICAL CHEMISTRY VOLUME 285NUMBER 12• MARCH 19, 2010
by guest on September 18, 2015http://www.jbc.org/Downloaded from
Churko, Qing Shao and Dale W. Laird
Ruchi Bhalla-Gehi, Silvia Penuela, Jared M.
Interactions
Surface Dynamics, and Cytoskeletal
Pannexin1 and Pannexin3 Delivery, Cell
Cell Biology:
doi: 10.1074/jbc.M109.082008 originally published online January 10, 2010
2010, 285:9147-9160.J. Biol. Chem.
10.1074/jbc.M109.082008Access the most updated version of this article at doi:
.JBC Affinity SitesFind articles, minireviews, Reflections and Classics on similar topics on the
Alerts:
When a correction for this article is posted When this article is cited
to choose from all of JBC's e-mail alertsClick here
Supplemental material:
http://www.jbc.org/content/suppl/2010/01/10/M109.082008.DC1.html
http://www.jbc.org/content/285/12/9147.full.html#ref-list-1
This article cites 60 references, 30 of which can be accessed free at
by guest on September 18, 2015http://www.jbc.org/Downloaded from
... 43 Panx1 channels are typically localized in the sarcolemma, but Panx1 has also been found in other cell compartments such as the endoplasmic reticulum and Golgi apparatus. 37,44 Trafficking of Panx1 is regulated by its glycosylation status 45 and mediated by its interaction with cytoskeletal proteins, such as actin microfilaments. 44 Indeed, the previous proteomic analysis identified cytoskeletal proteins as well as mitochondrial proteins like VDAC1, ANT2, and heat shock protein 70 (Hsc70) as Panx1 partners. ...
... 37,44 Trafficking of Panx1 is regulated by its glycosylation status 45 and mediated by its interaction with cytoskeletal proteins, such as actin microfilaments. 44 Indeed, the previous proteomic analysis identified cytoskeletal proteins as well as mitochondrial proteins like VDAC1, ANT2, and heat shock protein 70 (Hsc70) as Panx1 partners. 34,44,46 Here, we demonstrated for the first time that Panx1 is present in the SSM of murine cardiomyocytes ( Figure 4A and E). ...
... 44 Indeed, the previous proteomic analysis identified cytoskeletal proteins as well as mitochondrial proteins like VDAC1, ANT2, and heat shock protein 70 (Hsc70) as Panx1 partners. 34,44,46 Here, we demonstrated for the first time that Panx1 is present in the SSM of murine cardiomyocytes ( Figure 4A and E). Altered mitochondrial morphology has been reported in the setting of cardiac I/R (for a review, see Ref. 47 ) where prevention of mitochondrial fission protects the myocardium and promotes recovery. ...
Article
Full-text available
Aims: No effective therapy is available in clinics to protect the heart from ischaemia/reperfusion (I/R) injury. Endothelial cells are activated after I/R, which may drive the inflammatory response by releasing ATP through pannexin1 (Panx1) channels. Here, we investigated the role of Panx1 in cardiac I/R. Methods and results: Panx1 was found in cardiac endothelial cells, neutrophils, and cardiomyocytes. After in vivo I/R, serum Troponin-I, and infarct size were less pronounced in Panx1-/- mice, but leukocyte infiltration in the infarct area was similar between Panx1-/- and wild-type mice. Serum Troponin-I and infarct size were not different between mice with neutrophil-specific deletion of Panx1 and Panx1fl/fl mice, suggesting that cardioprotection by Panx1 deletion rather involved cardiomyocytes than the inflammatory response. Physiological cardiac function in wild-type and Panx1-/- hearts was similar. The time to onset of contracture and time to maximal contracture were delayed in Panx1-/- hearts, suggesting reduced sensitivity of these hearts to ischaemic injury. Moreover, Panx1-/- hearts showed better recovery of left ventricle developed pressure, cardiac contractility, and relaxation after I/R. Ischaemic preconditioning failed to confer further protection in Panx1-/- hearts. Panx1 was found in subsarcolemmal mitochondria (SSM). SSM in WT or Panx1-/- hearts showed no differences in morphology. The function of the mitochondrial permeability transition pore and production of reactive oxygen species in SSM was not affected, but mitochondrial respiration was reduced in Panx1-/- SSM. Finally, Panx1-/- cardiomyocytes had a decreased mitochondrial membrane potential and an increased mitochondrial ATP content. Conclusion: Panx1-/- mice display decreased sensitivity to cardiac I/R injury, resulting in smaller infarcts and improved recovery of left ventricular function. This cardioprotective effect of Panx1 deletion seems to involve cardiac mitochondria rather than a reduced inflammatory response. Thus, Panx1 may represent a new target for controlling cardiac reperfusion damage.
