ArticlePDF Available

Polycystin-1, the PKD1 gene product, is in a complex containing E- cadherin and the catenins

Authors:

Abstract and Figures

Autosomal dominant polycystic kidney disease (ADPKD) is a common human genetic disease characterized by cyst formation in kidney tubules and other ductular epithelia. Cells lining the cysts have abnormalities in cell proliferation and cell polarity. The majority of ADPKD cases are caused by mutations in the PKD1 gene, which codes for polycystin-1, a large integral membrane protein of unknown function that is expressed on the plasma membrane of renal tubular epithelial cells in fetal kidneys. Because signaling from cell-cell and cell-matrix adhesion complexes regulates cell proliferation and polarity, we speculated that polycystin-1 might interact with these complexes. We show here that polycystin-1 colocalized with the cell adhesion molecules E-cadherin and alpha-, beta-, and gamma-catenin. Polycystin-1 coprecipitated with these proteins and comigrated with them on sucrose density gradients, but it did not colocalize, coprecipitate, or comigrate with focal adhesion kinase, a component of the focal adhesion. We conclude that polycystin-1 is in a complex containing E-cadherin and alpha-, beta-, and gamma-catenin. These observations raise the question of whether the defects in cell proliferation and cell polarity observed in ADPKD are mediated by E-cadherin or the catenins.
Content may be subject to copyright.
Introduction
Autosomal dominant polycystic kidney disease (ADPKD)
is one of the most common genetic diseases in humans. It
is a multisystem disease that causes renal cysts and renal
failure as well as cysts in the liver, pancreas, and other
organs, aneurysms of the cerebral and other arteries, car-
diac valvular insufficiency, and colonic diverticula (1).
ADPKD is caused by mutations in at least 3 genes. Two
of these, PKD1 and PKD2, have been cloned and
sequenced (2–5). Families with ADPKD unlinked to
PKD1 or PKD2 have been identified, so at least 1 more
locus must exist (6–9). Both PKD1 and PKD2 code for
novel molecules whose biochemical functions are
unknown. The PKD1 gene product, polycystin-1, is a large
intrinsic membrane protein with a molecular weight of
approximately 485 kDa that is expressed in a develop-
mentally regulated manner. The amino terminal domain
of polycystin-1 is predicted to be extracytoplasmic. The
carboxyl terminal domain of polycystin-1 is predicted to
have between 7 and 11 transmembrane helices. Poly-
cystin-2, the PKD2 gene product, is an intrinsic mem-
brane protein with a molecular weight of 110 kDa. It is
predicted to have intracellular amino and carboxyl ter-
mini and has significant similarities to transient receptor
potential (Trp) channel subunits. The observation that
carboxyl terminal fragments of polycystin-1 and -2 inter-
act in vitro and in transfected cells, coupled with the sim-
ilarity of the disease caused by mutations in either PKD1
or PKD2, suggests that these molecules act together in an
uncharacterized biochemical pathway (10, 11).
The primary structure of polycystin-1 suggests that it
might be a receptor. The large extracellular domain
begins with 2 leucine-rich repeats and has 16 novel
repeats with an Ig-like fold, an LDL-A domain, a calci-
um-dependent lectin domain, and a domain of nearly
1,000 amino acids that is similar to the sea urchin recep-
tor for egg jelly. These domains mediate protein-protein
and protein-carbohydrate interactions in other proteins,
and by analogy, are likely to mediate similar interactions
in polycystin-1. We and others have shown that poly-
cystin-1 is expressed in the plasma membrane of epithe-
lial cells (12–17). This observation is consistent with the
hypothesis that polycystin-1 might interact with an
extracellular ligand or ligands.
We had observed that polycystin-1 was concentrated
on the apical and lateral aspects of epithelial cells in
developing kidneys (12). This localization pattern was
similar to that of the cell adhesion molecule E-cadherin
in developing kidneys, and suggested that polycystin-1
might interact with E-cadherin. This hypothesis was par-
ticularly intriguing because of the roles that E-cadherin
plays in orchestrating cell proliferation and polarization
of epithelial monolayers. E-cadherin modulates prolif-
eration and polarity through associated cytoplasmic
polypeptides, the catenins (18–20). E-cadherin regulates
cell proliferation and gene expression by participating in
the regulation of β-catenin metabolism (21, 22). Cyst-lin-
ing epithelial cells in ADPKD express markers for cell
proliferation in vivo and are abnormally sensitive to
growth factors in vitro, suggesting that there is a
derangement of the control of cell growth in ADPKD
(23). E-cadherin orchestrates cell polarity by directing a
cytoskeletal network containing ankyrin and spectrin to
the basolateral aspect of the epithelial cell (24, 25). Na,K-
ATPase binds to ankyrin and is retained in the basolat-
eral membrane by its interaction with the ankyrin/spec-
trin cytoskeleton (26, 27). When these complexes are
disrupted, for example by ischemia, Na,K-ATPase
The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10 1459
Polycystin-1, the
PKD1
gene product, is in a complex
containing E-cadherin and the catenins
Yonghong Huan and Janet van Adelsberg
Department of Medicine, Columbia University, New York, New York 10032, USA
Address correspondence to: Janet van Adelsberg, Department of Medicine, Columbia University, 630 West 168th Street, Box 84,
New York, New York 10032, USA. Phone: (212) 305-4476; Fax: (212) 305-3475; E-mail: jsv1@columbia.edu.
Received for publication September 2, 1998, and accepted in revised form October 5, 1999.
Autosomal dominant polycystic kidney disease (ADPKD) is a common human genetic disease characterized
by cyst formation in kidney tubules and other ductular epithelia. Cells lining the cysts have abnormalities in
cell proliferation and cell polarity. The majority of ADPKD cases are caused by mutations in the PKD1 gene,
which codes for polycystin-1, a large integral membrane protein of unknown function that is expressed on
the plasma membrane of renal tubular epithelial cells in fetal kidneys. Because signaling from cell-cell and
cell-matrix adhesion complexes regulates cell proliferation and polarity, we speculated that polycystin-1
might interact with these complexes. We show here that polycystin-1 colocalized with the cell adhesion mol-
ecules E-cadherin and α-, β-, and γ-catenin. Polycystin-1 coprecipitated with these proteins and comigrated
with them on sucrose density gradients, but it did not colocalize, coprecipitate, or comigrate with focal adhe-
sion kinase, a component of the focal adhesion. We conclude that polycystin-1 is in a complex containing
E-cadherin and α-, β-, and γ-catenin. These observations raise the question of whether the defects in cell pro-
liferation and cell polarity observed in ADPKD are mediated by E-cadherin or the catenins.
J. Clin. Invest. 104:1459–1468 (1999).
becomes depolarized and appears in both lateral and api-
cal plasma membranes (28, 29). Some, but not all, inves-
tigators have found that Na,K-ATPase is depolarized in
ADPKD epithelia (30–32). These observations suggest-
ed that polycystin-1 might modulate cell proliferation
and cell polarity by interacting with E-cadherin or the
catenins. We found that polycystin-1 is part of a complex
containing E-cadherin and the catenins, suggesting that
some of the defects observed in ADPKD might be medi-
ated by E-cadherin or the catenins.
Methods
Antibodies. Rabbit polyclonal antibody against polycystin
peptide B145 has been described previously (12). Mouse
monoclonal antibodies against human γ-catenin were
obtained from Zymed Laboratories (South San Francis-
co, California, USA) and from Transduction Laborato-
ries (Lexington, Kentucky, USA). Mouse monoclonal
antibodies against human β1-integrin was obtained from
GIBCO BRL (Gaithersburg, Maryland, USA) and Trans-
duction Laboratories. Monoclonal antibodies against
human E-cadherin, α- and β-catenin, and focal adhesion
kinase (FAK) were obtained from Transduction Labora-
tories. Cy3 and horseradish peroxidase–labeled second-
ary antibodies absorbed against human, rat, mouse, and
rabbit serum proteins were obtained from Jackson
ImmunoResearch Laboratories Inc. (West Grove, Penn-
sylvania, USA). Oregon Green 488–labeled secondary
antibodies were obtained from Molecular Probes Inc.
(Eugene, Oregon, USA).
Immunofluorescence. Indirect immunofluorescence on
frozen sections was performed as described previously
(12). Briefly, the sections were fixed in methanol and ace-
tone (1:1, vol/vol) at –20°C for 10 minutes. The sections
were air dried, rehydrated in PBS, and blocked for 15–30
minutes in 0.2% (wt/vol) BSA in PBS. Primary antibod-
ies were diluted in blocking buffer containing the fol-
lowing: polycystin-1, either 1:10 (affinity purified) or
1:50 (anti-serum); E-cadherin, 1:25; α-catenin, 1:5; β-
catenin, 1:50; γ-catenin (plakoglobin), 1:100; FAK, 1:5;
and β1-integrin, 1:5. Human fetal kidneys of 20–24 weeks
gestational age were obtained under a protocol approved
by the Institutional Review Board of Columbia Univer-
sity. The sections were postfixed in 4% paraformaldehyde
before being mounted for confocal microscopy in Pro-
long Antifade (Molecular Probes Inc.).
Confocal microscopy. Confocal microscopy was per-
formed with a Zeiss LSM 410 laser scanning on a Zeiss
Axiovert 100 microscope platform using an argon-kryp-
ton laser (Carl Zeiss North America, Thornwood, New
York, USA). The 488-nm and 568-nm laser lines were
used for Oregon Green and Cy3 detection, respectively.
Images were collected with a 515–540-nm band pass fil-
ter (green channel) and a 590-nm long pass filter (red
channel). Sections were scanned at low power on the red
channel (polycystin-1 staining) to select tubules strong-
ly stained for polycystin. These tubules were then exam-
ined at high power with both red and green channels.
Red and green images were collected simultaneously and
stored for analysis as both individual red or green images
and as the dual-scanned image. For each image, tissue
sections were viewed at ×100 or ×40. The images (512 ×
512 pixels) were projected at 600 pixels/inch using
Adobe Photoshop 4.0 (Adobe Systems Inc., Mountain
View, California, USA), which was also used to make the
montages shown in Figures 1 and 5.
Cell lines. The human pancreatic adenocarcinoma (HPAC)
cell line was obtained from the American Type Culture Col-
lection (Rockville, Maryland, USA). Cells were grown in a
1:1 mixture of DMEM and Ham’s F12 medium containing
1.2 g/L sodium bicarbonate, 15 mM HEPES, 0.002 mg/mL
insulin, 0.005 mg/mL transferrin, 40 ng/mL hydrocorti-
sone, and 10 ng/mL EGF, with 5% FBS (HyClone Labora-
tories, Logan, Utah, USA). Insulin, transferrin, and EGF
were purchased from Fisher Scientific (Atlanta, Georgia,
USA). DMEM, F12, and HEPES were purchased from
GIBCO BRL. Hydrocortisone was obtained from Sigma
Chemical Co. (St. Louis, Missouri, USA).
