ArticlePDF Available

ENaC Activity is Increased in Isolated, Split-Open Cortical Collecting Ducts From Protein Kinase Cα Knockout Mice.

Authors:

Abstract and Figures

ENaC is negatively regulated by protein kinase C (PKC) as shown using PKC activators in a cell culture model. To determine whether PKCα influences ENaC activity in vivo, we examined the regulation of ENaC in renal tubules from PKCα(-/-) mice. Cortical collecting ducts were dissected and split open and the exposed principal cells were subjected to cell-attached patch clamp. In the absence of PKCα, open probability (Po) of ENaC was increased three-fold vs wild-type SV129 mice (0.52 ± 0.04 vs 0.17 ± 0.02). The number of channels per patch was also increased. Using confocal microscopy, we observed an increase in membrane localization of α, β, and γ subunits of ENaC in principal cells in the cortical collecting ducts of PKCα(-/-) mice compared to wild-type mice. To confirm this increase, one kidney from each animal was perfused with biotin and membrane protein was pulled down with streptavidin. The nonbiotinylated kidney was used to assess total protein. While total ENaC protein did not change in PKCα(-/-) mice, membrane localization of all the ENaC subunits was increased. The increase in membrane ENaC could be explained by the observation that ERK1/2 phosphorylation was decreased in the knockout mice. These results imply a reduction in ENaC membrane accumulation and Po by PKCα in vivo. The PKC-mediated increase in ENaC activity was associated with an increase in blood pressure in knockout mice fed a high-salt diet.
ENaC subunits are in closer association with AQP2 in principal cells of KO animals than in WT animals. Kidney slices were stained with rabbit anti-ENaC subunit antibodies and goat anti-AQP2 antibody. Following treatment with appropriate fluorescent secondary antibodies, ENaC subunits (green) and AQP2 (red) were examined by confocal microscopy using an Olympus FV-1000 confocal microscope. The images in the 2 left columns are images from WT animals. The left image of the pair is a composite merged image of the red and green channels. From top to bottom are images for each of the 3 ENaC subunits. The yellow pixels in the composite image show the close association of an ENaC subunit with AQP2. The second column on the left is an image of the same slice as the merged image that was analyzed for colocalization of red and green pixels using a quantitative algorithm (colocalization threshold plugin in the ImageJ program; see METHODS). Pixels that have an intensity for both green and red above the green and red thresholds represent colocalization and are recolored in white. The other areas of the image represent a traditional merge of the green and red channels. Yellow areas may represent additional colocalization, but the intensities in the red and green channels are lower than the white highlighted pixels. Two right columns are the merged image and colocalized image for each of the 3 ENaC subunits for KO mice, respectively. There are significantly more white pixels in slices from KO animals than in WT animals (P 0.001 by z-test; see Table1). Scale bars 5 m in all panels.
… 
Content may be subject to copyright.
ENaC activity is increased in isolated, split-open cortical collecting ducts
from protein kinase Cknockout mice
Hui-Fang Bao,
1,2
Tiffany L. Thai,
1,2
Qiang Yue,
1,2
He-Ping Ma,
1,2
Amity F. Eaton,
1,2
Hui Cai,
1,2,3
Janet D. Klein,
1,2,3
Jeff M. Sands,
1,2,3
and Douglas C. Eaton
1,2
1
Department of Physiology, Emory University School of Medicine, Atlanta, Georgia;
2
The Center for Cell and Molecular
Signaling, Emory University School of Medicine, Atlanta, Georgia; and
3
Department of Medicine, Renal Division, Emory
University School of Medicine, Atlanta, Georgia
Submitted 23 September 2013; accepted in final form 4 December 2013
Bao HF, Thai TL, Yue Q, Ma HP, Eaton AF, Cai H, Klein JD,
Sands JM, Eaton DC. ENaC activity is increased in isolated, split-
open cortical collecting ducts from protein kinase Cknockout mice.
Am J Physiol Renal Physiol 306: F309 –F320, 2014. First published
December 11, 2013; doi:10.1152/ajprenal.00519.2013.—The epithe-
lial Na channel (ENaC) is negatively regulated by protein kinase C
(PKC) as shown using PKC activators in a cell culture model. To
determine whether PKCinfluences ENaC activity in vivo, we
examined the regulation of ENaC in renal tubules from PKC
/
mice. Cortical collecting ducts were dissected and split open, and the
exposed principal cells were subjected to cell-attached patch clamp. In
the absence of PKC, the open probability (P
o
) of ENaC was
increased three-fold vs. wild-type SV129 mice (0.52 0.04 vs.
0.17 0.02). The number of channels per patch was also increased.
Using confocal microscopy, we observed an increase in membrane
localization of -, -, and -subunits of ENaC in principal cells in the
cortical collecting ducts of PKC
/
mice compared with wild-type
mice. To confirm this increase, one kidney from each animal was
perfused with biotin, and membrane protein was pulled down with
streptavidin. The nonbiotinylated kidney was used to assess total
protein. While total ENaC protein did not change in PKC
/
mice,
membrane localization of all the ENaC subunits was increased. The
increase in membrane ENaC could be explained by the observation
that ERK1/2 phosphorylation was decreased in the knockout mice.
These results imply a reduction in ENaC membrane accumulation and
P
o
by PKCin vivo. The PKC-mediated increase in ENaC activity
was associated with an increase in blood pressure in knockout mice
fed a high-salt diet.
protein kinase C; ENaC; renal tubules; single channels; knockout
mice; hypertension
EPITHELIAL NA CHANNELS (ENaC) are sodium-permeable ion
channels located in the apical membrane of polarized epithelial
cells primarily in the distal nephron, lung, and distal colon. In
the distal nephron, ENaC activity is the rate-limiting step for
Na
reabsorption (16, 34); therefore, ENaC activity is critical
for the physiological maintenance of systemic Na
homeosta-
sis and long-term control of blood pressure. Because of its
central role in responding to changes in Na
uptake, ENaC
activity is tightly regulated; dysregulation of this channel has
been linked to abnormal blood pressure in several genetic
disorders including Liddle’s syndrome (18, 37) and pseudohy-
poaldosteronism type 1 (9, 33, 41).
ENaC can be regulated either by altering the amount of time
the channel spends open (open probability or P
o
) or by altering
the density of functional channels (N) in the apical membrane
of distal nephron epithelial cells. One signaling molecule that
appears to alter ENaC activity is protein kinase C (PKC).
Activation of PKC with phorbol esters reduces ENaC activity
in the apical membrane of A6 cells, an amphibian renal cell
line, and in rat principal cells (15). In contrast to the inhibitory
effect on ENaC due to activating PKC, pharmacologically
inhibiting PKC increases ENaC P
o
(23, 49). A6 cells, on which
many of the experiments described above were performed,
contain several different PKC isoforms; so that it is difficult to
determine which isoform is responsible for the changes in
ENaC activity after stimulation or inhibition of all the PKC
isoforms. There are conflicting reports in the literature about
which isoforms of PKC are present in principal cells of the
mouse cortical collecting duct. One report suggested that, in
mice, there was no PKC present in principal cells (29); subse-
quent work suggested that PKCwas present, but no other
typical or novel isoforms (22). Therefore, we made use of
PKCknockout mice to examine PKC regulation of ENaC in
isolated split-open collecting duct principal cells.
METHODS
Animals. PKC
/
mice were initially graciously obtained from
the laboratory of Jeffery Molkentin at the University of Cincinnati
(17) and maintained in house. Control mice were developed from
backcrossing PKC
/
mice to SV129 controls 10 times to obtain
littermate wild-type animals. Once established, the wild-type control
line was maintained and used as controls for the PKCknockouts.
Mice were kept on a 12:12-h light-dark cycle and fed standard
laboratory chow and tap water ad libitum. Mice fed a high-salt diet
were fed standard laboratory chow containing 8% NaCl in place of
standard chow ad libitum for 14 –21 days before euthanasia. All of our
animal protocols and procedures in this paper were approved by the
Emory Institutional Animal Care and Use Committee.