... Few studies have examined determinants of Panx1 delivery to the cell membrane and its residence there, although they were shown to be regulated in part by interaction with actin microfilaments [43]. In order to address the distribution of Panx1 and the dynamic organization in specific membrane surface microdomains, their activity and dependence on cholesterol in the cell plasma membrane, we have examined the effects of cholesterol depletion on cell surface mobility of Panx1 and activity of the channels that it forms. ...
... Recovery kinetics and mobile fraction were comparable to those seen for mobility of other proteins of similar size in a non-junctional membrane (e.g., occludin and Cx30 connexons [51]). Previous FRAP studies performed at room temperature on GFP-tagged mouse Panx1 expressed in a breast cancer cell line (BICR-MIRk) [43] and zebrafish Panx1a expressed in N2a cells [52] revealed diffusion kinetics comparable to those reported here, with recovery of 45-60% at one min after bleaching. ...
... Recovery kinetics and mobile fraction were comparable to those seen for mobility of other proteins of similar size in a nonjunctional membrane (e.g., occludin and Cx30 connexons [51]). Previous FRAP studies performed at room temperature on GFP-tagged mouse Panx1 expressed in a breast cancer cell line (BICR-MIRk) [43] and zebrafish Panx1a expressed in N2a cells [52] revealed diffusion kinetics comparable to those reported here, with recovery of 45-60% at one min after bleaching. ...
Article
Full-text available
Pannexin1 (Panx1) is expressed in both neurons and glia where it forms ATP-permeable channels that are activated under pathological conditions such as epilepsy, migraine, inflammation, and ischemia. Membrane lipid composition affects proper distribution and function of receptors and ion channels, and defects in cholesterol metabolism are associated with neurological diseases. In order to understand the impact of membrane cholesterol on the distribution and function of Panx1 in neural cells, we used fluorescence recovery after photobleaching (FRAP) to evaluate its mobility and electrophysiology and dye uptake to assess channel function. We observed that cholesterol extraction (using methyl-β-cyclodextrin) and inhibition of its synthesis (lovastatin) decreased the lateral diffusion of Panx1 in the plasma membrane. Panx1 channel activity (dye uptake, ATP release and ionic current) was enhanced in cholesterol-depleted Panx1 transfected cells and in wild-type astrocytes compared to non-depleted or Panx1 null cells. Manipulation of cholesterol levels may, therefore, offer a novel strategy by which Panx1 channel activation might modulate various pathological conditions.
... Tolvaptan plays a pivotal role as a vasopressin V2 receptor inhibitor and is a key factor in the regulation of water and salt reabsorption in the kidneys (Berl, 2015). Tolvaptan is a vasopressin receptor antagonist that specifically targets V2 receptors and disrupts the reabsorption of free water (Torres et al., 2012;Tamma et al., 2017). ...
... Pannexin-1 channel activity and probenecid: Pannexin-1 plays a crucial role in the formation of large-pore membrane channels that facilitate the passage of ions and metabolites and promote ATP release (Chiu et al., 2018;Whyte-Fagundes and Zoidl, 2018;Wei et al., 2021). This channel not only interacts with various cytoskeletal proteins and influences cell polarity but also contributes to microtubule stability (Silverman et al., 2008;Xu et al., 2018b) Additionally, its interaction with actin and factors regulating cell surface localization and mobility are vital for maintaining normal cellular functions (Bhalla-Gehi et al., 2010;Wicki-Stordeur and Swayne, 2013). The preferential concentration of pannexin-1 channels in the apical membrane domain of polarized cells, particularly monolayer sheets or spheroids, facilitates diverse cellular actions. ...