RT-PCR. Total RNA was prepared from cultured cells
using RNAzol B (Tel-Test Inc., Friendswood, Texas,
USA). Primers for the 3end of the PKD1 gene were
selected using the Prime program (Wisconsin Package,
version 8.0) from Genetics Computer Group Inc. (Madi-
son, Wisconsin, USA). The forward primer (5to 3)
sequence was GGC TGT TAT TCT CCG CTG
(nucleotides 12,441–12,458, HSU24497). The reverse
primer (5to 3) sequence was GGG TGG ACC TTG TTC
TTG (nucleotides 13,034–13,017). RT-PCR was carried
out using the Access RT-PCR System from Promega
Corp. (Madison, Wisconsin, USA), according to the man-
ufacturer’s instructions, except that 10% DMSO was
included in the reaction. Human β-actin primers from 2
exons (Stratagene, La Jolla, California, USA) were used as
a positive control; RNA treated with RNase A (Sigma
Chemical Co.) was used as a negative control. RT-PCR
products were separated on a 3% NuSieve 3:1 agarose gel
(FMC BioProducts, Rockland, Maine, USA) in 1×Tris-
acetate-EDTA buffer, and were purified using QIAEX
resin (QIAGEN Inc., Valencia, California, USA) for direct
sequencing at the DNA facility at Columbia University.
Immunoprecipitation and immunoblotting. HPAC cells
were labeled overnight with [35S]methionine/cysteine
(EXPRESS label; NEN Life Science Products, Inc.,
Boston, Massachusetts, USA) at 0.1 mCi/mL in methio-
nine/cysteine–free medium with dialyzed FBS. Cells were
suspended in lysis buffer containing 1% (vol/vol) Triton
X-100, 0.5% (vol/vol) NP-40, 150 mM NaCl, 10 mM Tris-
HCl (pH 7.4), 1 mM EDTA, 1 mM EGTA, 0.2 mM
Na3VO4, 1 mM PMSF, 2 µg/mL aprotinin, 1 µg/mL E-64,
10 µg/mL leupeptin, and 1 µg/mL pepstatin A. Cells
were lysed by passage through a 22-gauge needle several
times. Insoluble material was removed by centrifugation
at 10,000 gfor 30 minutes at 4°C. Lysates were pre-
cleared by incubation with preimmune rabbit serum and
protein A–agarose (Oncogene Science Inc., Uniondale,
New York, USA) for 30 minutes at 4°C. After removal of
protein A–agarose by centrifugation, lysates were incu-
bated for 1 hour at 4°C with equivalent concentrations
of protein A–purified IgG from preimmune serum,
immune serum from rabbits immunized with peptide
B145, or immune serum plus peptide B145 at 0.4
mg/mL. For most experiments, antibody to peptide
B145 was purified on a B145 peptide column as
described previously (12). Affinity-purified antibody was
1460 The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10
used at a concentration of 1–5 µg/mL. After 1 hour, pro-
tein A–agarose was added and the lysates were incubat-
ed for an additional 30 minutes. Immunoprecipitates
were washed 4 times with lysis buffer and then boiled in
SDS-PAGE sample buffer. Cleared cell lysates and
immunoprecipitates were separated on a 4–10% linear
gradient polyacrylamide gel and examined by autoradi-
ography using a PhosphorImager from Molecular
Dynamics (Sunnyvale, California, USA).
For coprecipitation experiments, unlabeled HPAC
cell lysates were prepared as above, except that affini-
ty-purified antibody to peptide
B145 was used for all experi-
ments. Cell lysates and
immunoprecipitates were sepa-
rated on a 4–15% gradient gel,
transferred to nitrocellulose,
and probed with mouse mono-
clonal antibodies against differ-
ent components of cell adhesion
complexes using an HRP-
labeled anti-mouse secondary
antibody (Jackson ImmunoRe-
search Laboratories Inc.) and
enhanced chemiluminescence
(Amersham Pharmacia AB, Upp-
sala, Sweden) to detect antibody
binding to the nitrocellulose membrane as described
previously. Antibodies were used at the following dilu-
tions: E-cadherin, 1:500; α-catenin, 1:500; β-catenin,
1:1,000; γ-catenin (plakoglobin), 1:1,000; and FAK,
1:1,000. The HRP-labeled secondary antibodies were
used at 1:10,000 to 1:50,000, depending on the lot.
For double immunoprecipitation experiments, meta-
bolically labeled HPAC cell lysates were prepared and
cleared as described above, except that both protein A and
recombinant protein G–agarose (Roche Molecular Bio-
chemicals, Mannheim, Germany) were used during the
The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10 1461
Figure 1
Confocal images of sections of human fetal kid-
ney subjected to double-stain indirect immuno-
fluorescence with antibodies to polycystin-1
(red) and to junctional proteins (green). Each
horizontal series shows, from left to right, the
green channel (junctional protein), the red
channel (polycystin-1), and the total fluores-
cence (dual-scanned image). Colocalization is
seen in the right-hand panels as yellow staining.
(ac) Double staining for polycystin-1 and E-
cadherin, where (a) is E-cadherin, (b) is poly-
cystin-1, and (c) is the dual-scanned image.
(df) Double staining for α-catenin, where (d)
is α-catenin, (e) is polycystin-1, and (f) is the
dual-scanned image. (gi) Double staining for
β-catenin, where (g) is β-catenin, (h) is poly-
cystin-1, and (i) is the dual-scanned image. (jl)
Double staining for γ-catenin, where (j) is γ-
catenin, (k) is polycystin-1, and (l) is the dual-
scanned image. In all these images, polycystin-
1 and E-cadherin or the catenins are colocalized
(yellow) at the apical and lateral aspects of the
epithelial cells. There are areas where distinct
green or red staining can be seen, showing that
not all of the polycystin-1 colocalizes with the
junctional proteins. These images were acquired
with the ×100 lens. (mo) Confocal images of
antibodies to polycystin-1 (red) and β-catenin
(green) incubated in the presence of the peptide
against which the polycystin-1 antibody was
raised. The peptide blocked polycystin-1 stain-
ing (n) but did not affect β-catenin staining (m,
o). Some precipitated anti-rabbit antibody is
seen as red dots (n). These images were
acquired with the ×40 lens.
clearing step. Monoclonal antibodies to E-cadherin or
β-catenin were precipitated with protein G–agarose in the
first immunoprecipitation. The precipitates were washed
3 times with 1 mL lysis buffer and the precipitated pro-
teins were released with 1% SDS at 37°C for 30 minutes.
The eluate was diluted to a final concentration of 0.1%
SDS with lysis buffer, and was reprecipitated with affin-
ity-purified antibody to peptide B145 as described above.
To estimate the molar ratio of polycystin-1 to E-cad-
herin and β-catenin in precipitated complexes, the
results of PhosphorImager analysis (Molecular Dynam-
ics) were normalized to the number of methionines and
cysteines in each molecule. The molar ratio of E-cadherin
to β-catenin obtained by this method was 1:1 (range
0.7–1.1), agreeing with previous results (33, 34).
Sucrose density gradient centrifugation. HPAC cell lysates
prepared as above were layered onto linear 5–20%
(wt/vol) sucrose gradients in lysis buffer and centrifuged
in an SW-40 rotor (Beckman Instruments Inc., Fullerton,
California, USA) at 100,000 gfor 24 hours at 4°C. Twen-
ty-four fractions (0.5 mL each) were collected from each
gradient. Ten-microliter aliquots of each fraction were
used for immunoblotting. The remainder of each frac-
tion was immunoprecipitated with affinity-purified
antibody against polycystin-1. The samples were sepa-
rated on a 7.5% gel, transferred to PVDF (Millipore
Corp., Bedford, Massachusetts, USA), and probed with
mouse monoclonal antibodies against β-catenin. The
blots were then stripped and reprobed with antibodies
to FAK. Densitometry was performed on a densitometer
from Molecular Dynamics.
To determine the distribution of polycystin-1 in gra-
dients, metabolically labeled cell lysates were subjected
to density gradient centrifugation as above. Ten-micro-
liter fractions were separated on a 10% gel and trans-
ferred to PVDF. This was cut into strips at the 66 kDa
molecular weight marker, and probed with antibodies
to β-catenin and actin (Roche Molecular Biochemicals).
The remainder of each fraction was immunoprecipitat-
ed with affinity-purified antibody to polycystin-1 and
separated on a 7.5% gel. Autoradiograms of the
immunoprecipitates were viewed with a Storm Phos-
phorImager System from Molecular Dynamics.
Results
Polycystin colocalizes with E-cadherin and the catenins. We and
others had previously found that polycystin-1 was most
highly expressed in renal epithelial cells of fetal kidney
(12, 35). By indirect immunofluorescence, polycystin-1
was inconsistently detectable in the collecting duct of
adult kidney. To test the hypothesis that polycystin inter-
acts with complexes containing E-cadherin, we stained
sections of fetal kidney with antibodies to polycystin-1
and E-cadherin. We selected tubules that were strongly
stained for polycystin and then scanned those tubules
for simultaneous localization of both polycystin-1 and
cell adhesion molecules. The majority of the tubules
were located just below the nephrogenic zone in the
developing kidneys examined. The staining pattern and
morphology of the tubules suggested that the majority
were of ureteric bud origin. Figure 1 shows that anti-
bodies to E-cadherin (Figure 1a, green) and polycystin-1
(Figure 1b, red) both stained the plasma membrane of
tubular epithelial cells. Figure 1c shows that the majori-
ty of polycystin-1 colocalized with E-cadherin (yellow),
suggesting that the 2 proteins might be part of a com-
plex. There was a significant amount of E-cadherin in
the apical plasma membrane of the tubule because this
fetal tubule is not yet completely polarized. Polycystin-1
and E-cadherin were colocalized in the apical and later-
al plasma membrane. There were areas where polycystin-
1 and E-cadherin did not colocalize, shown as distinct
green or red staining (Figure 1c). These data suggest that
there are pools of E-cadherin and polycystin-1 in the
plasma membrane that do not interact.
The cytoplasmic domain of E-cadherin interacts with
cytoplasmic molecules known as the catenins. These mol-
ecules both regulate the adhesive function of E-cadherin
and mediate signaling from E-cadherin. The observation
that polycystin-1 colocalized with E-cadherin suggested
that polycystin-1 should also colocalize with the catenins.