Blood pressure measurements. Systolic blood pressure and heart
rate were measured by tail cuff as previously described (43). Blood
pressures were measured for 5 consecutive days before and during
weeks 1 and 2of a high-salt diet. Data from the first 2 days of each
cycle were discarded as this was considered a transition period in
which the mice become accustomed to the procedure. Between mea-
surement times, mice were allowed to rest for 2 days to avoid
extraordinarily high stress levels. Blood pressures were measured on
a warmed platform (BP-2000, Visitech Systems), and mice were
allowed to rest on the platform for 15 min before measurement. Five
preliminary measurements were made and discarded to accustom
mice to the procedure. Blood pressures and heart rates are an average
of 10 measurements each day.
Address for reprint requests and other correspondence: D. C. Eaton, Emory
Univ. School of Medicine, Dept. of Physiology, Whitehead Biomedical
Research Bldg., 615 Michael St., Atlanta, GA 30322 (e-mail: deaton
@emory.edu).
Am J Physiol Renal Physiol 306: F309–F320, 2014.
First published December 11, 2013; doi:10.1152/ajprenal.00519.2013.
1931-857X/14 Copyright ©2014 the American Physiological Societyhttp://www.ajprenal.org F309
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
SDS-PAGE and immunoblotting. Freshly isolated kidneys were
minced and washed once with PBS and then homogenized using an
Omni TH homogenizer (Warrenton, VA) in tissue protein extraction
reagent (TPER; Thermo Scientific), both solutions containing protease
and phosphatase inhibitors (Thermo Scientific). Tissue lysates were
centrifuged at 1,000 rpm at 4°C for 10 min to remove debris. The
supernatant was then centrifuged at 18,000 gfor6htosediment a
total membrane fraction; this pellet was suspended in a 150-l lysis
buffer. Protein concentration was calculated for cell and tissue lysates
using the BCA protein assay (Thermo Scientific). Forty micrograms
of total protein prepared in Laemmli sample buffer (Bio-Rad, Hercu-
les, CA) was loaded and resolved on Bio-Rad Any KD gradient gels
using the Criterion or Protean electrophoresis systems (Bio-Rad). The
separated proteins were electrically transferred onto Immobilon-P
transfer membranes (Millipore, Billerica, MA). The membranes were
blocked in 5% wt/vol milk in TBST (Bio-Rad) at room temperature
for 1 h. The membranes were washed once with TBST and then
incubated with primary antibodies at a dilution of 1:1,000 in 5%
wt/vol milk in TBST at 4°C for 8 h. The membranes were washed
three times with TBST for 5-min intervals before being incubated
with horseradish peroxidase-conjugated goat anti-rabbit secondary
antibody at a dilution of 1:5,000 in blocking solution. The membranes
were incubated with SuperSignal Dura Chemiluminescent Substrate
for 5 min before being developed using a Kodak Gel Logic 2200
Imager and Molecular Imaging software (Carestream Health, Roch-
ester, NY). This method was used to detect ENaC subunits (with
in-house antibodies) (1, 4, 8, 40, 46), ERK1/2 (9102, Cell Signaling)
and phosphoERK1/2 (9101a, Cell Signaling). PKC isoforms were
detected with antibodies obtained from Cell Signaling, (In particular,
PKCwas detected with Cell Signaling no. 9375.)
Antibody production. Restricted segments of the (H554-N643)-
and -C terminus (D566-N647) were subcloned into the pGEX
expression vector. A segment of the -extracellular domain
250
KIGFQ....SNLWMS
347
from a rat was subcloned into a
maltose-binding-protein vector. The constructs were transformed
into competent bacterial cells, induced with IPTG for expression,
and batch purified from inclusion bodies using glutathione Sephar-
ose 4B (2, 3) or an amylose column. A peptide corresponding to
599
CVDNPI...RIQSAF
647
from the Xenopus -subunit was synthe-
sized. The subunit-specific antibodies were raised in rabbits against a
synthetic peptide sequence or fusion proteins described above. Poly-
clonal antibodies against the carboxy terminal domain of -ENaC
(ENaC 59) and -ENaC (ENaC 60) and the extracellular domain of
(890)- and the C-terminal domain of (2102) were generated in
White New Zealand rabbits by Bio-Synthesis (Lewisville, TX). Each
batch of serum was supplemented with sodium azide and evaluated
for specificity and cross-reactivity using protein from the wheat germ
in vitro translation system (Promega) and mouse renal tissue lysates.
Single-channel patch clamp. Renal tubules were manually dis-
sected, and the cortical collecting duct was identified by morphology.
Tubules were placed in physiological saline [(in mM) 140 NaCl, 5
KCl, 1 CaCl
2
, and 10 HEPES adjusted to pH 7.4 with NaOH] in a
plastic dish before being split open to reveal the apical surface of the
cells before single-channel patch clamp as previously described for
patch clamp of cells in culture (6, 40, 45, 46, 49). Briefly, a micro-
electrode was filled with physiological buffer solution in which
lithium was substituted for sodium (in mM: 140 LiCl, 5 KCl, 1 CaCl
2
,
and 10 HEPES adjusted to pH 7.4 with NaOH) and lowered to a single
cell before application of a small amount of suction to achieve a 1
Gseal. ENaC were identified by characteristic channel kinetics
(long mean open and closed times 0.5 s) and the current-voltage
relationship of the channel (unit conductance close to 6 pS and a very
positive, 40 mv, reversal potential).
Measurement of plasma aldosterone. Blood samples were taken
from anesthetized mice. Blood samples were extracted with 4
volume of methylene chloride three times before evaporating the
solvent under dry nitrogen. The extracted samples were then prepared
for ELISA according to the manufacturer’s instructions (aldosterone
EIA kit- monoclonal, Cayman Biologicals). Plasma aldosterone was
calculated by comparison with a standard curve prepared from known
concentrations of aldosterone.
In situ biotinylation. For in situ biotinylation, a protocol established
by Frindt and Palmer (12, 14) for biotinylation in rats was modified
for use in mice. Mice were anesthetized by injection of 40 –50 mg/kg
pentobarbital sodium (ip). The abdominal cavity was opened to the
diaphragm, and a butterfly needle was inserted into the abdominal
aorta at the bifurcation of the iliac arteries. The aorta was tied above
the level of the renal arteries, and the left renal vein was cut to allow
exit of perfusate. The mouse was then moved to an ice bath for cold
perfusion. Both kidneys were perfused with PBS for 10 min, after
which the left renal artery and vein were tied and the left kidney was
removed to serve as a nonbiotinylated control. The right renal vein
was then cut, and the right kidney was perfused with PBS containing
0.5 mg/ml sulfosuccinimidyl-2-[biotinamido]ethyl-1,3-dithiopropi-
onate (Pierce) for 15 min, after which a biotin-quenching solution
(PBS in which 25 mM Tris·HCl replaced 25 mM NaCl) was perfused
for 25 min to remove excess biotin. Whole kidneys were homoge-
nized, and protein was extracted in buffer containing 250 mM sucrose
and 10 mM triethanolamine (pH 7.4). The supernatant was then
centrifuged at 18,000 gfor6hat4°Ctosediment a total membrane
fraction; this pellet was suspended in a 150-l lysis buffer. The
membrane fraction was equally loaded on streptavidin beads (high-
capacity neutravadin-agarose resin, Thermo Scientific) and incubated
overnight. Beads were washed with a solution of 150 mM NaCl, 5
mM EDTA, 50 mM Tris, and 1% Triton X-100 (pH 7.4) three times
followed by a high-salt wash (wash buffer with 500 mM NaCl) and
two no-salt washes (10 mM Tris, pH 7.4) to remove unbound protein.
Bound protein was eluted in sample buffer containing 0.5 M DTT.
Biotinylated and unbiotinylated membrane fractions were run on
Bio-Rad Any KD gradient gels and probed with in-house antibodies to
-, -, or -ENaC.
Fluorescent immunohistochemistry. Fluorescent immunohistochem-
istry was performed as previously described (43, 45). Briefly, kidneys
were perfused in situ with PBS followed by 2% paraformaldehyde.
Kidneys were removed, put in 20% sucrose solution at 4°C overnight,
dehydrated, and embedded in paraffin wax. Four-micrometer sections
were made, and slices were rehydrated before addition of in-house
ENaC or aquaporin-2 (AQP2) antibodies or commercially available
antibodies against PKC(SC-8393, Santa Cruz Biotechnology). An-
tibodies were detected using an appropriate fluorescently conjugated
secondary antibody (Molecular Probes).