Article
Full-text available
Autosomal dominant polycystic kidney disease (ADPKD), a congenital genetic disorder, is a notable contributor to the prevalence of chronic kidney disease worldwide. Despite the absence of a complete cure, ongoing research aims for early diagnosis and treatment. Although agents such as tolvaptan and mTOR inhibitors have been utilized, their effectiveness in managing the disease during its initial phase has certain limitations. This review aimed to explore new targets for the early diagnosis and treatment of ADPKD, considering ongoing developments. We particularly focus on cell polarity, which is a key factor that influences the process and pace of cyst formation. In addition, we aimed to identify agents or treatments that can prevent or impede the progression of renal fibrosis, ultimately slowing its trajectory toward end-stage renal disease. Recent advances in slowing ADPKD progression have been examined, and potential therapeutic approaches targeting multiple pathways have been introduced. This comprehensive review discusses innovative strategies to address the challenges of ADPKD and provides valuable insights into potential avenues for its prevention and treatment.
... The local ATP increase will then activate the ionotropic or the metabotropic purinergic receptors responsible for, respectively, Ca 2+ entry or Ca 2+ release from the intracellular stores [121]. Of note, the C-terminus of Panx1 has been shown to interact directly with F-actin, actin-related protein 3 (Arp3), and collapsin response mediator protein 2 (Crmp2), regulators of microtubule polymerization and stabilization in neural cells [122][123][124]. Thus, the crosstalk between pannexins, purinergic signaling, Ca 2+ regulation, and cytoskeleton remodeling suggests a potential role of pannexons in several steps of the epithelium regeneration. ...
... Another fundamental question is to discriminate between the channel-dependent or channel-independent role of connexins and pannexins in repairing airway epithelial cells. In this context, a particular focus must be placed on their localization, as connexins and pannexins can be found in intracellular compartments [122,144,145]. The intracellular relocalization of connexins and pannexins depends on the regulation of their trafficking and degradation, which are controlled by phosphorylation, acetylation, glycosylation, and ubiquitination [12,146]. ...
Article
Full-text available
Connexins and pannexins are transmembrane proteins that can form direct (gap junctions) or indirect (connexons, pannexons) intercellular communication channels. By propagating ions, metabolites, sugars, nucleotides, miRNAs, and/or second messengers, they participate in a variety of physiological functions, such as tissue homeostasis and host defense. There is solid evidence supporting a role for intercellular signaling in various pulmonary inflammatory diseases where alteration of connexin/pannexin channel functional expression occurs, thus leading to abnormal intercellular communication pathways and contributing to pathophysiological aspects, such as innate immune defense and remodeling. The integrity of the airway epithelium, which is the first line of defense against invading microbes, is established and maintained by a repair mechanism that involves processes such as proliferation, migration, and differentiation. Here, we briefly summarize current knowledge on the contribution of connexins and pannexins to necessary processes of tissue repair and speculate on their possible involvement in the shaping of the airway epithelium integrity.
... In support of a potential regulatory role of Panx1 on the actin cytoskeleton, it has been previously reported that Panx1 channels influence cellular changes that require cytoskeletal modifications, including migration, differentiation, and proliferation [33,97]. Interestingly, Panx1 directly interacts with F-actin and its regulator Arp3 [33,98], which is the main component of the Arp2/3 complex involved in the nucleation and branching of F-actin [99]; thus, suggesting that F-actin polymerization and nucleation of new microfilaments might directly control Panx1 localization. In this context, a higher spine density and complexity network has been reported in global and conditional Panx1-KO [32] compared to control animals, suggesting that Panx1 activity limits the neuronal actin cytoskeleton remodeling. ...
Article
Full-text available
Enhanced activity and overexpression of Pannexin 1 (Panx1) channels contribute to neuronal pathologies such as epilepsy and Alzheimer’s disease (AD). The Panx1 channel ablation alters the hippocampus’s glutamatergic neurotransmission, synaptic plasticity, and memory flexibility. Nevertheless, Panx1-knockout (Panx1-KO) mice still retain the ability to learn, suggesting that compensatory mechanisms stabilize their neuronal activity. Here, we show that the absence of Panx1 in the adult brain promotes a series of structural and functional modifications in the Panx1-KO hippocampal synapses, preserving spontaneous activity. Compared to the wild-type (WT) condition, the adult hippocampal neurons of Panx1-KO mice exhibit enhanced excitability, a more complex dendritic branching, enhanced spine maturation, and an increased proportion of multiple synaptic contacts. These modifications seem to rely on the actin–cytoskeleton dynamics as an increase in the actin polymerization and an imbalance between the Rac1 and the RhoA GTPase activities were observed in Panx1-KO brain tissues. Our findings highlight a novel interaction between Panx1 channels, actin, and Rho GTPases, which appear to be relevant for synapse stability.