Distinct complexes containing E-cadherin and subsets of
the catenins have been identified in epithelial cell cultures,
suggesting that these complexes might have different sig-
naling properties. The common components of these
complexes were E-cadherin and α-catenin; β-catenin and
1462 The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10
Figure 2
Expression of polycystin-1 in HPAC cells. (a) RT-PCR with PKD1-specific
primers yielded the predicted 564-bp product (PKD1, – RNase). RT-PCR
with human β-actin primers from 2 exons yielded the mRNA-specific
product of 661 bp (actin, – RNase). The sizes of the markers on the left
are 1,000, 700, 500, and 300 bp. Treatment of the RNA template with
RNase A before RT-PCR eliminated both the actin and the PKD1 prod-
ucts (+ RNase). (b) Polycystin-1 was precipitated from [35S]methion-
ine/cysteine–labeled HPAC cells as described, displayed on a 4–10% lin-
ear gradient gel, and viewed with a PhosphorImager analysis system
(Molecular Dynamics). IM + peptide: precipitation with immune serum
in the presence of the immunizing peptide. IM: precipitation with
immune serum. PI: precipitation with preimmune serum. Note that the
polycystin polypeptide of approximately 500 kDa is precipitated only by
immune serum and only in the absence of peptide.
γ-catenin were found in a mutually exclusive manner in
these complexes. It was possible that polycystin-1 might
be identified in only 1 type of E-cadherin complex.
Accordingly, we stained sections of fetal kidney with anti-
bodies to α-, β-, and γ-catenin and to polycystin-1. We
found that polycystin-1 colocalized with α-catenin (Fig-
ure 1, d–f), β-catenin (Figure 1, g–i), and γ-catenin (Figure
1, j–l). There was no difference in the location or amount
of polycystin-1 associated with β- and α-catenin, suggest-
ing that polycystin-1 did not preferentially interact with a
subset of E-cadherin complexes. All 3 catenins localized to
the apical and lateral plasma membranes (Figure 1, d, g,
and j), paralleling the apicolateral distribution of E-cad-
herin in these immature epithelia. All 3 catenins were
expressed in mesenchymal cells as well as epithelial cells of
the developing kidney. The expression of β-and γ-catenin
was quite high and is easily seen (Figure 1, g and j), where-
as the expression of α-catenin was lower and is not well
seen in this figure (Figure 1d). As we had observed for E-
cadherin, the association of polycystin-1 with the catenins
was not complete. There were areas where polycystin-1 did
not colocalize with the catenins, seen as distinct areas of
green or red staining (Figure 1, f, i, and l).
To demonstrate the specificity of polycystin-1 staining,
we incubated peptide B145 (the antigen against which
the polycystin-1 antibody was raised) with the primary
antibodies during immunostaining. The result of a rep-
resentative experiment is shown in Figure 1, m–o. Pep-
tide abolished polycystin-1 staining (Figure 1n) but had
no effect on β-catenin staining (Figure 1, m and o). These
data show, first, that polycystin-1 staining was peptide
specific. Second, the peptide had no effect on β-catenin
staining, demonstrating that there were no nonspecific
effects of peptide on antibody reactivity. Third, these
images show that there was excellent optical separation
between the red channel (polycystin-1) and the green
channel (in this case, β-catenin), because no β-catenin
staining was found in the red channel.
HPAC pancreatic epithelial cells express polycystin-1. To
determine whether polycystin-1 interacts with E-cad-
herin and the catenins, we sought a human epithelial cell
line that expressed all the proteins of interest. Polycystin-
1 is expressed in the pancreatic ductal epithelium (van
Adelsberg, unpublished observations, and ref. 17), sug-
gesting that pancreatic ductal cell lines would be an
appropriate in vitro model system. Patients with ADPKD
develop pancreatic cysts, as do mice with a targeted
mutation in the PKD1 gene (36). Because the disease is
manifested in the pancreatic duct in vivo, cell lines
derived from the pancreatic duct should be appropriate
for studying the relevant biochemical interactions of
polycystin-1 in vitro. We identified a human pancreatic
epithelial cell line, HPAC, that grows as a monolayer and
expresses the proteins E-cadherin, α-, β-, and γ- catenin,
β1integrins, and FAK (see Figure 3). We screened this cell
line for PKD1 gene and protein expression.
To determine whether the PKD1 gene was expressed
in HPAC cells, we performed RT-PCR on total RNA
isolated from these cells. Figure 2a shows that primers
from the unique 3end of PKD1 amplified the expect-
ed product of 564 bp. This product was purified and
sequenced; the sequence was identical to the pub-
lished PKD1 sequence. Primers from 2 exons of the
human β-actin gene amplified the expected product
of 661 bp. There was no larger product, suggesting
that there was no contamination by genomic DNA.
Treatment of the RNA template with RNase A before
RT-PCR eliminated both the PKD1 and the actin PCR
products (Figure 2a, + RNase), providing additional
The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10 1463
Figure 3
Coimmunoprecipitation of polycystin-1 with junctional complex pro-
teins. (a) Immunoprecipitation of HPAC cell lysates with affinity-purif ied
antibody to polycystin-1. Each panel is an immunoblot probed with anti-
body to the protein named at the top of the panel. In each panel, the left
lane is an aliquot of the cell lysate (lysate). The middle and right lanes
were immunoprecipitated with affinity-purified antibody against poly-
cystin-1 in the absence (–) or presence (+) of the immunizing peptide.
Each set of samples was probed with antibody to E-cadherin, α-catenin,
β-catenin, γ-catenin, β1-integrin, or FAK. Note that E-cadherin and the
catenins coprecipitated with polycystin-1 only in the absence of compet-
ing peptide. The doublet of 120 kDa and 140 kDa in the lysate probed
with antibody to β1-integrin represents reduced (140 kDa) and unre-
duced (120 kDa) β1-integrin. Results are representative of at least 5 inde-
pendent experiments. (b) Double immunoprecipitation of metabolically
labeled HPAC lysates with antibodies to E-cadherin or β-catenin (first
immunoprecipitation) followed by affinity-purified antibody to poly-
cystin-1 (second immunoprecipitation). The left pair of lanes shows the
first immunoprecipitation with monoclonal antibody to β-catenin (β-cat)
or E-cadherin (E-cad). Note the high-molecular-weight polypeptide that
is precipitated as part of the complex. The center pair of lanes shows the
reimmunoprecipitation of the complex with affinity-purified antibody to
polycystin-1. The high-molecular-weight polypeptide seen in the left pan-
els was reprecipitated by polycystin antibody and is therefore poly-
cystin-1. Note that some of the E-cadherin/catenin complex remains
associated with polycystin-1 (arrows). The right pair of lanes shows con-
trol experiments performed on aliquots of the same HPAC lysate. The
lane marked “PKD1” was precipitated with affinity-purified antibody to
polycystin-1. The lane marked “IgG” was precipitated with the same con-
centration of nonimmune rabbit anti-mouse IgG. This figure is a single
gel. The results are representative of 3 independent experiments.
evidence that the PCR product was amplified from
reverse transcribed PKD1 message.
To discover whether there was significant polycystin-1
protein expression in HPAC cells, confluent cultures
were labeled with [35S]methionine/cysteine and lysates
were immunoprecipitated with the anti-peptide anti-
body against polycystin-1. Figure 2b shows that the
immune serum precipitated a polypeptide of approxi-
mately 500 kDa that was not precipitated by the preim-
mune serum. Incubating the immune serum with lysate
in the presence of the immunizing peptide blocked pre-
cipitation of this polypeptide, showing that the
immunoprecipitation was specific for the peptide. Iden-
tical results were obtained with 2 different antibodies
raised against the extracellular domain of polycystin-1
(not shown). These results show that the HPAC cells
express polycystin-1, and that the anti-peptide antibody
specifically precipitates polycystin-1 from HPAC cell
lysates. We were unable to detect polycystin-1 by
immunoblotting lysates of the HPAC cell line, suggest-
ing that polycystin-1 is expressed at low levels in these
cells or that the efficiency of electroblotting this large
membrane protein is very low, or both.
E-cadherin and the catenins are coprecipitated with polycystin-
1. We precipitated HPAC cell lysates with antibody to poly-
cystin-1 and probed the precipitates with antibodies to E-
cadherin and α-, β-, and γ-catenin. We found that
E-cadherin and the catenins were coprecipitated with
polycystin-1 (Figure 3a, – peptide). Adding B145 peptide
to the immunoprecipitation reaction blocked the precip-
itation of E-cadherin and the catenins (Figure 3a, + pep-
tide), demonstrating that the coprecipitation of E-cad-
herin and the catenins was specific for the precipitation of
polycystin-1. These data show that polycystin-1 interacts
with a complex containing E-cadherin and the catenins.
To determine whether polycystin-1 was coprecipitated
with E-cadherin and β-catenin, we performed double
immunoprecipitation (Figure 3b). Antibodies to either
E-cadherin or β-catenin (2 left lanes) precipitated a com-
plex containing a high-molecular-weight polypeptide as
well as E-cadherin, α-catenin, and β-catenin. The molec-
ular weight of this polypeptide was the same as that of
precipitated polycystin-1 (fifth lane, PKD1). Reprecipi-
tation of the complex with anti–polycystin-1 antibodies
showed that the high-molecular-weight polypeptide was
polycystin-1 (center pair of lanes). Note that a small frac-
tion of the E-cadherin/catenin complex was reprecipi-
tated with antibody to polycystin-1 despite treatment of
the precipitated proteins with 1% SDS. From densito-
metric analysis, we determined that only 6.0 ± 0.5% (n =
3) of precipitable E-cadherin or β-catenin complexes
contained polycystin-1.
Polycystin-1 comigrates with
β
-catenin in sucrose density gradi-
ents. The coprecipitation of polycystin-1 with E-cadherin
and the catenins suggested that these proteins were in a
macromolecular complex. Both the colocalization data
and the coimmunoprecipitation data showed that only a
fraction of E-cadherin/catenin complexes contained poly-
cystin-1. To characterize this complex further, we subject-
ed HPAC cell lysates to sucrose density centrifugation. We
chose to use β-catenin as a marker for the
E-cadherin/catenin complex because antibodies to β-
catenin gave the strongest signal in coprecipitation exper-
iments and because β-catenin metabolism is a major deter-
minant of cell proliferation. Polycystin-1 migrated in a
fairly broad range near the center of the gradient (Figure 4,
top). The distribution of β-catenin was broader than that
of polycystin-1, extending into lighter fractions of the gra-
dient. The distribution of polycystin-1 overlapped with
that of β-catenin only in the denser parts of the β-catenin
range. As anticipated, polycystin-1 coprecipitated β-catenin
only in those fractions where polycystin-1 was present (Fig-
ure 4, bottom). The distribution of the complex contain-
ing polycystin-1 and β-catenin was not homogeneous. The
amount of β-catenin coprecipitated with polycystin-1 was
much greater in the lighter part of the polycystin-1 distri-
bution. This skewed distribution of polycystin-catenin
complexes was seen in 4 of 4 independent experiments.