Colocalization analysis. To determine whether ENaC subunits
colocalize more strongly with AQP2 in knockout mice compared with
wild-type, kidney slices from wild-type and knockout mice were
stained with rabbit primary antibodies to -, -, or -ENaC and goat
anti-AQP2. Following treatment with fluorescent secondary antibod-
ies, ENaC (green) and AQP2 (red) were examined using confocal
microscopy on an Olympus Fluoview 1000 confocal microscope. To
determine colocalization, we first determined that there was no fluo-
rescence bleed through from the green channel to the red channel or
vice versa. Then we merged images from the green channel and the
red channel; yellow in the merged image was an indication of
colocalization. To produce a more quantitative measure of the colo-
calization, we used the colocalization algorithms in the image analysis
program ImageJ (11, 28, 35). These algorithms examine the merged
images pixel by pixel for red and green intensities and establish
thresholds for intensity below which there is no significant correlation
of red and green. This represents an unbiased method for determining
thresholds. Pixels that have both red and green intensities above the
thresholds are analyzed and recolored white. In addition, the number
and fraction of pixels with red-green colocalization were calculated
and red-green and green-red correlation coefficients were calculated
(Manders coefficients, m
1
and m
2
) according to Eq. 1
F310 PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
m1
is1i·s2i
i(s1i)2m2
is1i·s2i
i(s2i)2(1)
where s1
i
and s2
i
are the red and green intensities of the ith pixel in
the image. Only pixels with significant intensities in both channels
will contribute to the coefficient.
Data analysis and statistics. All data acquisition and analysis were
performed as described previously (45, 47). Data are reported as
means SE. Statistical analysis was performed with SigmaPlot and
SigmaStat software (Jandel Scientific). Differences between groups
were evaluated with one-way ANOVA or for blood pressure with
two-way ANOVA, and results were considered significant if P
0.05.
We used the FIJI variant of the image-analysis program ImageJ to
analyze changes in Western blots. ImageJ calculates the cumulative
sum of pixel values above baseline for specific bands in a blot. We
averaged these cumulative pixel values for at least three experiments
to determine whether there was a significant difference in band
densities.
RESULTS
Channels in isolated, split-open tubules. Figure 1Ashows an
isolated tubule before (top) and after one end of the tubule from
a wild-type SV-129 mouse kidney was split open. A patch
electrode on a principal cell is visible in the bottom left of the
image. Principal cells were identified by their characteristic
morphology in the split-open tubule. Specifically, in Fig. 1A,
principal cells appear in the Hoffman modulation image to be
large, polygonal, or round cells with concave surfaces; inter-
calated cells have asymmetric shapes with convex, but convo-
luted surfaces. Figure 1Bshows single-channel records from
the patch electrode on the cell in Fig. 1A. The currents are
inward with long mean open and mean closed times charac-
teristic of ENaC. Figure 1Cshows the current-voltage relation-
ship for the channel in Fig. 1B. The inward rectification and the
highly positive reversal potential are also characteristic of
ENaC. The conductance of the channel between 100 and 0 mV
was 13.1 1.43 pS, similar to that reported in rat connecting
tubules by Frindt and Palmer (13).
PKC
knockout mice. We used mice in which PKCis
globally knocked out to study the effect of PKCon ENaC
activity in renal principal cells. Western blotting could not
detect PKCin kidney lysates from knockout mice (Fig. 2,
left). Immunohistochemistry showed that SV-129 wild-type
mice ubiquitously express PKCin the kidney including in
AQP2 (a marker for cortical collecting duct principal cells)-
positive cells (Fig. 2, top right). As expected, the knockout
mice have no detectable PKC, and, in particular, have none in
AQP2-positive principal cells (Fig. 2, bottom right). In data not
shown, we blotted renal cortical lysates for at least one isoform
from each of the major PKC types: typical, novel, and atypical.
As expected, we detected low levels of some other PKC
isoforms; but, except for PKC, there was no difference in the
isoforms between wild-type and knockout kidneys, implying
that there was no compensation of other PKC isoforms for the
loss of PKC.
ENaC activity in PKC
knockout mice on a normal-salt diet.
ENaC activity was recorded from cell-attached patches on
principal cells (as in Fig. 1) from wild-type mice or from
PKCknockout mice fed a normal mouse diet (Fig. 3A). The
top two traces are long representative records from tubular
cells from wild-type or knockout tubules. ENaC activity in the
knockout cells is substantially increased above that in the wild-
type. The regions marked 1 and 2 in the top records are
expanded in the two bottom traces, which also show that the
activity of ENaC in knockout mice is substantially higher than
in wild-type mice. Figure 3, BD, summarizes the results from
patches on a large number of tubules from both wild-type and
knockout mice. Wild-type data are from 31 individual patches;
knockout data are from 21 individual patches. The patches
were from 21 cortical collecting ducts isolated from wild-type
-Vp(mV)
-120
-100
-80
-60
-40
-20
3sec
2pA
BA
C
Voltage (-Vpipette mV)
-120 -80 -40 40
Current (pA) -2.5
-2.0
-1.5
-1.0
-0.5
0
0
20
3sec
0.2pA
Fig. 1. Channels in isolated, split-open tubules.
A: an isolated tubule before (top) and after
splitting open of one end of a tubule from a
wild-type SV-129 mouse kidney. A patch elec-
trode on a principal cell is visible in the bottom
left of the image. Principal cells were identified
by their characteristic morphology. B: single-
channel records from the patch electrode on the
cell in A. The currents are inward with long
mean open and mean closed times characteristic
of epithelial sodium channels (ENaC). Note that
the vertical scale has been expanded 10-fold for
the bottom 2 traces to more easily show the small
unitary currents at these potentials. C: current-
voltage relationship for the channel in B. The
inward rectification and the very positive reversal
potential are also characteristic of ENaC. The line
through the data is the best nonlinear least-squares
fit to the Goldman-Hodgkin-Katz equation and
predicts that intracellular sodium is 12 6.6 mM
and principal cell sodium permeability is (6.4
3.38) 10
7
cm/s. The conductance of the chan-
nel between 100 and 0 mV was 13.1 1.43 pS,
similar to that reported in rat connecting tubule by
Frindt and Palmer (13).
F311PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
and 18 cortical collecting ducts from PKCknockout mice.
The cortical collecting ducts were obtained from 14 wild-type
mice and 10 PKCknockout mice. ENaC activity, measured as
NP
o
, the product of the number of channels (N) times P
o
,is
increased more than threefold in knockout tubules compared
with wild-type (Fig. 3B). When we examined the individual
components of activity, we found that P
o
more than doubled
(Fig. 3C). Given prior reports describing the effects of PKC
inhibitors, this increase in P
o
was not too surprising. What was
somewhat surprising to us was that the number of channels per
PKCα,
,
green, AQP2, red
Wild Type
PKCαKO
PKC WT
98
64
PKC KO
98
64
Fig. 2. PKCknockout mice. Left: Western blots of renal cortical lysates from wild-type (WT) and PKCknockout (KO) mice showing that the knockout mice
have no PKC. The expected molecular weight is 80 82 KDa. Twenty-five micrograms of lysate were loaded per well. Right: immunohistochemistry shows
that SV-129 WT mice ubiquitously express PKCin the kidney including in aquaporin-2 (AQP2; a marker for cortical collecting duct principal cells)-positive
cells (top). The bottom panels show mice in which PKCis globally knocked out. As expected, the KO mice have no detectable PKC, and, in particular, have
none in AQP2-positive principal cells. Scale bars (red lines) 5m in all panels.
PKCα
α
KO mice
Po
0.0
0.2
0.4
0.6
0.8
Wildtype
Mice
**
0.0
0.5
1.0
1.5
NPo
(Channel activity)
Wildtype
Mice
**
PKC
α
KO mice
0
1
2
3
N (channel density)
Wildtype
Mice
**
PKC
α
KO mice
B
CD
A
1
2
1
2
Fig. 3. ENaC activity in PKCKO mice.
ENaC activity was recorded from cell-at-
tached patches on principal cells (as in Fig. 1)
from WT mice or PKCKO mice. A:top 2
traces are long representative records from
WT or KO cells (pipette potential is 60
mv). The activity of the KO cell is substan-
tially increased above that of WT. The regions
marked 1 and 2 in the top traces are expanded
10-fold in the bottom traces to emphasize the
difference in the activity of WT and KO cells.