... In addition, mechanical forces between the actin cytoskeleton and substrate are critical for successful migration. Pannexin-1 is a mechanosensitive ion channel that binds to F-actin in the carboxy terminus (Bao et al., 2004;Bhalla-Gehi et al., 2010). Inhibition of pannexin-1 in corneas from 12-week mice with 10PanX caused decreases in calcium signaling after injury and alterations at the leading edge with a reduction in lamellipodia (Rhodes et al., 2021). ...
Article
Full-text available
The cornea is exposed daily to a number of mechanical stresses including shear stress from tear film and blinking. Over time, these stressors can lead to changes in the extracellular matrix that alter corneal stiffness, cell-substrate structures, and the integrity of cell-cell junctions. We hypothesized that changes in tissue stiffness of the cornea with age may alter calcium signaling between cells after injury, and the downstream effects of this signaling on cellular motility and wound healing. Nanoindentation studies revealed that there were significant differences in the stiffness of the corneal epithelium and stroma between corneas of 9- and 27-week mice. These changes corresponded to differences in the timeline of wound healing and in cell signaling. Corneas from 9-week mice were fully healed within 24 h. However, the wounds on corneas from 27-week mice remained incompletely healed. Furthermore, in the 27-week cohort there was no detectable calcium signaling at the wound in either apical or basal corneal epithelial cells. This is in contrast to the young cohort, where there was elevated basal cell activity relative to background levels. Cell culture experiments were performed to assess the roles of P2Y2, P2X7, and pannexin-1 in cellular motility during wound healing. Inhibition of P2Y2, P2X7, or pannexin-1 all significantly reduce wound closure. However, the inhibitors all have different effects on the trajectories of individual migrating cells. Together, these findings suggest that there are several significant differences in the stiffness and signaling that underlie the decreased wound healing efficacy of the cornea in older mice.
... The alternative mechanism is that the PANX channel is gated via the cytoskeleton or extracellular matrix, which interacts with the PANX channel ( Figure 1B). A study using co-immunoprecipitation and co-sedimentation assays has shown that PANX1 interacts with actin through its C-terminus [58]. Using liquid chromatography and tandem mass spectrometry (LC/MS), both actin and actin-related protein 3 (Arp3) have been identified to bind PANX1 directly [59]. ...
Article
Full-text available
Pannexins (PANX) were cloned based on their sequence homology to innexins (Inx), invertebrate gap junction proteins. Although there is no sequence homology between PANX and connexins (Cx), these proteins exhibit similar configurations. The PANX family has three members, PANX1, PANX2 and PANX3. Among them, PANX1 has been the most extensively studied. The PANX1 channels are activated by many factors, including high extracellular K+ ([K+]e), high intracellular Ca2+ ([Ca2+]i), Src family kinase (SFK)-mediated phosphorylation, caspase cleavage and mechanical stimuli. However, the mechanisms mediating this mechanosensitivity of PANX1 remain unknown. Both force-from-lipids and force-from-filaments models are proposed to explain the gating mechanisms of PANX1 channel mechanosensitivity. Finally, both the physiological and pathological roles of mechanosensitive PANX1 are discussed.
Article
Chronic pain complicates many diseases and is notoriously difficult to treat. In search of new therapeutic targets, pannexin-1 (Panx1) channels have sparked intense interest as a key mechanism involved in a variety of chronic pain conditions. Panx1 channels are transmembrane proteins that release ions and small molecules, such as adenosine triphosphate (ATP). They are expressed along important nodes of the pain pathway, modulating activity of diverse cell types implicated in the development and progression of chronic pain caused by injury or pathology. This review highlights advances that have unlocked the core structure and machinery controlling Panx1 function with a focus on understanding and treating chronic pain.