These data provide additional evidence that polycystin-1 is
in a complex containing β-catenin. These results are con-
sistent with both the immunohistochemical and copre-
cipitation data, which showed that polycystin-1 and β-
catenin were not completely colocalized or coprecipitated.
The data suggest that there are 2 pools of polycystin-1, only
1 of which interacts with a complex containing β-catenin.
To discover whether the actin cytoskeleton could par-
ticipate in the formation of a unique complex containing
polycystin-1 and β-catenin, we examined the distribution
of actin within the gradient (Figure 4). Actin comigrated
1464 The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10
Figure 4
Sucrose density gradient centrifugation of HPAC cell lysates. Top: Distrib-
ution of polycystin-1, β-catenin, and actin within the gradient. Polycystin-1
was detected by immunoprecipitation and PhosphorImager analysis (Mol-
ecular Dynamics). Immunoblotting was used to detect β-catenin and actin.
Results are representative of 3 independent experiments. Bottom: Distrib-
ution of β-catenin, FAK, and complexes of β-catenin and polycystin-1 with-
in gradient. The distributions of β-catenin and FAK were determined by
immunoblotting. The distributions of complexes containing polycystin-1
and either β-catenin or FAK were determined by immunoprecipitating each
fraction of the gradient with affinity-purified antibodies to polycystin-1 and
probing immunoblots of the precipitates with antibodies to β-catenin (IP:
polycystin-1, blot: β-catenin) or FAK (IP: polycystin-1, blot: FAK). No FAK
was coprecipitated with polycystin-1. Results are representative of 3 inde-
pendent experiments.
with polycystin-1, although the distribution of actin was
broader than that of polycystin-1. The peak distribution
of polycystin-1 in the gradient matched the peak distri-
bution of actin. These results suggest that actin could be
present in polycystin-1 complexes. The distribution of
actin overlapped, but did not precisely mimic the distri-
bution of β-catenin. This result is not surprising, because
β-catenin is found both in complexes containing E-cad-
herin and in separate complexes containing APC, glyco-
gen synthase kinase, and axin.
Polycystin-1 colocalizes with
β
1integrins but not with FAK.
Although the majority of polycystin-1 was concentrated
in the apical and lateral aspects of the tubular epithelial
cells, there was consistently a small amount of peribasal
staining that did not colocalize with E-cadherin or α-, β-,
or γ-catenin. We had previously found that epithelial cells
from the cpk mouse model of autosomal recessive poly-
cystic kidney disease had increased β1-integrin–mediated
adhesion to collagen and laminin (37). We speculated that
polycystin-1 might interact with both E-cadherin and
integrin-type cell adhesion complexes. Accordingly, we
performed double-label immunofluorescence with anti-
bodies against the human β1-integrin subunit or against
FAK, the well characterized signal transduction molecule
of the focal adhesion. We found that the β1-integrin sub-
unit was not polarized in these primitive renal epithelia,
and that it colocalized with polycystin-1 (Figure 5a). FAK
was detected only in the branching tubules of the ureteric
bud, which also expressed polycystin-1. FAK was confined
to the basal plasma membrane in these epithelia, as was
paxillin, another protein of the focal adhesion (data not
shown). However, FAK did not colocalize with polycystin-
1 (Figure 5b). These data suggest that, although polycystin
could be present in a complex with a β1-integrin, it is not
present in focal adhesions containing FAK and paxillin.
Polycystin-1 does not coprecipitate or comigrate with FAK or
β
1-
integrin. To discover if a small proportion of FAK interact-
ed with polycystin-1, we probed polycystin-1 immuno-
precipitates with antibodies to β1-integrin and FAK. We
found that neither β1-integrin nor FAK were coprecipitat-
ed with polycystin-1 (Figure 3). Paxillin was precipitated
nonspecifically by agarose beads despite extensive pre-
clearing, so we could not assess coprecipitation of paxillin
with polycystin-1. In sucrose density gradient centrifuga-
tion, the distribution of FAK overlapped minimally with
that of polycystin-1 (Figure 4, bottom). FAK did not copre-
cipitate with polycystin-1 in fractions isolated by sucrose
density gradient centrifugation (Figure 4, bottom). These
data show that complexes containing FAK do not contain
polycystin-1, and that polycystin-1 is not part of the focal
adhesion. The coprecipitation data suggest that, if there
is an interaction between polycystin-1 and β1integrins, the
interaction is not strong. The observation that neither β1-
integrin nor FAK could be coprecipitated with polycystin-
1 antibodies provides additional evidence that the copre-
cipitation of E-cadherin and the catenins by polycystin-1
antibodies was specific.
Discussion
We have demonstrated that polycystin-1 is present in com-
plexes containing E-cadherin and the catenins. E-cadherin
and the catenins regulate cell proliferation, cell polarity,
and tissue morphogenesis. In ADPKD, the cysts develop
from tubules, a striking example of abnormal morpho-
genesis. The creation of this abnormal structure, the cyst,
is accompanied by perturbations in cell proliferation and
cell polarity. Our discovery that polycystin-1 interacts with
E-cadherin and α-, β-, and γ-catenin raises obvious ques-
tions about whether polycystin-1 affects the assembly of
junctional complexes or signaling from these complexes.
Because the catenins have effects on such diverse cellular
functions as the polarization of the cytoskeleton (18), reg-
ulation of gene transcription (38), and the formation of
desmosomes (39, 40), the range of possibilities is large, but
is consistent with the protean manifestations of ADPKD.
One of the first questions that arises from our work is
The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10 1465
Figure 5
Confocal images of sections of human fetal kidney stained with antibody to polycystin-1 (red) and either β1-integrin (a) or FAK (green) (b). Only the
dual-scanned images are shown. The areas where polycystin-1 colocalized with β1-integrin (a) are yellow. Polycystin-1 was expressed only in the epithe-
lial structures of the ureteric bud and the S-shaped body (arrowheads), whereas β1-integrin was ubiquitously expressed in both the epithelia (yellow
staining due to colocalization with polycystin-1) and the surrounding mesenchyme (green staining, because polycystin-1 is not expressed in the mes-
enchyme). The absence of yellow staining in (b) shows that polycystin-1 did not colocalize with FAK.
whether polycystin-1 interacts directly with E-cadherin
or with 1 of the catenins. Our data do not allow us to dis-
tinguish between a direct or an indirect interaction. Yeast
2-hybrid analysis conducted in several labs did not iden-
tify E-cadherin or any of the catenins as binding partners
for the carboxyl terminal domain of polycystin-1
(41–45), nor have preliminary biochemical experiments
identified any of these proteins in pull-down assays from
epithelial cell lysates (van Adelsberg, Huan, and Walz,
unpublished observations). Polycystin-1 is a polytopic
integral membrane protein. Current models of its struc-
ture favor 11 transmembrane domains with 5 intracel-
lular loops. Any of these loops could bind directly to E-
cadherin, to a catenin, or to an adapter molecule that
could provide a bridge between polycystin-1 and the
E-cadherin/catenin complex. β-catenin and γ-catenin
bind directly to α-catenin, giving rise to 2 distinct cad-
herin-catenin complexes (46, 47). Polycystin-1 interacts
with both of these complexes. The most parsimonious
explanation is that polycystin-1 (or the hypothetical
adapter) interacts with E-cadherin or with α-catenin,
which are common to both complexes.
We found that the interaction between polycystin-1
and the E-cadherin/catenin complex was not stoichio-
metric. The colocalization data showed separate areas of
polycystin-1 and cadherin/catenin staining, suggesting
that polycystin-1 interacts only with a discrete pool of
cadherin and catenins. Sucrose density gradient cen-
trifugation also identified a pool of polycystin-1 that did
not interact with β-catenin. These data suggest that the
interaction of polycystin-1 with the E-cadherin/catenin
complex could be regulated. Complexes containing
E-cadherin, the catenins, and polycystin-1 might have a
specialized role in signal transduction.
What is the role of an interaction between polycystin-1
and the E-cadherin/catenin complex? One possibility is
that polycystin-1 is involved in the formation of the
E-cadherin/catenin complex. This hypothesis appears
unlikely, because PKD1 null mice die during late organo-
genesis, not at the peri-implantation stage as do E-cad-
herin or β-catenin null mice (48–51). Preliminary work
(Huan and van Adelsberg, unpublished observations)
suggests that E-cadherin/catenin complexes are intact in
cyst-lining epithelial cells isolated from ADPKD kidneys.
The interaction of polycystin-1 with the E-cad-
herin/catenin complex could be involved in correct tar-
geting of polycystin-1. Mutations that affect targeting
rather than biochemical activity are well documented.
One well characterized example is the F508 mutation
of the cystic fibrosis transmembrane conductance reg-
ulator (CFTR) ion channel, which produces a func-
tional channel that is not targeted to the plasma mem-
brane (52). Retention of targeted proteins in the
correct plasma membrane domain is one mechanism
for maintaining cell polarity. This mechanism has been
well characterized for Na,K-ATPase. This protein inter-
acts directly with ankyrin (27, 53), which is complexed
with nonerythrocyte spectrin (fodrin), and then,
through actin and α-catenin, to the E-cadherin/catenin
complex (24, 33, 34, 54, 55). The hypothesis that the
interaction between polycystin-1 and the E-cad-
herin/catenin complex stabilizes correctly polarized
polycystin-1 in a manner analogous to that of Na,K-
ATPase predicts several results. First, the half-life of
mutant polycystin-1 in ADPKD cyst-lining epithelial
cells should be reduced. Second, the complex should
contain cytoskeletal elements like actin and possibly
ankyrin or spectrin. We found that polycystin-1 and
actin comigrated in sucrose gradients, suggesting that
they could be part of a complex. Two of the candidate
proteins that interact with the carboxyl terminal
domain of polycystin-1 contain spectrin repeats, sug-
gesting that polycystin-1 could interact with spectrin
or fodrin. Third, polycystin-1 should be incorrectly tar-
geted in cyst-lining epithelial cells of ADPKD. A corol-
lary to the hypothesis is that only some mutations in
PKD1 would disrupt the interaction of polycystin-1
with E-cadherin and the catenins, whereas other muta-
tions might cause different effects on polycystin-1 pro-
cessing or function. Apparent mislocalization of poly-
cystin-1 has been observed in at least 1 case of ADPKD.