All recordings were made at 60 mV (differ-
ence in potential between the inside of the cell
and the patch pipette. If there is a significant
basal membrane potential, it will add to the
pipette potential). BD: summary of all single-
channel data. The graph in Bshows that ENaC
activity [NP
o
; measured as the number of
channels (N) times the open probability (P
o
)]
increases over 3-fold in the PKCKO mice
compared with WT (P0.001). When the
components of activity are examined individ-
ually, both P
o
(C) and N(D) increase signif-
icantly (P0.01). WT data are from 31
individual patches; KO data are from 21
individual patches. The patches were from 21
cortical collecting ducts isolated from WT
and 18 cortical collecting ducts from PKC
KO. The cortical collecting ducts were ob-
tained from 14 WT mice and 10 PKCKO
mice.
F312 PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
patch also increased substantially (Fig. 3D). Therefore, we
decided to more carefully examine this increase in N.
Knocking out PKC
increases the density of ENaC in or
near the apical membrane of principal cells. An increase in the
number of active channels in the apical membrane could be due
to recruitment of new channels to the apical membrane or the
conversion of otherwise silent channels to active channels. In
the latter case, the number of channels in or near the apical
membrane would not change; in the former case, there should
be more ENaC near the apical membrane of principal cells. To
examine which possibility seemed more likely, we used im-
munohistochemistry and confocal microscopy of kidney slices
to examine the number and localization of ENaC subunits in
the apical membranes of AQP2-positive cells. Figure 4 shows
that, in knockout compared with wild-type animals, -ENaC is
more strongly colocalized with AQP2 (indicated in yellow) in,
or very near, the apical membranes of cells that we presume are
principal cells because of the AQP2 staining. Figures 5 and 6
αENaC, AQP2
Wild Type
PKCαKO
Fig. 4. Membrane -ENaC is increased in the prin-
cipal cells of PKCknockout mice. We prepared
kidney slices from WT and KO before fixing and
treating with AQP2 and -ENaC antibodies. Subse-
quently, we used secondary antibodies that labeled
AQP2 with a ds-Red monomer and -ENaC with
enhanced green fluorescent protein. Left: 4 panels
are, from bottom right, differential interference con-
trast images, AQP2 in red, merged image, and
-ENaC in green. Right: expanded images from the
areas outlined in white in the merged images on the
left.-ENaC appears to more strongly colocalize
(yellow) with AQP2 in the principal cells from
PKCKO mice than in WT mice. Scale bars (red
lines) 5m in all panels.
βENaC, AQP2
Wild Type
PKCαKO
Fig. 5. Membrane -ENaC is increased in the
principal cells of PKCKO mice. We prepared
kidney slices from WT and KO before fixing and
treating with AQP2 and -ENaC antibodies. Sub-
sequently, we used secondary antibodies that la-
beled AQP2 with a ds-Red monomer and -ENaC
with enhanced green fluorescent protein. Left:4
panels are, from bottom right, differential interfer-
ence contrast images, AQP2 in red, merged image,
and -ENaC in green. Right: expanded images
from the areas outlined in white in the merged
images on the left.-ENaC appears to more
strongly colocalize (yellow) with AQP2 in the prin-
cipal cells from PKCKO mice than in WT mice.
Scale bars (red lines) 5m in all panels.
F313PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
show that - and -ENaC are also more strongly associated
with the apical membrane and AQP2 in principal cells from
knockout animals.
Although the yellow color of the merged images in Figs.
4 6 suggests that there is more colocalization of ENaC and
AQP2 in the knockout mice, we wanted a more quantifiable
measure of colocalization. Therefore, we used the image-
analysis program ImageJ to quantify colocalization (as de-
scribed in METHODS). Figure 7 shows, for slices from wild-type
and knockout mice, the merged images for AQP2 and each
ENaC subunit; to the right of the merged images are merged
images in which pixels that have significant red and green
intensities have been recolored white. There is more colocal-
ization in the knockout mice than in wild-type. Table 1 gives
actual values for the red-green and green-red overlap and the
number and percentage of pixels with significant overlap.
Association of the red channel (AQP2) with the green channel
(ENaC subunits) was calculated using the colocalization plugin
of ImageJ. The last two columns show that ENaC is signifi-
cantly more likely to be associated with ENaC in the knockout
mice compared with wild-type (P0.001 for all subunits).
Examination of the red-green and green-red localization shows
that AQP2 is almost always associated with an ENaC subunit,
but that ENaC subunits (particularly in wild-type mice) are less
likely to be associated with AQP2. We interpret these results to
imply that ENaC moves in knockout mice to regions of the cell
near AQP2, i.e., near the apical membrane. The percent values
for knockout vs. wild-type are significantly different (P
0.001 for each subunit by z-test).
Although confocal microscopy makes it appear possible that
ENaC is concentrated in the apical membranes of principal
cells in knockout animals, we could not rule out the possibility
that ENaC subunits were very close to the inner surface of the
apical membranes, but not actually in the membrane as parts of
active channels. Therefore, we adapted an approach previously
used in rats (12, 14) to mice in which we perfused one kidney
with biotin to label proteins on the luminal surface of the
tubules and confirm that ENaC is actually in the apical mem-
brane (Fig. 8A). After precipitating biotin-labeled proteins
from kidney lysates, we resolved the avidin precipitates and
blotted for the three ENaC subunits (Fig. 8B). The quantity of
each of the subunits was increased in the PKCknockout
mice, confirming that indeed ENaC was increased in the apical
membranes of principal cells as we had shown in the confocal
images (Fig. 8C).
Plasma aldosterone in PKC
knockout mice is not signifi-
cantly different from wild-type mice. The increase in P
o
and
channel density could be explained if plasma aldosterone levels
were substantially elevated in PKCknockout mice. There-
fore, we took blood samples from 24 mice (10 wild-type and 7
knockout animals on a normal diet; 3 wild-type and 4 knockout
on a high-salt diet). After sample preparation, plasma aldoste-
rone was determined using a commercially available enzyme-
linked immunoassay kit. As we anticipated, because a normal
mouse diet contains moderate amounts of salt and a high-salt
diet even more, aldosterone levels were low but not signifi-
cantly different in wild-type or knockout mice [normal diet
wild-type 0.57 0.14 nM (n10) vs. knockout 0.76
0.20 nM (n7); high-salt diet wild-type 0.54 0.24 nM
(n3) vs. knockout 0.33 0.21 nM (n4; means SE)].
Therefore, differences in aldosterone concentration cannot ex-
plain the increase in ENaC activity in PKCa knockout mice.
ERK1/2 activity is reduced in PKC
knockout mice. ENaC
density in the apical membrane is controlled by the activity of
the ubiquitin ligase Nedd4-2 (39, 53). The ability of Nedd4-2
to ubiquitinate ENaC and promote internalization is augmented
by PKC-mediated ERK1/2 phosphorylation of ENaC (7).
Therefore, we examined ERK1/2 phosphorylation as a measure
of ERK1/2 activity in wild-type and PKCknockout mice. We
found that the ERK1/2 amount was unchanged, but phosphor-
ylation was significantly reduced in knockout mice (Fig. 9).
γ
γ
ENaC, AQP2
Wild Type
PKCαKO
Fig. 6. Membrane -ENaC is increased in the
principal cells of PKCknockout mice. We pre-
pared kidney slices from WT and KO before fixing
and treating with AQP2 and -ENaC antibodies.
Subsequently, we used secondary antibodies that
labeled AQP2 with a ds-Red monomer and -
ENaC with enhanced green fluorescent protein.
Left: 4 panels are, from bottom right, differential
interference contrast images, AQP2 in red, merged
image, and -ENaC in green. Right: expanded im-
ages from the areas outlined in white in the merged
images on the left.-ENaC appears to more
strongly colocalize (yellow) with AQP2 in the prin-
cipal cells from PKCKO mice than in WT mice.
Scale bars (red lines) 5m in all panels.
F314 PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
This presumably explains the increased channel density of
ENaC in the apical membrane of principal cells.