Article
The skin is a highly organized tissue composed of multiple layers and cell types that require coordinated cell to cell communication to maintain tissue homeostasis. In skin cancer, this organized structure and communication is disrupted, prompting the malignant transformation of healthy cells into melanoma, basal cell carcinoma or squamous cell carcinoma tumours. One such family of channel proteins critical for cellular communication is pannexins (PANX1, PANX2, PANX3), all of which are present in the skin. These heptameric single-membrane channels act as conduits for small molecules and ions like ATP and Ca2+ but have also been shown to have channel-independent functions through their interacting partners or action in signalling pathways. Pannexins have diverse roles in the skin such as in skin development, aging, barrier function, keratinocyte differentiation, inflammation, and wound healing, which were discovered through work with pannexin knockout mice, organotypic epidermis models, primary cells, and immortalized cell lines. In the context of cutaneous cancer, PANX1 is present at high levels in melanoma tumours and functions in melanoma carcinogenesis, and both PANX1 and PANX3 expression is altered in non-melanoma skin cancer. PANX2 has thus far not been implicated in any skin cancer. This review will discuss pannexin isoforms, structure, trafficking, post-translational modifications, interactome, and channel activity. We will also outline the expression, localization, and function of pannexin channels within the diverse cell types of the epidermis, dermis, hypodermis, and adnexal structures of the skin, and how these properties are exploited or abrogated in instances of skin cancer.
Article
Full-text available
Pannexin1 proteins form communication channels at the cell plasma membrane surface, which allow the transfer of small molecules and ions between the intracellular compartment and extracellular environment. In this way, pannexin1 channels play an important role in various cellular processes and diseases. Indeed, a plethora of human pathologies is associated with the activation of pannexin1 channels. The present paper reviews and summarizes the structure, life cycle, regulation and (patho)physiological roles of pannexin1 channels, with a particular focus on the relevance of pannexin1 channels in liver diseases.
Article
Full-text available
The ATP-sensitive potassium (K(ATP)) channel controls insulin secretion by coupling glucose metabolism to excitability of the pancreatic beta-cell membrane. The channel comprises four subunits each of Kir6.2 and the sulphonylurea receptor (SUR1), encoded by KCNJ11 and ABCC8, respectively. Mutations in these genes that result in reduced activity or expression of K(ATP) channels lead to enhanced beta-cell excitability, insulin hypersecretion and hypoglycaemia, and in humans lead to the clinical condition congenital hyperinsulinism (CHI). Here we have investigated the molecular basis of the focal form of CHI caused by one such mutation in Kir6.2, E282K. The study led to the discovery that Kir6.2 contains a di-acidic ER exit signal, (280)DLE(282), which promotes concentration of the channel into COPII-enriched ER exit sites prior to ER export via a process that requires Sar1-GTPase. The E282K mutation abrogates the exit signal, and thereby prevents the ER export and surface expression of the channel. When co-expressed, the mutant subunit was able to associate with the wild-type Kir6.2 and form functional channels. Thus unlike most mutations, the E282K mutation does not cause protein mis-folding. Since in focal CHI, maternal chromosome containing the K(ATP) channel genes is lost, beta-cells of the patient would lack wild-type Kir6.2 to rescue the mutant Kir6.2 subunit expressed from the paternal chromosome. The resultant absence of functional K(ATP) channels leads to insulin hypersecretion. Taken together, we conclude that surface expression of K(ATP) channels is critically dependent on the Sar1-GTPase-dependent ER exit mechanism and abrogation of the di-acidic ER exit signal leads to CHI.
Article
Full-text available
Gap junctions (GJ), specialised membrane structures that mediate cell-to-cell communication in almost all animal tissues, are composed of intercellular channel-forming integral membrane proteins termed connexins (Cxs), innexins or pannexins. The activity of these channels is closely regulated, particularly by intramolecular modifications as phosphorylation of proteins, via the formation of multiprotein complexes where pore-forming subunits bind to auxiliary channel subunits and associate with scaffolding proteins that play essential roles in channel localization and activity. Scaffolding proteins link signalling enzymes, substrates, and potential effectors (such as channels) into multiprotein signalling complexes that may be anchored to the cytoskeleton. Protein-protein interactions play essential roles in channel localization and activity and, besides their cell-to-cell channel-forming functions, gap junctional proteins now appear involved in different cellular functions (e.g. transcriptional and cytoskeletal regulation). The present review summarizes the recent progress regarding the proteins capable of interacting with junctional proteins and their functional importance.