These data support the idea that the interaction of
E-cadherin/catenin complex with polycystin-1 could be
involved in polycystin-1 targeting.
A third possible role for the interaction of polycystin-1
with E-cadherin/catenin complexes is that of modulat-
ing signaling from these complexes, in particular by
effects on β-catenin metabolism. This is an attractive
hypothesis for a variety of reasons. First, expression of
the carboxyl terminal tail of polycystin-1 as a membrane-
targeted chimera protected cytoplasmic β-catenin from
degradation by the proteasome and increased transcrip-
tion from a reporter for Siamois, a genuine β-catenin tar-
get (56). Second, high expression of c-myc, a likely
β-catenin target, is found in both human ADPKD cyst-
lining epithelial cells and cyst-lining epithelial cells from
2 rodent models of polycystic kidney disease (57–59).
Transgenic mice that overexpress c-myc in renal epithe-
lial cells develop fatal polycystic kidney disease, demon-
strating that c-myc is in the pathway to cyst formation
(60). Third, nuclear β-catenin staining is found in the bcl-
2null mouse model of polycystic kidney disease (61).
The interaction of β-catenin with LEF transcription fac-
tors is known to target β-catenin to the nucleus (62).
These data suggest that aberrant β-catenin signaling
could be a common feature of polycystic kidney diseases.
This hypothesis predicts that the half-life of β-catenin
will be prolonged in cyst-lining epithelial cells in poly-
cystic kidney disease, and that expression of other
β-catenin target genes will be upregulated in these cells.
Our data show that polycystin-1 does not interact
directly with focal adhesions in the steady state that is
characteristic of epithelial cells in tissue or confluent cul-
tures. We had previously observed increased β1-inte-
grin–mediated adhesion to extracellular matrix in cells
isolated from a mouse model of polycystic kidney dis-
ease, suggesting that defects in cell-matrix interaction
might contribute to cyst formation (37). Our new data
suggest that this phenotype is unlikely to be caused by a
direct effect of polycystin-1 on focal adhesions, but
could be an example of cross-talk between cadherins and
integrin (63, 64). We had found that the integrin that
mediates increased matrix adhesion in cpk mouse epithe-
lial cells is α1β1-integrin (van Adelsberg and Almonte,
1466 The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10
unpublished observations). Expression of this integrin
is upregulated in autosomal recessive polycystic kidney
disease and in ADPKD (65), suggesting that upregula-
tion of α1β1-integrin–mediated adhesion is part of a
common pathway in cyst formation. This hypothesis
would place the regulation of α1β1-integrin expression
downstream of the PKD and ARPKD genes.
Our observation that β1integrins are not polarized dur-
ing renal development is not new (66–68). However, we
are the first to describe the localization of FAK and pax-
illin in developing kidney. Our data show that paxillin
and FAK are confined to the basal plasma membrane.
The β1integrins, however, are basal, lateral, and apical.
This result suggests that either these apical and lateral
integrins are not involved in signaling or that other sig-
naling complexes containing integrins but not the clas-
sic molecules of the focal adhesion are present in renal
epithelial cells. Because Madin-Darby canine kidney
(MDCK) cells express apical β1integrins and respond to
them by remodeling the axis of apicobasal polarity when
collagen is overlaid on an MDCK monolayer (69, 70), it
seems likely that the apical β1integrins are functional.
Acknowledgments
Janet van Adelsberg is an Established Investigator of the
American Heart Association. This project was funded in
part by grant 1-FY97-0513 from the March of Dimes.
Confocal microscopy was performed at the Confocal
Microscopy Facility of Columbia University, which was
established by National Institutes of Health (NIH)
Shared Instrumentation Grant 1S10-RR10406 and is
supported by NIH grant 5-P30-CA13696 as part of the
Herbert Irving Cancer Center at Columbia University.
We are grateful to Theresa Swayne of the Confocal
Microscopy Facility for her invaluable assistance in
obtaining the confocal images shown here, and to
Vivette D’Agati and Llewellyn Ward of the Department
of Pathology at Columbia University for their help in
obtaining and analyzing the kidney sections shown here.
1.Gabow, P.A. 1993. Autosomal dominant polycystic kidney disease. N.
Engl. J. Med. 329:332–342.
2. European Polycystic Kidney Disease Consortium. 1994. The polycystic
kidney disease 1 gene encodes a 14 kb transcript and lies within a dupli-
cated region on chromosome 16. Cell. 77:881–894.
3. American PKD1 Consortium. 1995. Analysis of the genomic sequence
for the autosomal dominant polycystic kidney disease gene predicts the
presence of a leucine-rich repeat. Hum. Mol. Genet. 4:575–582.
4. International Polycystic Kidney Disease Consortium. 1995. Polycystic
kidney disease: the complete structure of the PKD1 gene and its protein.
Cell. 81:289–298.
5. Hughes, J., et al. 1995. The polycystic kidney disease 1 gene encodes a
novel protein with multiple cell recognition domains. Nat. Genet.
10:151–159.
6.Ariza, M., et al. 1997. A family with a milder form of adult dominant
polycystic kidney disease not linked to the PKD1 (16p) or PKD2 (4q)
genes. J. Med. Genet. 34:587–589.
7.Daoust, M.C., Reynolds, D.M., Bichet, D.G., and Somlo, S. 1995. Evi-
dence for a third genetic locus for autosomal dominant polycystic kid-
ney disease. Genomics. 25:733–736.
8. de Almeida, S., et al. 1995. Autosomal dominant polycystic kidney dis-
ease: evidence for the existence of a third locus in a Portuguese family.
Hum. Genet. 96:83–88.
9.Turco, A.E., Clementi, M., Rossetti, S., Tenconi, R., and Pignatti, P.F.
1996. An Italian family with autosomal dominant polycystic kidney dis-
ease unlinked to either the PKD1 or PKD2 gene. Am. J. Kidney Dis.
28:759–761.
10.Qian, F., et al. 1997. PKD1 interacts with PKD2 through a probable
coiled-coil domain. Nat. Genet. 16:179–183.
11.Tsiokas, L., Kim, E., Arnould, T., Sukhatme, V.P., and Walz, G. 1997.
Homo- and heterodimeric interactions between the gene products of
PKD1 and PKD2. Proc. Natl. Acad. Sci. USA. 94:6965–6970.
12.van Adelsberg, J.S., Chamberlain, S., and D’Agati, V. 1997. Polycystin
expression is temporally and spatially regulated during renal develop-
ment. Am. J. Physiol. 272:F602–F609.
13. Palsson, R., et al. 1996. Immunolocalization of PKD1 in normal kidneys.
Mol. Med. 2:702–711.
14. Ibraghimov-Beskrovnaya, O., et al. 1997. Polycystin: in vitro synthesis, in
vivo tissue expression, and subcellular localization identifies a large
membrane-associated protein. Proc. Natl. Acad. Sci. USA. 94:6397–6402.
15. Ong, A.C., Ward, C.J., Biddolph, S., Migone, N., and Harris, P.C. 1997.
Polycystin expression in PKD1, infantile PKD1, and TSC-2/PKD1 cystic
kidney: evidence against a two-hit disease mechanism in cyst initiation.
J. Am. Soc. Nephrol. 8:378A. (Abstr.)
16.Griffin, M.D., et al. 1997. Expression of polycystin in mouse
metanephros and extra-metanephric tissues. Kidney Int. 52:1196–1205.
17.Geng, L., et al. 1996. Identification and localization of polycystin, the
PKD1 gene product. J. Clin. Invest. 98:2674–2682.
18. Nelson, W.J. 1993. Regulation of cell surface polarity in renal epithelia.
Pediatr. Nephrol. 7:599–604.
19. Gumbiner, B.M., and McCrea, P.D. 1993. Catenins as mediators of the
cytoplasmic functions of cadherins. J. Cell Sci. Suppl. 17:155–158.
20. Aberle, H., Schwartz, H., and Kemler, R. 1996. Cadherin-catenin com-
plex: protein interactions and their implications for cadherin function.
J. Cell Biochem. 61:514–523.
21.Gumbiner, B. 1995. Signal transduction by β-catenin. Curr. Opin. Cell
Biol. 7:634–640.
22. Barth, A.I., Nathke, I.S., and Nelson, W.J. 1997. Cadherins, catenins and
APC protein: interplay between cytoskeletal complexes and signaling
pathways. Curr. Opin. Cell Biol. 9:683–690.
23. Grantham, J.J. 1990. Polycystic kidney disease: neoplasia in disguise. Am.
J. Kidney Dis. 15:110–116.
24. Nelson, W.J., Shore, E.M., Wang, A.Z., and Hammerton, R.W. 1990. Iden-
tification of a membrane-cytoskeletal complex containing the cell adhe-
sion molecule uvomorulin (E-cadherin), ankyrin, and fodrin in Madin-
Darby canine kidney epithelial cells. J. Cell Biol. 110:349–357.
25. McNeill, H., Ozawa, M., Kemler, R., and Nelson, W.J. 1990. Novel func-
tion of the cell adhesion molecule uvomorulin as an inducer of cell sur-
face polarity. Cell. 62:309–316.
26.Graeve, L., Drickamer, K., and Rodriguez-Boulan, E. 1989. Polarized
endocytosis by Madin-Darby canine kidney cells transfected with func-
tional chicken liver glycoprotein receptor. J. Cell Biol. 109:2809–2816.
27. Devarajan, P., Scaramuzzino, D.A., and Morrow, J.S. 1994. Ankyrin binds
to two distinct cytoplasmic domains of Na,K-ATPase alpha subunit.
Proc. Natl. Acad. Sci. USA. 91:2965–2969.
28. Alejandro, V.S., et al. 1995. Postischemic injury, delayed function and
Na+/K(+)-ATPase distribution in the transplanted kidney. Kidney Int.
48:1308–1315.
29. Molitoris, B.A., Geerdes, A.M., and McIntosh, J.R. 1991. Dissociation and
redistribution of Na+,K(+)-ATPase from its surface membrane actin
cytoskeletal complex during cellular ATP depletion. J. Clin. Invest.
88:462–469.
30. Wilson, P.D., et al. 1991. Reversed polarity of Na(+)-K(+)-ATPase: mislo-
cation to apical plasma membranes in polycystic kidney disease epithe-
lia. Am. J. Physiol. 260:F420–F430.