Blood pressure is increased in PKC
knockout mice. If the
increase in ENaC activity in knockout mice is accompanied by
increased distal nephron sodium transport, then one might
expect that the blood pressure of knockout mice would be
elevated above that in wild-type mice. In particular, challeng-
ing the mice with a high-sodium diet should increase blood
pressure. Figure 10 shows that when blood pressure was
measured by tail cuff (as described in METHODS), wild-type
mice challenged with a high-salt diet (8% NaCl) for 2 wk had
little if any change in blood pressure while blood pressure in
knockout mice increased dramatically.
ENaC activity in PKC
knockout mice on a high-salt diet.
ENaC activity was recorded from cell-attached patches on
principal cells (as in Figs. 1 and 3) from wild-type mice or
from PKCknockout mice fed a high-salt diet (Fig. 11A). The
top two traces are long representative records from tubular
cells from wild-type or knockout tubules. ENaC activity in the
knockout cells is substantially increased above that in the
wild-type. Figure 11, BD, summarizes the results from
patches on a large number of tubules from both wild-type and
knockout mice. Wild-type data are from 33 individual patches;
knockout data are from 42 individual patches. The patches
were from 11 cortical collecting ducts isolated from WT and 15
cortical collecting ducts from PKCKO. The cortical collect-
α
β
γ
merged co-localized overlay merged co-localized overlay
Green = ENaC Red = AQP2 White = Co-localized ENaC and AQP2
Fig. 7. ENaC subunits are in closer association with AQP2 in principal cells of KO animals than in WT animals. Kidney slices were stained with rabbit anti-ENaC
subunit antibodies and goat anti-AQP2 antibody. Following treatment with appropriate fluorescent secondary antibodies, ENaC subunits (green) and AQP2 (red)
were examined by confocal microscopy using an Olympus FV-1000 confocal microscope. The images in the 2 left columns are images from WT animals. The
left image of the pair is a composite merged image of the red and green channels. From top to bottom are images for each of the 3 ENaC subunits. The yellow
pixels in the composite image show the close association of an ENaC subunit with AQP2. The second column on the left is an image of the same slice as the
merged image that was analyzed for colocalization of red and green pixels using a quantitative algorithm (colocalization threshold plugin in the ImageJ program;
see METHODS). Pixels that have an intensity for both green and red above the green and red thresholds represent colocalization and are recolored in white. The
other areas of the image represent a traditional merge of the green and red channels. Yellow areas may represent additional colocalization, but the intensities
in the red and green channels are lower than the white highlighted pixels. Two right columns are the merged image and colocalized image for each of the 3 ENaC
subunits for KO mice, respectively. There are significantly more white pixels in slices from KO animals than in WT animals (P0.001 by z-test; see Table1).
Scale bars 5m in all panels.
Table 1. Colocalization of ENaC subunits with AQP2 in wild-type and knockout mice
Images AQP2 with ENaC ENaC with AQP2 Number of Colocalized Pixels % Colocalized pixels
Manders coefficients
Knockout AQP2 vs. ␣⫺ENaC 0.915 0.849 6,673 15.6%
WT AQP2 vs. ␣⫺ENaC 0.951 0.568 3,196 7.66%
Knockout AQP2 vs. ␤⫺ENaC 0.923 0.801 4,162 9.88%
WT AQP2 vs. ␤⫺ENaC 0.896 0.668 1,099 2.61%
Knockout AQP2 vs. ␥⫺ENaC 0.955 0.802 7,197 13.6%
WT AQP2 vs. ␥⫺ENaC 0.910 0.773 1,345 2.58%
ENaC, epithelial Na channel; AQP2, aquaporin-2; WT, wild-type.
F315PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
ing ducts were obtained from four wild-type mice and four
PKCknockout mice. ENaC activity, measured as NP
o
, the
product of the number of channels (N) times P
o
, is approxi-
mately doubled from 0.170 0.0386 in wild-type to 0.327
0.0566 in knockout (P0.033) (Fig. 11B). When we exam-
ined the individual components of activity, we found that P
o
increased 50% from 0.101 0.0180 to 0.152 0.0186 (P
0.028) (Fig. 11C). The number of channels per patch also
increased by a small but significant amount (Fig. 11D) from
1.35 0.097 to 1.67 0.095 (P0.023). Interestingly, in the
knockout but not the wild-type animals, we observed several
patches with more channels than we could easily measure
(12). This implies that NP
o
,N, and P
o
might have all been
underestimated in the knockout animals.
DISCUSSION
The effect of PKC on distal nephron ENaC has been previ-
ously described in the literature (7, 15, 23, 23, 40, 49). With a
few exceptions, previous reports using cultured cells have
generally shown that PKC activity inhibits ENaC. Interest-
ingly, despite the role of PKC as a protein kinase, ENaC does
not appear to be directly phosphorylated by PKC (48, 49).
Therefore, PKC must act indirectly to phosphorylate one or
more ENaC-regulatory proteins. The purpose of this work was
to investigate which proteins might be modulated by PKC to
alter ENaC activity.
We were particularly interested in the PKCknockout
animals because, while there are several different renal PKC
isoforms, principal cells appear to contain mostly the -iso-
form. Originally, it was reported that there were no PKC
isoforms in principal cells (29). Subsequently, it was shown
that principal cells did contain PKCbut no other conventional
isoforms (or ) or novel isoforms (,ε,,or) (22).
However, there is evidence for at least one atypical isoform ()
(19); however, the atypical isoforms usually are associated
with regulation of nuclear gene expression so that only the
-isoform appears relevant to membrane signaling. Therefore,
we concentrated on PKCsignaling mechanisms. In our work,
we also show (Fig. 2) that PKCis ubiquitously expressed in
the kidney of wild-type mice (including AQP2-positive prin-
Alpha
Beta
Wildtype
Mice
PKCα
α
KO mice
Wildtype
Mice
PKC
α
KO mice
αENaC
68 kDa
βENaC
95 kDa
Gamma
Wildtype
Mice
PKC
α
KO mice
γENaC
67 kDa
Descending
aorta
Perfusion cannula
ligations
0
2
4
6
8alpha
WT PKCαKO
Band Density
(in thousands)
*
0
10
20
30
Band Density
(in thousands)
gamma
WT PKCαKO
*
0
5
10
15
20
25 beta
WT PKCα KO
Band Density
(in thousands)
*
Fig. 8. In situ biotinylation of mouse kidney. A: schematic of the method. Mice were anesthetized by injection of 80 –90 mg/kg pentobarbital sodium (ip). The
abdominal cavity was opened to the diaphragm, and a butterfly needle was inserted into the abdominal aorta at the bifurcation of the iliac arteries. The aorta was
tied above the level of the renal arteries, and the left renal vein was cut to allow exit of the perfusate. Both kidneys were perfused with PBS for 5 min, after
which the left renal artery and vein were tied and the left kidney was removed to serve as a nonbiotinylated control. The right renal vein was then cut, and the
right kidney perfused with PBS containing 0.5 mg/ml sulfosuccinimidyl-2-[biotinamido]ethyl-1,3-dithiopropionate (Pierce) for 5 min, after which a biotin-
quenching solution was perfused for 25 min to remove excess biotin. B: biotinylated ENaC subunits. Whole kidneys were homogenized, and protein was
extracted. A membrane fraction was equally loaded on streptavidin beads and incubated overnight. After washing, protein was eluted with sample buffer and
resolved on gels and detected with ENaC-specific antibodies. The amount of each of the subunits was greater in the PKCKO mice than WT. C: quantification
of the amounts of ENaC subunits. Mean densitometric analysis is shown of 3 typical experiments for each subunit. We used ImageJ to quantify the blots. The
program calculated the cumulative sum of the pixel values above the background for specific bands. Asterisks indicate significant differences in KO density
compared with wild-type (by t-test: -ENaC: P0.026; -ENaC: P0.046; by rank sum test: -ENaC: P0.029).
F316 PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
cipal cells) but is not detectable in knockout mice, confirming
the previous work of Madsen and coworkers (22).