Article
Full-text available
Pannexins have been proposed to play a role in gap junctional intercellular communication and as single-membrane channels, although many of their molecular characteristics differ from connexins. Localization of untagged Panx1 and Panx3 exogenously expressed in five cultured cell lines revealed a cell surface distribution profile with limited evidence of cell surface clustering and variable levels of intracellular pools. However, N-glycosylation-defective mutants of pannexins exhibited a more prominent intracellular distribution with decreased cell surface labeling, suggesting an important role for pannexin glycosylation in trafficking. Similar to wild-type pannexins, the glycosylation-defective mutants failed to noticeably transfer microinjected fluorescent dyes to neighboring cells, suggesting that few, or no functional intercellular channels were formed. Finally, varied distribution patterns of endogenous Panx1 and Panx3 were observed in cells of osteoblast origin and Madin-Darby canine kidney cells. Collectively, diverse expression and distribution profiles of Panx1 and Panx3 suggest that they may have multiple cellular functions.
Article
Full-text available
Pannexins are mammalian orthologs of innexins and have a predicted topological folding pattern similar to that of connexins, except they are glycosylated. Rat pannexin 1 is glycosylated at N254 and this residue is important for plasma membrane targeting. Here we demonstrate that cell surface expression levels of the rat pannexin 1 N254Q mutant are rescued by coexpression with the wild-type protein. In paired Xenopus oocytes, the functional effect of this rescue is inconsequential; however, cell surface deglycosylation by PNGase F significantly enhanced functional gap junction formation. In mammalian cells, wild-type oligomers traffic at a slower rate than Myc-or tetracysteine domain-tagged versions, a behavior opposite to that of tagged connexins. The temporal differences of Panx1 trafficking correlate with spatial differences of intracellular localizations induced by Golgi blockage by Brefeldin-A or glycosylation prevention by tunicamycin. Therefore, Panx1 has kinetics and dynamics that make it unique to serve distinct functions separate from connexin-based channels.
Article
Full-text available
In this review, we briefly summarize what is known about the properties of the three families of gap junction proteins, connexins, innexins and pannexins, emphasizing their importance as intercellular channels that provide ionic and metabolic coupling and as non-junctional channels that can function as a paracrine signaling pathway. We discuss that two distinct groups of proteins form gap junctions in deuterostomes (connexins) and protostomes (innexins), and that channels formed of the deuterostome homologues of innexins (pannexins) differ from connexin channels in terms of important structural features and activation properties. These differences indicate that the two families of gap junction proteins serve distinct, complementary functions in deuterostomes. In several tissues, including the CNS, both connexins and pannexins are involved in intercellular communication, but have different roles. Connexins mainly contribute by forming the intercellular gap junction channels, which provide for junctional coupling and define the communication compartments in the CNS. We also provide new data supporting the concept that pannexins form the non-junctional channels that play paracrine roles by releasing ATP and, thus, modulating the range of the intercellular Ca(2+)-wave transmission between astrocytes in culture.
Article
Full-text available
The pannexin family of mammalian proteins, composed of Panx1, Panx2, and Panx3, has been postulated to be a new class of single-membrane channels with functional similarities to connexin gap junction proteins. In this study, immunolabeling and coimmunoprecipitation assays revealed that Panx1 can interact with Panx2 and to a lesser extent, with Panx3 in a glycosylation-dependent manner. Panx2 strongly interacts with the core and high-mannose species of Panx1 but not with Panx3. Biotinylation and dye uptake assays indicated that all three pannexins, as well as the N-glycosylation-defective mutants of Panx1 and Panx3, can traffic to the cell surface and form functional single-membrane channels. Interestingly, Panx2, which is also a glycoprotein and seems to only be glycosylated to a high-mannose form, is more abundant in intracellular compartments, except when coexpressed with Panx1, when its cell surface distribution increases by twofold. Functional assays indicated that the combination of Panx1 and Panx2 results in compromised channel function, whereas coexpressing Panx1 and Panx3 does not affect the incidence of dye uptake in 293T cells. Collectively, these results reveal that the functional state and cellular distribution of mouse pannexins are regulated by their glycosylation status and interactions among pannexin family members.