31. Brill, S.R., et al. 1996. Immunolocalization of ion transport proteins in
human autosomal dominant polycystic kidney disease. Proc. Natl. Acad.
Sci. USA. 93:10206–10211.
32. Carone, F.A., et al. 1994. Cell polarity in human renal cystic disease. Lab.
Invest. 70:648–655.
33.Ozawa, M., and Kemler, R. 1992. Molecular organization of the uvo-
morulin-catenin complex. J. Cell Biol. 115:989–996.
34. Jou, T.S., Stewart, D.B., Stappert, J., Nelson, W.J., and Marrs, J.A. 1995.
Genetic and biochemical dissection of protein linkages in the cadherin-
catenin complex. Proc. Natl. Acad. Sci. USA. 92:5067–5071.
35. van Adelsberg, J.S. 1999. The role of the polycystins in kidney develop-
ment. Pediatr. Nephrol. 13:454–459.
36. Lu, W., et al. 1997. Perinatal lethality with kidney and pancreas defects
in mice with a targeted Pkd1 mutation. Nat. Genet. 17:179–181.
37.van Adelsberg, J.S. 1994. Murine polycystic kidney epithelial cell lines
have increased integrin-mediated adhesion to collagen. Am. J. Physiol.
276:F1082–F1093.
38. Bullions, L.C., and Levine, A.J. 1998. The role of beta-catenin in cell adhe-
sion, signal transduction, and cancer. Curr. Opin. Oncol. 10:81–87.
39. Marrs, J.A., et al. 1995. Plasticity in epithelial cell phenotype: modulation
by expression of different cadherin cell adhesion molecules. J. Cell Biol.
129:507–519.
40.Lewis, J.E., et al. 1997. Cross-talk between adherens junctions and
desmosomes depends on plakoglobin. J. Cell Biol. 136:919–934.
41. Qian, F., Germino, J., Zhang, F., and Germino, G. 1997. Putative binding
partners of polycystin. J. Am. Soc. Nephrol. 8:380A. (Abstr.)
The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10 1467
42. Maeda, Y., et al. 1997. Identification of potential binding proteins for
PKD2. J. Am. Soc. Nephrol. 8:376A. (Abstr.)
43. Kim, E., Arnould, T., and Walz, G. 1997. Isolation of polycystin-inter-
acting proteins. J. Am. Soc. Nephrol. 8:375A. (Abstr.)
44. Kim, E., et al. 1999. Interaction between RGS7 and polycystin. Proc. Natl.
Acad. Sci. USA. 96:6371–6376.
45. Kim, K., Sharma, C.P., and Arnaout, M.A. 1997. An interaction of 14-3-
3 adapter proteins with polycystin. J. Am. Soc. Nephrol. 8:375A. (Abstr.)
46. Hinck, L., Näthke, I.S., Papkoff, J., and Nelson, W.J. 1994. Dynamics of
cadherin/catenin complex formation: novel protein interactions and
pathways of complex assembly. J. Cell Biol. 125:1327–1340.
47. Näthke, I.S., Hinck, L., Swedlow, J.R., Papkoff, J., and Nelson, W.J. 1994.
Defining interactions and distributions of cadherin and catenin com-
plexes in polarized epithelial cells. J. Cell Biol. 125:1341–1352.
48. Larue, L., Ohsugi, M., Hirchenhain, J., and Kemler, R. 1994. E-cadherin
null mutant embryos fail to form a trophectoderm epithelium. Proc.
Natl. Acad. Sci. USA. 91:8263–8267.
49.Riethmacher, D., Brinkmann, V., and Birchmeier, C. 1995. A targeted
mutation in the mouse E-cadherin gene results in defective preimplan-
tation development. Proc. Natl. Acad. Sci. USA. 92:855–859.
50. Levine, E., Lee, C.H., Kintner, C., and Gumbiner, B.M. 1994. Selective dis-
ruption of E-cadherin function in early Xenopus embryos by a dominant
negative mutant. Development. 120:901–909.
51. Haegel, H., et al. 1995. Lack of beta-catenin affects mouse development
at gastrulation. Development. 121:3529–3537.
52. Kopito, R.R. 1999. Biosynthesis and degradation of CFTR. Physiol. Rev.
79:S167–S173.
53. Jordan, C., Puschel, B., Koob, R., and Drenckhahn, D. 1995. Identifica-
tion of a binding motif for ankyrin on the alpha-subunit of Na+,K(+)-
ATPase. J. Biol. Chem. 270:29971–29975.
54. McCrea, P.D., Turck, C.W., and Gumbiner, B.M. 1991. A homolog of the
armadillo protein in Drosophila (plakoglobin) associated with E-cadherin.
Science. 254:1359–1361.
55. Hirano, S., Kimoto, N., Shimoyama, Y., Hirohashi, S., and Takechi, M.
1992. Identification of a neural α-catenin as a key regulator of cadherin
function and multicellular organization. Cell. 70:293–301.
56. Kim, E., et al. 1999. The polycystic kidney disease 1 gene product mod-
ulates Wnt signaling. J. Biol. Chem. 274:4947–4953.
57.Cowley, B.D., Jr., Smardo, F.L., Jr., Grantham, J.J., and Calvet, J.P. 1987.
Elevated c-myc protooncogene expression in autosomal recessive poly-
cystic kidney disease. Proc. Natl. Acad. Sci. USA. 84:8394–8398.
58. Klingel, et al. 1992. Expression of differentiation antigens and growth-
related genes in normal kidney, autosomal dominant polycystic kidney
disease, and renal cell carcinoma. Am. J. Kidney Dis. 19:22–30.
59. Cowley, B.D., Jr., et al. 1993. Autosomal-dominant polycystic kidney dis-
ease in the rat. Kidney Int. 43:522–534.
60. Trudel, M., D’Agati, V.D., and Cost antini, F. 1991. C-myc as an induc-
er of polycystic kidney disease in transgenic mice. Kidney Int.
39:665–671.
61. Sorenson, C.M. 1999. Nuclear localization of beta-catenin and loss of
apical brush border actin in cystic tubules of bcl-2 –/– mice. Am. J. Phys-
iol. 276:F210–F217.
62. Huber, O., et al. 1996. Nuclear localization of beta-catenin by interaction
with transcription factor LEF-1. Mech. Dev. 59:3–10.
63. Monier-Gavelle, F., and Duband, J.L. 1997. Cross talk between adhesion
molecules: control of N-cadherin activity by intracellular signals elicit-
ed by beta1 and beta3 integrins in migrating neural crest cells. J. Cell Biol.
137:1663–1681.
64. Novak, A., et al. 1998. Cell adhesion and the integrin-linked kinase reg-
ulate the LEF-1 and beta-catenin signaling pathways. Proc. Natl. Acad. Sci.
USA. 95:4374–4379.
65. Daikha-Dahmane, F., et al. 1997. Distribution of alpha-integrin subunits
in fetal polycystic kidney diseases. Pediatr. Nephrol. 11:267–273.
66. Korhonen, M., Ylanne, J., Laitinen, L., and Virtanen, I. 1990. The alpha
1-alpha 6 subunits of integrins are characteristically expressed in distinct
segments of developing and adult human nephron. J. Cell Biol.
111:1245–1254.
67. Simon, E.E., and McDonald, J.A. 1990. Extracellular matrix receptors in
the kidney cortex. Am. J. Physiol. 259:F783–F792.
68. Korhonen, M., Laitinen, L.,Ylänne, J., Gould, V.E., and Virtanen, I. 1992.
Integrins in developing, normal, and malignant human kidney. Kidney
Int. 41:641–644.
69. Schwimmer, R., and Ojakian, G.K. 1995. The alpha 2 beta 1 integrin reg-
ulates collagen-mediated MDCK epithelial membrane remodeling and
tubule formation. J. Cell Sci. 108:2487–2498.
70. Zuk, A., and Matlin, K.S. 1996. Apical beta 1 integrin in polarized MDCK
cells mediates tubulocyst formation in response to type I collagen over-
lay. J. Cell Sci. 109:1875–1889.
1468 The Journal of Clinical Investigation | November 1999 | Volume 104 | Number 10
... While PC1 exhibits high expression in fetal renal tissue, its expression is subdued in adult tissue. It is found in the cilium, plasma membrane, and adhesion complex in polarized epithelial cells, suggesting its involvement in protein-protein interactions, cell-cell adhesion, and cell-matrix interactions [22][23][24]. PC2 is an integral six-transmembrane protein with intracellular N and C termini. It functions as a nonselective calcium-permeable transient receptor potential channel [25,26]. ...
Article
Full-text available
Autosomal dominant polycystic kidney disease (ADPKD) is the most common hereditary kidney disorder, but kidneys are not the only organs involved in this systemic disorder. Individuals with the condition may display additional manifestations beyond the renal system, involving the liver, pancreas, and brain in the context of cystic manifestations, while involving the vascular system, gastrointestinal tract, bones, and cardiac valves in the context of non-cystic manifestations. Despite kidney involvement remaining the main feature of the disease, thanks to longer survival, early diagnosis, and better management of kidney-related problems, a new wave of complications must be faced by clinicians who treated patients with ADPKD. Involvement of the liver represents the most prevalent extrarenal manifestation and has growing importance in the symptom burden and quality of life. Vascular abnormalities are a key factor for patients’ life expectancy and there is still debate whether to screen or not to screen all patients. Arterial hypertension is often the earliest onset symptom among ADPKD patients, leading to frequent cardiovascular complications. Although cardiac valvular abnormalities are a frequent complication, they rarely lead to relevant problems in the clinical history of polycystic patients. One of the newest relevant aspects concerns bone disorders that can exert a considerable influence on the clinical course of these patients. This review aims to provide the “state of the art” among the extrarenal manifestation of ADPKD.
... While PC1 exhibits high expression in fetal renal tissue, its expression is subdued in adult tissue. It is found in the cilium, plasma membrane, and adhesion complex in polarized epithelial cells, suggesting its involvement in protein-protein interactions, cell-cell adhesion, and cell-matrix interactions [24][25][26]. PC2 is an integral six-3 transmembrane protein with intracellular N and C termini. It functions as a nonselective calciumpermeable transient receptor potential channel [27,28]. ...