ENaC P
o
is increased in PKCa knockout mice. Previous
work examining the effects of PKC activation or inhibition on
single ENaC all suggested that there was an effect on channel
P
o
: PKC activation decreased P
o
; PKC inhibition increased P
o
(5, 7, 23, 24, 40, 48, 49). Therefore, it was not surprising to us
that in principal cells in which PKC was knocked out P
o
should
be increased. It was interesting, however, that knocking out
only one isoform, PKC, was sufficient to produce an increase
in P
o
as large as any seen in previous work using inhibitors that
inhibited all typical and most novel isoforms of PKC. How-
ever, in retrospect, given that PKCappears to be the only
isoform present in AQP2-positive principal cells, the signifi-
cant effect of knockout on P
o
might be expected. On the other
hand, in other sodium-transporting epithelia that express mul-
tiple PKC isoforms (such as in the lung), the effect of knocking
out only PKCmight have much more complicated effects.
The mechanism by which P
o
is increased likely involves
phosphatidylinositol 4,5-bisphosphate (PIP
2
) interaction with
the channel. That ENaC gating depends upon PIP
2
has been
well known for a long time (20, 25–27, 30, 50, 52). Recently,
however, Alli et al. (1) have shown that the local concentration
of PIP
2
in the membrane is controlled by association with the
apical membrane of a specialized protein, myristoylated ala-
nine-rich C-kinase substrate (MARCKS; or the very similar
MARCKS-related protein). The ability of MARCKS to control
the local concentration of PIP
2
in the membrane near ENaC
depends upon the state of MARCKS phosphorylation: 1) when
dephosphorylated MARCKS associates with the membrane
and membrane PIP
2
concentrations are elevated, which leads to
increased ENaC P
o
; and 2) when phosphorylated MARCKS
leaves the membrane and enters the cytosol, which reduces
PIP
2
concentrations near ENaC, after which ENaC activity
decreases. The primary kinase that phosphorylates MARCKS
is PKC. Thus, in the absence of PKC, MARCKS remains
associated with the membrane and increases PIP
2
concentra-
tions near ENaC and ENaC P
o
increases (as we observed).
ENaC membrane density is increased in PKCa knockout
mice. Besides an increase in P
o
, we also observed an increase
in the membrane density of ENaC. While it might be possible
that increased PIP
2
in the apical membrane alone could stabi-
lize ENaC and allow an accumulation in the membrane, other
mechanisms appear to contribute. Activation of PKC with
phorbol esters is known to reduce ENaC subunit protein in the
membrane (7, 40). It is also known that phosphorylation of
ENaC by the active, phosphorylated form of ERK1/2 does
promote interaction of ENaC with the ubiquitin ligase (7)
Nedd4-2. Nedd4-2 ubiquitination promotes ENaC removal
from the membrane and internalization, reducing the overall
levels of ENaC in the membrane (21, 32). PKC can phosphor-
ylate and activate ERK1/2 (7, 10). We showed that in the
absence of PKCthe active phosphorylated form of ERK1/2
was significantly reduced, thereby leading to reduced ENaC
internalization and to the increase in membrane ENaC we
observed.
Blood pressure is elevated in PKCa knockout mice. One
might expect that if there are substantial increases in distal
nephron sodium reabsorption, there would be a concomitant
increase in blood volume and blood pressure. One previous
study did not observe any significant difference in blood
pressure between wild-type and PKCknockout mice (43).
However, those experiments were done on mice fed a normal-
salt diet, which might not reveal a sodium balance problem.
Even mice with a mutation which produces constitutively
active ENaC do not have noticeably elevated blood pressure in
the absence of a high-salt challenge (31). Therefore, we fed our
mice an 8% sodium chloride diet and did observe a significant
increase in blood pressure in the knockout animals. We con-
Systolic Blood Pressure
(mm Hg)
100
120
140
160
180
Wild Type
PKC KO
*
Fig. 10. Blood pressure is increased in PKCknockout mice. When blood
pressure was measured by tail cuff (as described in METHODS), WT mice
challenged with a high-salt diet (8% NaCl) had little if any change in systolic
blood pressure, while blood pressure in KO mice increased significantly
(marked with asterisk; n4/group). P0.001 for week 2 on a high-salt for
KO vs. WT by 2-way ANOVA.
ERK1/2 phosphoERK1/2
Band Density
(in thousands)
0
10
20
30
40
Wild type
PKC-αKO
Wild type
Wild type
PKC-αKO
PKC-αKO
2/1KREohpsohp2/1KRE
*
Wild type
Wild type
PKC-αKO
PKC-αKO
α-tubulin α-tubulin
Fig. 9. Active ERK is reduced in PKCKO mice. We measured total ERK1/2
and phosphoERK in kidney lysates (each lane represents a separate animal).
We used ImageJ to quantify the blots using a area that included both ERK1 and
ERK2 bands. The program calculated the cumulative sum of the pixel values
above the background for specific bands. The mean values of the densitometry
results from 3 separate experiments show that total ERK is the same (t-test,
P0.59), but phosphoERK is significantly reduced in KO compared with WT
mice (t test, P0.001).
F317PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
firmed using single-channel measurements that ENaC activity
in knockout animals was increased and could, therefore, ac-
count for the high blood pressure.
Figure 12 shows a schematic diagram of PKC signaling in
wild-type and knockout mice. The situation in wild-type mice
is shown in Fig. 12A. PKCis active and phosphorylates
MARCKS protein. When phosphorylated, MARCKS leaves
the membrane and does not sequester and present PIP
2
to
ENaC, causing ENaC P
o
to decrease. Active PKC also phos-
phorylates ERK, which in turn phosphorylates ENaC. ERK
phosphorylation of ENaC promotes Nedd4-2 interaction with
and ubiquitination of ENaC, with subsequent internalization.
This reduces the apical density of ENaC. The situation in
PKCknockout mice is shown in Fig. 12B. PKC is absent, and
MARCKS protein is not phosphorylated. Therefore, MARCKS
remains associated with the membrane and presents PIP
2
to
ENaC to increase ENaC P
o
. ERK is also not phosphorylated so
that ENaC internalization is reduced and apical membrane
ENaC is increased.
A recent genome-wide association study (44) found that
polymorphic changes in only a very limited number of genes
were associated with increases in blood pressure. The strongest
linkage involved polymorphisms in the gene for PKC. The
study provides no information about the specific effects that the
polymorphisms have on PKC activity. We speculate based on
our results that the polymorphisms lead to decreases in PKC
activity. We will be interested in expressing specific PKC
mutations in heterologous expression systems that also express
ENaC to determine the effect of the polymorphisms on ENaC
activity.
Our work has shown an important role for PKCin the
regulation of ENaC with a potential for significant hyperten-
sion in animals or patients with reduced PKC activity. One
therapeutic agent in common use, rapamycin, is a PKC inhib-
PKCα
α
KO mice
Po
Wildtype
Mice
N (channel density)
Wildtype
Mice
PKC
α
KO mice
*
B
CD
A
0.0
0.2
0.4
NPo
(Channel activity)
Wildtype
Mice
PKC
α
KO mice
0.00
0.05
0.10
0.15
0.20
0.0
0.4
0.8
1.2
1.6
2.0
**
C
-Vp = -60mv
PKC-
α
knock out mice
Control mice
C
10sec
1pS
High salt 2 weeks
Fig. 11. ENaC activity from tubules in high-salt
diet mice. ENaC activity was recorded from cell-
attached patches on principal cells (as in Figs. 1
and 3) from WT mice or PKCKO mice. A:top
trace is a representative record from WT, and
bottom trace from a KO cell. The activity of the
KO cell is substantially increased above that of
the WT. All recordings were made at 60 mV
(difference in potential between the inside of the
cell and the patch pipette. If there is a significant
basal membrane potential, it will add to the pi-
pette potential). BD: summary of all single-
channel data The graph in Bshows that ENaC
NP
o
[measured as the number of channels (N)
times the P
o
] increases 50% in the PKCKO
mice compared with WT (P0.033). When the
components of activity are examined individually,
both P
o
(C) and N(D) increase significantly (P
0.03). WT data are from 33 individual patches; KO
data are from 42 individual patches. The patches
were from 11 cortical collecting ducts isolated from
WT and 15 cortical collecting ducts from PKC
KO. The cortical collecting ducts were obtained
from 4 WT mice and 4 PKCKO mice.