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
Voltage-dependent sodium channels are distributed nonuniformly over the surface of nerve cells and are localized to morphologically distinct regions. Fluorescent neurotoxin probes specific for the voltage-dependent sodium channel stain the axon hillock 5-10 times more intensely than the cell body and show punctate fluorescence confined to the axon hillock which can be compared with the more diffuse and uniform labeling in the cell body. Using fluorescence photobleaching recovery (FPR) we measured the lateral mobility of voltage-dependent sodium channels over specific regions of the neuron. Nearly all sodium channels labeled with specific neurotoxins are free to diffuse within the cell body with lateral diffusion coefficients on the order of 10(-9) cm2/s. In contrast, lateral diffusion of sodium channels in the axon hillock is restricted, apparently in two different ways. Not only do sodium channels in these regions diffuse more slowly (10(-10)-10(-11) cm2/s), but also they are prevented from diffusing between axon hillock and cell body. No regionalization or differential mobilities were observed, however, for either tetramethylrhodamine-phosphatidylethanolamine, a probe of lipid diffusion, or FITC-succinyl concanavalin A, a probe for glycoproteins. During the maturation of the neuron, the plasma membrane differentiates and segregates voltage-dependent sodium channels into local compartments and maintains this localization perhaps either by direct cytoskeletal attachments or by a selective barrier to channel diffusion.
Article
Gap junctions in the heart play an important functional role by electrically coupling cells, thereby organizing the pattern of current flow to allow co-ordinated muscle contraction. Cardiac gap junctions are therefore intimately involved in normal conduction as well as the genesis of potentially lethal arrhythmias. We recently utilized electron cryo-microscopy and image analysis to examine frozen–hydrated 2D crystals of a recombinant, C-terminal truncated form of connexin43 (Cx43; 1), the principal cardiac gap junction protein. The projection map at 7 Å resolution revealed that each 30 kDa connexin subunit has a transmembrane -helix that lines the aqueous pore and a second -helix in close contact with the membrane lipids. The distribution of densities allowed us to propose a model in which the two apposing connexons that form the channel are staggered by ∼30°. We are now recording images of tilted, frozen–hydrated 2D crystals, and a preliminary 3D map has been computed at an in-plane resolution of ∼7.5 Å and a vertical resolution of ∼25 Å. As predicted by our model, the two apposing connexons that form the channel are staggered with respect to each other for certain connexin molecular boundaries within the hexamer. Within the membrane interior each connexin subunit displays four rods of density, which are consistent with an -helical conformation for the four transmembrane domains. Preliminary studies of BHK hamster cells that express the truncated Cx43 designated 1Cx263T demonstrate that oleamide, a sleep inducing lipid, blocks in vivo dye transfer, suggesting that oleamide causes closure of 1Cx263T channels. The comparison of the 3D structures in the presence and absence of oleamide may provide an opportunity to explore the conformational changes that are associated with oleamide-induced blockage of dye transfer. The structural details revealed by our analysis will be essential for delineating the molecular basis for intercellular current flow in the heart, as well as the general molecular design and functional properties of this important class of channel proteins.
Article
For all animals, cell migration is an essential and highly regulated process. Cells migrate to shape tissues, to vascularize tissues, in wound healing, and as part of the immune response. Unfortunately, tumor cells can also become migratory and invade surrounding tissues. Some cells migrate as individuals, but many cell types will, under physiological conditions, migrate collectively in tightly or loosely associated groups. This includes invasive tumor cells. This review discusses different types of collective cell migration, including sheet movement, sprouting and branching, streams, and free groups, and highlights recent findings that provide insight into cells' organization and behavior. Cells performing collective migration share many cell biological characteristics with independently migrating cells but, by affecting one another mechanically and via signaling, these cell groups are subject to additional regulation and constraints. New properties that emerge from this connectivity can contribute to shaping, guiding, and ultimately ensuring tissue function.