Preprint
Full-text available
Autosomal dominant polycystic kidney disease (ADPKD) is the most common hereditary kidney disorder but kidneys are not the only organs involved in this systemic disorder. Individuals with the condition may display additional manifestations beyond the renal system, involving liver, pancreas and brain in the con-text of cystic manifestations, while vascular system, gastrointestinal tract, bones, and cardiac valves in the context of non-cystic manifestations. Despite kidney involvement remain the main feature of the disease, thanks to new generation therapies, early diagnosis and better management of kidney related problems, a new wave of complications must be faced from clinicians who treated polycystic patients. Involvement of the liver represents the most prevalent extrarenal manifestation and grows importance in symptom burden and quality of life. Vascular abnormalities are a key factor for patients’ life expectancy and there’s still de-bate whether to screen or not to screen all patients. Arterial hypertension is often the earliest onset among polycystic patients leading to frequent cardiovascular complications. Although cardiac valvular abnormali-ties is a frequent complication, it rarely leads to relevant problems in the clinical history of polycystic pa-tients. One of the new relevant aspects concern bone disorders, that can exert a considerable influence on the clinical course of these patients. This review aims to provide the “state of art” among extrarenal mani-festation of ADPKD.
... PC1 forms a complex containing E-cadherin and alpha-, beta-, and gamma-catenin. E-cadherin is a signaling center for cellular pathways and is constantly trafficked through an endocytic recycling pathway, which regulates cell adhesion and polarity [7,9]. PC2 is a multi-pass membrane protein that functions as a non-selective cationic channel permeable to calcium. ...
Article
Full-text available
Autosomal-Dominant Polycystic Kidney Disease (ADPKD) is a monogenic disorder initiated by mutations in either PKD1 or PKD2 genes, responsible for encoding polycystin 1 and polycystin 2, respectively. These proteins are primarily located within the primary cilia. The disease follows an inexorable progression, leading most patients to severe renal failure around the age of 50, and extra-renal complications are frequent. A cure for ADPKD remains elusive, but some measures can be employed to manage symptoms and slow cyst growth. Tolvaptan, a vasopressin V2 receptor antagonist, is the only drug that has been proven to attenuate ADPKD progression. Recently, autophagy, a cellular recycling system that facilitates the breakdown and reuse of aged or damaged cellular components, has emerged as a potential contributor to the pathogenesis of ADPKD. However, the precise role of autophagy in ADPKD remains a subject of investigation, displaying a potentially twofold impact. On the one hand, impaired autophagy may promote cyst formation by inducing apoptosis, while on the other hand, excessive autophagy may lead to fibrosis through epithelial to mesenchymal transition. Promising results of autophagy inducers have been observed in preclinical studies. Clinical trials are warranted to thoroughly assess the long-term safety and efficacy of a combination of autophagy inducers with metabolic and/or aquaferetic drugs. This research aims to shed light on the complex involvement of autophagy in ADPKD, explore the regulation of autophagy in disease progression, and highlight the potential of combination therapies as a promising avenue for future investigations.
... The large 54 kb human PKD1 gene and 462 kDa PC1 protein are differentially expressed in a wide range of adult tissues, including epithelial and non-epithelial cell types. 11,13,[15][16][17] Interestingly, PKD1 expression is regulated developmentally in the kidneys. PC1 has the highest levels during fetal life, with readily detectable expression in proximal and distal tubular epithelial cells and collecting ducts. ...
Article
Full-text available
Autosomal dominant polycystic kidney disease (ADPKD) causes renal cysts and leads to end-stage renal disease in midlife due mainly to PKD1 gene mutations. Virtually no studies have explored gene therapeutic strategies for long-term effective treatment of PKD. Toward this aim, the severely cystic Pkd1-null mouse model was targeted with a series of transgene transfers using genomic Pkd1 under its regulatory elements (Pkd1wt), a kidney-targeted Pkd1 gene (SBPkd1), or Pkd1Minigene. The introduced Pkd1wt gene constructs with ∼8-fold overexpression display similar endogenous cellular profiles and full complementation of Pkd1-/- phenotype and establish the referral Pkd1 genomic length for proper regulation. SBPkd1 transgene transfer expressing 0.6- or 7-fold Pkd1 endogenous levels is sufficient to correct glomerular and proximal tubular cysts and to markedly postpone cysts in other tubular segments as well, showing that the small SB elements appreciably overlap with Pkd1 promoter/5' UTR regulation. Renal-targeted Pkd1Minigene at high copy numbers conveys an expression level similar to that of the endogenous Pkd1 gene, with widespread and homogeneous weak Pkd1 cellular signal, partially rescuing all cystic tubular segments. These transgene transfers determine that Pkd1 intragenic sequences regulate not only expression levels but also spatiotemporal patterns. Importantly, our study demonstrates that Pkd1 re-expression from hybrid therapeutic constructs can ameliorate, with considerably extended lifespan, or eliminate PKD.
... As one might expect from its structural complexity, a multitude of functions have been proposed for PC1. In addition to a role in cell adhesion based on its ECR domains, PC1 is reported to functionally interact with cadherins and to be localized to multiple plasma membrane domains, including adherens junctions, desmosomes, focal adhesions, and the primary cilium (Huan and van Adelsberg, 1999;Scheffers Frontiers in Molecular Biosciences frontiersin.org Yoder et al., 2002). ...
Article
Full-text available
Polycystin-1 (PC1) is an 11-transmembrane (TM) domain-containing protein encoded by the PKD1 gene, the most frequently mutated gene leading to autosomal dominant polycystic kidney disease (ADPKD). This large (> 462 kDal) protein has a complex posttranslational maturation process, with over five proteolytic cleavages having been described, and is found at multiple cellular locations. The initial description of the binding and activation of heterotrimeric Gαi/o by the juxtamembrane region of the PC1 cytosolic C-terminal tail (C-tail) more than 20 years ago opened the door to investigations, and controversies, into PC1’s potential function as a novel G protein-coupled receptor (GPCR). Subsequent biochemical and cellular-based assays supported an ability of the PC1 C-tail to bind numerous members of the Gα protein family and to either inhibit or activate G protein-dependent pathways involved in the regulation of ion channel activity, transcription factor activation, and apoptosis. More recent work has demonstrated an essential role for PC1-mediated G protein regulation in preventing kidney cyst development; however, the mechanisms by which PC1 regulates G protein activity continue to be discovered. Similarities between PC1 and the adhesion class of 7-TM GPCRs, most notably a conserved GPCR proteolysis site (GPS) before the first TM domain, which undergoes autocatalyzed proteolytic cleavage, suggest potential mechanisms for PC1-mediated regulation of G protein signaling. This article reviews the evidence supporting GPCR-like functions of PC1 and their relevance to cystic disease, discusses the involvement of GPS cleavage and potential ligands in regulating PC1 GPCR function, and explores potential connections between PC1 GPCR-like activity and regulation of the channel properties of the polycystin receptor-channel complex.
... Meanwhile, PKD1 ablation in airway cilia cells also reduced E-cadherin levels. Earlier studies have reported that the E-cadherin protein directly interacted with PC1 in human kidney cells and pancreatic adenocarcinoma cells [39,40]. Since E-cadherin is one of the most widely studied marker of adherens junction, the loss of E-cadherin in PKD1 +/− lungs may contribute to the impaired airway epithelial barrier function. ...
Article
Full-text available
Background Autosomal dominant polycystic kidney disease (ADPKD) is a prevalent genetic disorder, mainly characterized by the development of renal cysts, as well as various extrarenal manifestations. Previous studies have shown that ADPKD is related to bronchiectasis, while its pathogenic mechanism is unclear. In previous studies, we have generated the PKD1+/− pigs to simulate the progression of cyst formation and physiological alterations similar to those seen in ADPKD patients. Methods Phenotypic changes to airway epithelial cell and mesenchymal cell in PKD1+/− pigs were assessed by histological analysis. The molecular mechanisms driving these processes were investigated by using PKD1+/− pig lungs, human mesenchymal cells, and generating PKD1 deficient human epithelial cells. Results We identified bronchiectasis in PKD1+/− pigs, which is consistent with the clinical symptoms in ADPKD patients. The deficiency of PKD1 suppressed E-cadherin expression in the airway epithelial barrier, which aggravated invasion and leaded to a perpetuated inflammatory response. During this process, extracellular matrix (ECM) components were altered, which contributed to airway smooth muscle cell phenotype switch from a contractile phenotype to a proliferative phenotype. The effects on smooth muscle cells resulted in airway remodeling and establishment of bronchiectasis. Conclusion To our knowledge, the PKD1+/− pig provides the first model recapitulating the pathogenesis of bronchiectasis in ADPKD. The role of PKD1 in airway epithelial suggests a potential target for development of new strategies for the diagnosis and treatment of bronchiectasis.
Article
Apicobasal epithelial polarity controls the functional properties of most organs. Thus, there has been extensive research on the molecular intricacies governing the establishment and maintenance of cell polarity. Whereas loss of apicobasal polarity is a well-documented phenomenon associated with multiple diseases, less is known regarding another type of apicobasal polarity alteration – the inversion of polarity. In this Review, we provide a unifying definition of inverted polarity and discuss multiple scenarios in mammalian systems and human health and disease in which apical and basolateral membrane domains are interchanged. This includes mammalian embryo implantation, monogenic diseases and dissemination of cancer cell clusters. For each example, the functional consequences of polarity inversion are assessed, revealing shared outcomes, including modifications in immune surveillance, altered drug sensitivity and changes in adhesions to neighboring cells. Finally, we highlight the molecular alterations associated with inverted apicobasal polarity and provide a molecular framework to connect these changes with the core cell polarity machinery and to explain roles of polarity inversion in health and disease. Based on the current state of the field, failure to respond to extracellular matrix (ECM) cues, increased cellular contractility and membrane trafficking defects are likely to account for most cases of inverted apicobasal polarity.
Article
Autosomal recessive polycystic kidney disease is an early onset inherited hepatorenal disorder affecting around 1 in 20,000 births with no approved specific therapies. The disease is almost always caused by variations in the polycystic kidney and hepatic disease 1 gene, which encodes fibrocystin (FC), a very large, single‐pass transmembrane glycoprotein found in primary cilia, urine and urinary exosomes. By comparison to proteins involved in autosomal dominant PKD, our structural and molecular understanding of FC has lagged far behind such that there are no published experimentally determined structures of any part of the protein. Bioinformatics analyses predict that the ectodomain contains a long chain of immunoglobulin‐like plexin‐transcription factor domains, a protective antigen 14 domain, a tandem G8‐TMEM2 homology region and a sperm protein, enterokinase and agrin domain. Here we review current knowledge on the molecular function of the protein from a structural perspective.