β
βγ
α
PKC
Nedd4-2
ENaC
Internalization
(-)
PIP
2
ERK1/2
(+)
(+)
MARCKS
P
P
P
βγ
α
PKC
Nedd4-2
ENaC
Internalization
PIP
2
ERK1/2
MARCKS
Fig. 12. Schematic diagram of PKC signaling in
WT and KO mice. A: situation in WT mice. PKC
is active and phosphorylates myristoylated alanine-
rich C-kinase substrate (MARCKS) protein. When
phosphorylated, MARCKS leaves the membrane
and does not sequester and present phosphatidyl-
inositol 4,5-bisphosphate (PIP
2
) to ENaC, causing
a decrease in ENaC P
o
. Active PKC also phosphor-
ylates ERK, which in turn phosphorylates ENaC.
ERK phosphorylation of ENaC promotes Nedd4-2
interaction with and ubiquitination of ENaC, with
subsequent internalization. This reduces the apical
density of ENaC. B: situation in PKCKO mice.
PKC is absent, and MARCKS protein is not phos-
phorylated. Therefore, MARCKS remains associ-
ated with the membrane and presents PIP
2
to
ENaC to increase ENaC P
o
. ERK is also not
phosphorylated so that ENaC internalization is
reduced and apical membrane ENaC is increased.
F318 PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
itor (49) and does commonly induce hypertension (42). Given
the examination of PKC inhibitors for a variety of clinical
disorders (36, 38, 51), it seems appropriate to understand the
mechanisms by which the inhibitors can produce excessive
renal sodium transport and hypertension.
GRANTS
This work was supported by National Institutes of Health Grants R37
DK037963 to D. C. Eaton, R01 DK89828 to J. M. Sands, R01-DK067110 to
H.-P. Ma, and T32 DK07656 and AHA 13POST16820072 to T. L. Thai.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
AUTHOR CONTRIBUTIONS
Author contributions: H.-F.B., H.-P.M., H.C., J.D.K., J.M.S., and D.C.E.
provided conception and design of research; H.-F.B., T.L.T., and Q.Y. per-
formed experiments; H.-F.B., T.L.T., Q.Y., A.F.E., and D.C.E. analyzed data;
H.-F.B., T.L.T., H.-P.M., J.D.K., J.M.S., and D.C.E. interpreted results of
experiments; H.-F.B., T.L.T., A.F.E., and D.C.E. prepared figures; H.-F.B.,
T.L.T., H.-P.M., A.F.E., H.C., J.D.K., J.M.S., and D.C.E. edited and revised
manuscript; H.-F.B., T.L.T., Q.Y., H.-P.M., A.F.E., H.C., J.D.K., J.M.S., and
D.C.E. approved final version of manuscript; D.C.E. drafted manuscript.
REFERENCES
1. Alli AA, Bao HF, Alli AA, Aldrugh Y, Song JZ, Ma HP, Yu L,
Al-Khalili O, Eaton DC. Phosphatidylinositol phosphate-dependent reg-
ulation of Xenopus ENaC by MARCKS protein. Am J Physiol Renal
Physiol 303: F800 –F811, 2012.
2. Alli AA, Gower WR Jr. The C type natriuretic peptide receptor tethers
AHNAK1 at the plasma membrane to potentiate arachidonic acid-induced
calcium mobilization. Am J Physiol Cell Physiol 297: C1157–C1167,
2009.
3. Alli AA, Gower WR Jr. Molecular approaches to examine the phosphor-
ylation state of the C type natriuretic peptide receptor. J Cell Biochem 110:
985–994, 2010.
4. Alli AA, Song JZ, Al-Khalili O, Bao HF, Ma HP, Alli AA, Eaton DC.
Cathepsin B is secreted apically from Xenopus 2F3 cells and cleaves the
epithelial sodium channel (ENaC) to increase its activity. J Biol Chem
287: 30073–30083, 2012.
5. Awayda MS, Ismailov II, Berdiev BK, Fuller CM, Benos DJ. Protein
kinase regulation of a cloned epithelial Nachannel. J Gen Physiol 108:
49 –65, 1996.
6. Bao HF, Zhang ZR, Liang YY, Ma JJ, Eaton DC, Ma HP. Ceramide
mediates inhibition of the renal epithelial sodium channel by tumor
necrosis factor-through protein kinase C. Am J Physiol Renal Physiol
293: F1178 –F1186, 2007.
7. Booth RE, Stockand JD. Targeted degradation of ENaC in response to
PKC activation of the ERK1/2 cascade. Am J Physiol Renal Physiol 284:
F938 –F947, 2003.
8. Bou Matar RN, Malik B, Wang XH, Martin CF, Eaton DC, Sands JM,
Klein JD. Protein abundance of urea transporters and aquaporin 2 change
differently in nephrotic pair-fed vs. non-pair-fed rats. Am J Physiol Renal
Physiol 302: F1545–F1553, 2012.
9. Chang SS, Grunder S, Hanukoglu A, Rosler A, Mathew PM, Hanu-
koglu I, Schild L, Lu Y, Shimkets RA, Nelson-Williams C, Rossier BC,
Lifton RP. Mutations in subunits of the epithelial sodium channel cause
salt wasting with hyperkalaemic acidosis, pseudohypoaldosteronism type
1. Nat Genet 12: 248 –253, 1996.
10. Corbit KC, Trakul N, Eves EM, Diaz B, Marshall M, Rosner MR.
Activation of Raf-1 signaling by protein kinase C through a mechanism
involving Raf kinase inhibitory protein. J Biol Chem 278: 13061–13068,
2003.
11. Dunn KW, Kamocka MM, McDonald JH. A practical guide to evalu-
ating colocalization in biological microscopy. Am J Physiol Cell Physiol
300: C723–C742, 2011.
12. Frindt G, Ergonul Z, Palmer LG. Surface expression of epithelial Na
channel protein in rat kidney. J Gen Physiol 131: 617–627, 2008.
13. Frindt G, Palmer LG. Na channels in the rat connecting tubule. Am J
Physiol Renal Physiol 286: F669 –F674, 2004.
14. Frindt G, Palmer LG. Surface expression of sodium channels and
transporters in rat kidney: effects of dietary sodium. Am J Physiol Renal
Physiol 297: F1249 –F1255, 2009.
15. Frindt G, Palmer LG, Windhager EE. Feedback regulation of Na
channels in rat CCT. IV. Mediation by activation of protein kinase C. Am
J Physiol Renal Fluid Electrolyte Physiol 270: F371–F376, 1996.
16. Garty H, Palmer LG. Epithelial sodium channels: function, structure, and
regulation. Physiol Rev 77: 359 –396, 1997.
17. Hambleton M, Hahn H, Pleger ST, Kuhn MC, Klevitsky R, Carr AN,
Kimball TF, Hewett TE, Dorn GW, Koch WJ, Molkentin JD. Phar-
macological- and gene therapy-based inhibition of protein kinase Calpha/
beta enhances cardiac contractility and attenuates heart failure. Circulation
114: 574 –582, 2006.
18. Hansson JH, Nelson-Williams C, Suzuki H, Schild L, Shimkets RA,
Lu Y, Canessa CM, Iwasaki T, Rossier BC, Lifton RP. Hypertension
caused by a truncated epithelial sodium channel gamma subunit: genetic
heterogeneity of Liddle syndrome. Nat Genet 11: 76 –82, 1995.
19. Hao CM, Breyer RM, Davis LS, Breyer MD. Intrarenal distribution of
rabbit PKC zeta. Kidney Int 51: 1831–1837, 1997.
20. Helms MN, Liu L, Liang YY, Al-Khalili O, Vandewalle A, Saxena S,
Eaton DC, Ma HP. Phosphatidylinositol 3,4,5-trisphosphate mediates
aldosterone stimulation of epithelial sodium channel (ENaC) and interacts
with gamma-ENaC. J Biol Chem 280: 40885–40891, 2005.
21. Kabra R, Knight KK, Zhou R, Snyder PM. Nedd4 –2 induces endocy-
tosis and degradation of proteolytically cleaved epithelial Na
channels. J
Biol Chem 283: 6033–6039, 2008.
22. Kim WY, Jung JH, Park EY, Yang CW, Kim H, Nielsen S, Madsen
KM, Kim J. Expression of protein kinase C isoenzymes alpha, betaI, and
delta in subtypes of intercalated cells of mouse kidney. Am J Physiol Renal
Physiol 291: F1052–F1060, 2006.
23. Ling BN, Eaton DC. Effects of luminal Na
on single Na
channels in
A6 cells, a regulatory role for protein kinase C. Am J Physiol Renal Fluid
Electrolyte Physiol 256: F1094 –F1103, 1989.