Article
Full-text available
Calcium-dependent cell-cell adhesion is mediated by the cadherin family of cell adhesion proteins. Transduction of cadherin adhesion into cellular reorganization is regulated by cytosolic proteins, termed alpha-, beta-, and gamma-catenin (plakoglobin), that bind to the cytoplasmic domain of cadherins and link them to the cytoskeleton. Previous studies of cadherin/catenin complex assembly and organization relied on the coimmunoprecipitation of the complex with cadherin antibodies, and were limited to the analysis of the Triton X-100 (TX-100)-soluble fraction of these proteins. These studies concluded that only one complex exists which contains cadherin and all of the catenins. We raised antibodies specific for each catenin to analyze each protein independent of its association with E-cadherin. Extracts of Madin-Darby canine kidney epithelial cells were sequentially immunoprecipitated and immunoblotted with each antibody, and the results showed that there were complexes of E-cadherin/alpha-catenin, and either beta-catenin or plakoglobin in the TX-100-soluble fraction. We analyzed the assembly of cadherin/catenin complexes in the TX-100-soluble fraction by [35S]methionine pulse-chase labeling, followed by sucrose density gradient fractionation of proteins. Immediately after synthesis, E-cadherin, beta-catenin, and plakoglobin cosedimented as complexes. alpha-Catenin was not associated with these complexes after synthesis, but a subpopulation of alpha-catenin joined the complex at a time coincident with the arrival of E-cadherin at the plasma membrane. The arrival of E-cadherin at the plasma membrane coincided with an increase in its insolubility in TX-100, but extraction of this insoluble pool with 1% SDS disrupted the cadherin/catenin complex. Therefore, to examine protein complex assembly in both the TX-100-soluble and -insoluble fractions, we used [35S]methionine labeling followed by chemical cross-linking before cell extraction. Analysis of cross-linked complexes from cells labeled to steady state indicates that, in addition to cadherin/catenin complexes, there were cadherin-independent pools of catenins present in both the TX-100-soluble and -insoluble fractions. Metabolic labeling followed by chase showed that immediately after synthesis, cadherin/beta-catenin, and cadherin/plakoglobin complexes were present in the TX-100-soluble fraction. Approximately 50% of complexes were titrated into the TX-100-insoluble fraction coincident with the arrival of the complexes at the plasma membrane and the assembly of alpha-catenin. Subsequently, > 90% of labeled cadherin, but no additional labeled catenin complexes, entered the TX-100-insoluble fraction.(ABSTRACT TRUNCATED AT 400 WORDS)
Article
Cadherins comprise a family of calcium-dependent glycoproteins that function in mediating cell-cell adhesion in virtually all solid tissues of multicellular organisms. In epithelial cells, E-cadherin represents a key molecule in the establishment and stabilization of cellular junctions. On the cellular level, E-cadherin is concentrated at the adherens junction and interacts homophilically with E-cadherin molecules of adjacent cells. Significant progress has been made in understanding the extra- and intracellular interactions of E-cadherin. Recent success in solving the three-dimensional structure of an extracellular cadherin domain provides a structural basis for understanding the homophilic interaction mechanism and the calcium requirement of cadherins. According to the crystal structure, individual cadherin molecules cooperate to form a linear cell adhesion zipper. The intracellular anchorage of cadherins is regulated by the dynamic association with cytoplasmic proteins, termed catenins. The cytoplasmic domain of E-cadherin is complexed with either β-catenin or plakoglobin (γ-catenin). β-catenin and plakoglobin bind directly to α-catenin, giving rise to two distinct cadherin-catenin complexes (CCC). α-catenin is thought to link both CCC's to actin filaments. The anchorage of cadherins to the cytoskeleton appears to be regulated by tyrosine phosphorylation. Phosphorylation-induced junctional disassembly targets the catenins, indicating that catenins are components of signal transduction pathways. The unexpected association of catenins with the product of the tumor suppressor gene APC has led to the discovery of a second, cadherin-independent catenin complex. Two separate catenin complexes are therefore involved in the cross-talk between cell adhesion and signal transduction. In this review we focus on protein interactions regulating the molecular architecture and function of the CCC. In the light of a fundamental role of the CCC during mammalian development and tissue morphogenesis, we also discuss the phenotypes of embryos lacking E-cadherin or β-catenin. © 1996 Wiley-Liss, Inc.
Article
The complete genomic sequence of the gene responsible for the predominant form of polycystic kidney disease, PKD1, was determined to provide a framework for understanding the biology and evolution of the gene, and to aid in the development of molecular diagnostics. The DNA sequence of a 54 kb interval immediately upstream of the poly(A) addition signal sequence of the PKD1 transcript was determined, and then analyzed using computer methods. A leucine-rich repeat (LRR) motif was identified within the resulting predicted protein sequence of the PKD1 gene. By analogy with other LRR-containing proteins, this may explain some of the disease-related renal alterations such as mislocalization of membrane protein constituents and changes in the extracellular matrix organization. Finally, comparison of the genomic sequence and the published partial cDNA sequence showed several differences between the two sequences. The most significant difference detected predicts a novel carboxy-terminus for the PKD1 gene product.
Article
Autosomal dominant polycystic kidney disease (ADPKD) is a common genetic disorder that frequently results in renal fallure due to progressive cyst development. The major locus, PKD1, maps to 16p13.3. We identified a chromosome translocation associated with ADPKD that disrupts a gene (PBP) encoding a 14 kb transcript in the PKD1 candidate region. Further mutations of the PBP gene were found in PKD1 patients, two deletions (one a de novo event) and a splicing defect, confirming that PBP is the PKD1 gene. This gene is located adjacent to the TSC2 locus in a genomic region that is reiterated more proximally on 16p. The duplicate area encodes three transcripts substantially homologous to the PKD1 transcript. Partial sequence analysis of the PKD1 transcript shows that it encodes a novel protein whose function is at present unknown.
Article
We have studied the expression of the chicken hepatic glycoprotein receptor (chicken hepatic lectin [CHL]) in Madin-Darby canine kidney (MDCK) cells, by transfection of its cDNA under the control of a retroviral promotor. Transfected cell lines stably express 87,000 surface receptors/cell with a kd = 13 nM. In confluent monolayers, approximately 40% of CHL is localized at the plasma membrane. 98% of the surface CHL is expressed at the basolateral surface where it performs polarized endocytosis and degradation of glycoproteins carrying terminal N-acetylglucosamine at a rate of 50,000 ligand molecules/h. Studies of the half-life of metabolically labeled receptor and of the stability of biotinylated cell surface receptor after internalization indicate that transfected CHL performs several rounds of uptake and recycling before it gets degraded. The successful expression of a functional basolateral receptor in MDCK cells opens the way for the characterization of the mechanisms that control targeting and recycling of proteins to the basolateral membrane of epithelial cells.
Article
Autosomal dominant polycystic kidney disease (ADPKD) is a common genetic disorder that frequently results in renal failure due to progressive cyst development. The major locus, PKD1, maps to 16p13.3. We identified a chromosome translocation associated with APPKD that disrupts a gene (PBP) encoding a 14 kb transcript in the PKD1 candidate region. Further mutations of the PBP gene were found in PKD1 patients, two deletions (one a de novo event) and a splicing defect, confirming that PBP is the PKD1 gene. This gene is located adjacent to the TSC2 locus in a genomic region that is reiterated more proximally on 16p. The duplicate area encodes three transcripts substantially homologous to the PKD1 transcript. Partial sequence analysis of the PKD1 transcript shows that it encodes a novel protein whose function is at present unknown.
Article
Cell-cell contact is an important determinant in the formation of functionally distinct plasma membrane domains during the development of epithelial cell polarity. In cultures of Madin-Darby canine kidney (MDCK) epithelial cells, cell-cell contact induces the assembly and accumulation of the Na+,K+-ATPase and elements of the membrane-cytoskeleton (ankyrin and fodrin) at the regions of cell-cell contact. Epithelial cell-cell contact appears to be regulated by the cell adhesion molecule uvomorulin (E-cadherin) which also becomes localized at the lateral plasma membrane of polarized cells. We have sought to determine whether the colocalization of these proteins reflects direct molecular interactions which may play roles in coordinating cell-cell contact and the assembly of the basal-lateral domain of the plasma membrane. Recently, we identified a complex of proteins containing the Na+,K+-ATPase, ankyrin, and fodrin in extracts of whole MDCK cells (Nelson, W.J., and R. W. Hammerton. 1989. J. Cell Biol. 108:893-902). We have now examined cell extracts for protein complexes containing the cell adhesion molecule uvomorulin. Proteins were solubilized from whole MDCK cells and fractionated in sucrose gradients. The sedimentation profile of solubilized uvomorulin is well separated from the majority of cell surface proteins, suggesting that uvomorulin occurs in a protein complex. A distinct portion of uvomorulin (30%) cosediments with ankyrin and fodrin (approximately 10.5S). Further fractionation of cosedimenting proteins in nondenaturing polyacrylamide gels reveals a discrete band of proteins that binds antibodies specific for uvomorulin, Na+,K+-ATPase, ankyrin, and fodrin. Significantly, ankyrin and fodrin, but not Na+K+-ATPase, coimmunoprecipitate in a complex with uvomorulin using uvomorulin antibodies. This result indicates that separate complexes exist containing ankyrin and fodrin with either uvomorulin or Na+,K+-ATPase. These results are discussed in the context of the possible roles of uvomorulin-induced cell-cell contact in the assembly of the membrane-cytoskeleton and associated membrane proteins (e.g., Na+,K+-ATPase) at the contact zone and in the development of cell polarity.
Article
A primary function of cadherins is to regulate cell adhesion. Here, we demonstrate a broader function of cadherins in the differentiation of specialized epithelial cell phenotypes. In situ, the rat retinal pigment epithelium (RPE) forms cell-cell contacts within its monolayer, and at the apical membrane with the neural retina; Na+, K(+)-ATPase and the membrane cytoskeleton are restricted to the apical membrane. In vitro, RPE cells (RPE-J cell line) express an endogenous cadherin, form adherens junctions and a tight monolayer, but Na+,K(+)-ATPase is localized to both apical and basal-lateral membranes. Expression of E-cadherin in RPE-J cells results in restriction and accumulation of both Na+,K(+)-ATPase and the membrane cytoskeleton at the lateral membrane; these changes correlate with the synthesis of a different ankyrin isoform. In contrast to both RPE in situ and RPE-J cells that do not form desmosomes, E-cadherin expression in RPE-J cells induces accumulation of desmoglein mRNA, and assembly of desmosome-keratin complexes at cell-cell contacts. These results demonstrate that cadherins directly affect epithelial cell phenotype by remodeling the distributions of constitutively expressed proteins and by induced accumulation of specific proteins, which together lead to the generation of structurally and functionally distinct epithelial cell types.