24. Ling BN, Kokko KE, Eaton DC. Inhibition of apical Nachannels in
rabbit cortical collecting tubules by basolateral prostaglandin E2 is mod-
ulated by protein kinase C. J Clin Invest 90: 1328 –1334, 1992.
25. Ma HP, Chou CF, Wei SP, Eaton DC. Regulation of the epithelial
sodium channel by phosphatidylinositides: experiments, implications, and
speculations. Pflügers Arch 455: 169 –180, 2007.
26. Ma HP, Eaton DC. Acute regulation of epithelial sodium channel by
anionic phospholipids. J Am Soc Nephrol 16: 3182–3187, 2005.
27. Ma HP, Saxena S, Warnock DG. Anionic phospholipids regulate native
and expressed ENaC. J Biol Chem 277: 7641–7644, 2002.
28. Manders EMM, Vereek FJ, Aten JA. Measurement of co-localization of
objects in dual-colour confocal images. J Microsc 169: 375–382, 1993.
29. Pfaff IL, Wagner HJ, Vallon V. Immunolocalization of protein kinase C
isoenzymes alpha, beta1 and betaII in rat kidney. J Am Soc Nephrol 10:
1861–1873, 1999.
30. Pochynyuk O, Tong Q, Staruschenko A, Ma HP, Stockand JD.
Regulation of the epithelial Na
channel (ENaC) by phosphatidylinositi-
des. Am J Physiol Renal Physiol 290: F949 –F957, 2006.
31. Pradervand S, Wang Q, Burnier M, Beermann F, Horisberger JD,
Hummler E, Rossier BC. A mouse model for Liddle’s syndrome. JAm
Soc Nephrol 10: 2527–2533, 1999.
32. Raikwar NS, Thomas CP. Nedd4 –2 isoforms ubiquitinate individual
epithelial sodium channel subunits and reduce surface expression and
function of the epithelial sodium channel. Am J Physiol Renal Physiol 294:
F1157–F1165, 2008.
33. Schild L. The ENaC channel as the primary determinant of two human
diseases: Liddle syndrome and pseudohypoaldosteronism. Nephrologie
17: 395–400, 1996.
34. Schild L. The epithelial sodium channel: from molecule to disease. Rev
Physiol Biochem Pharmacol 151: 93–107, 2004.
35. Schneider CA, Rasband WS, Eliceiri KW. NIH Image to ImageJ: 25
years of image analysis. Nat Methods 9: 671–675, 2012.
36. Shen GX. Selective protein kinase C inhibitors and their applications.
Curr Drug Targets Cardiovasc Haematol Disord 3: 301–307, 2003.
37. Shimkets RA, Warnock DG, Bositis CM, Nelson-Williams C, Hansson
JH, Schambelan M, Gill JR Jr, Ulick S, Milora RV, Findling JW,
Canessa CM, Rossier BC, Lifton RP. Liddle’s syndrome: heritable
human hypertension caused by mutations in the B subunit of the epithelial
sodium channel. Cell 79: 407–414, 1994.
38. Skvara H, Dawid M, Kleyn E, Wolff B, Meingassner JG, Knight H,
Dumortier T, Kopp T, Fallahi N, Stary G, Burkhart C, Grenet O,
F319PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
Wagner J, Hijazi Y, Morris RE, McGeown C, Rordorf C, Griffiths
CE, Stingl G, Jung T. The PKC inhibitor AEB071 may be a therapeutic
option for psoriasis. J Clin Invest 118: 3151–3159, 2008.
39. Snyder PM. Minireview: regulation of epithelial Na
channel trafficking.
Endocrinology 146: 5079 –5085, 2005.
40. Stockand JD, Bao HF, Schenck J, Malik B, Middleton P, Schlanger
LE, Eaton DC. Differential effects of protein kinase C on the levels of
epithelial Na
channel subunit proteins. J Biol Chem 275: 25760–25765,
2000.
41. Strautnieks SS, Thompson RJ, Hanukoglu A, Dillon MJ, Hanukoglu
I, Kuhnle U, Seckl J, Gardiner RM, Chung E. Localisation of pseudo-
hypoaldosteronism genes to chromosome 16p12.2–1311 and 12p131-pter
by homozygosity mapping. Hum Mol Genet 5: 293–299, 1996.
42. Sundaram V, Abraham G, Sundaram V, Mathew M, Bhaskar S,
Ponnusamy J, Prabu S. Rapamycin-induced hypokalaemic nephropathy
in a middle-aged hypertensive male. Nephrol Dial Transplant 22: 1798,
2007.
43. Thai TL, Blount MA, Klein JD, Sands JM. Lack of protein kinase C-
leads to impaired urine concentrating ability and decreased aquaporin-2 in
angiotensin II-induced hypertension. Am J Physiol Renal Physiol 303:
F37–F44, 2012.
44. Turner ST, Boerwinkle E, O’Connell JR, Bailey KR, Gong Y, Chap-
man AB, McDonough CW, Beitelshees AL, Schwartz GL, Gums JG,
Padmanabhan S, Hiltunen TP, Citterio L, Donner KM, Hedner T,
Lanzani C, Melander O, Saarela J, Ripatti S, Wahlstrand B, Manunta
P, Kontula K, Dominiczak AF, Cooper-DeHoff RM, Johnson JA.
Genomic association analysis of common variants influencing antihyper-
tensive response to hydrochlorothiazide. Hypertension 62: 391–397, 2013.
45. Yu L, Bao HF, Self JL, Eaton DC, Helms MN. Aldosterone-induced
increases in superoxide production counters nitric oxide inhibition of
epithelial Na channel activity in A6 distal nephron cells. Am J Physiol
Renal Physiol 293: F1666 –F1677, 2007.
46. Yu L, Cai H, Yue Q, Alli AA, Wang D, Al-Khalili O, Bao HF, Eaton
DC. WNK4 inhibition of ENaC is independent of Nedd4-2-mediated
ENaC ubiquitination. Am J Physiol Renal Physiol 305: F31–F41, 2013.
47. Yu L, Eaton DC, Helms MN. Effect of divalent heavy metals on
epithelial Na
channels in A6 cells. Am J Physiol Renal Physiol 293:
F236 –F244, 2007.
48. Yue G, Eaton DC. Protein kinase C does not affect epithelial Na
channel
activity in inside-out patches from A6 cells (Abstract). FASEB J 12: A179.
1998.
49. Yue G, Edinger RS, Bao HF, Johnson JP, Eaton DC. The effect of
rapamycin on single ENaC channel activity and phosphorylation in A6
cells. Am J Physiol Cell Physiol 279: C81–C88, 2000.
50. Yue G, Malik B, Yue G, Eaton DC. Phosphatidylinositol 4,5-bisphos-
phate (PIP2) stimulates epithelial sodium channel activity in A6 cells. J
Biol Chem 277: 11965–11969, 2002.
51. Zarate CA, Manji HK. Protein kinase C inhibitors: rationale for use and
potential in the treatment of bipolar disorder. CNS Drugs 23: 569 –582,
2009.
52. Zhang ZR, Chou CF, Wang J, Liang YY, Ma HP. Anionic phospho-
lipids differentially regulate the epithelial sodium channel (ENaC) by
interacting with alpha, beta, and gamma ENaC subunits. Pflügers Arch
459: 377–387, 2010.
53. Zhou R, Patel SV, Snyder PM. Nedd4 –2 catalyzes ubiquitination and
degradation of cell surface ENaC. J Biol Chem 282: 20207–20212, 2007.
F320 PKCREGULATES ENaC
AJP-Renal Physiol doi:10.1152/ajprenal.00519.2013 www.ajprenal.org
Downloaded from www.physiology.org/journal/ajprenal (191.101.215.042) on July 22, 2019.
... With these values, the mean time between collision of PIP 2 and ENaC would be 6.3 ϫ 10 2 s, or approximately once in 10 min. In fact, in principal cells, ENaC channels open every 1 or 2 s in a typical patch (18). The implication of these observations is that there must be a mechanism by which the local concentration of PIP 2 could be increased close enough to ENaC to account for the apparently anonymously high opening rate. ...
... The open probability of ENaC is higher in isolated, split-open tubules from PKC␣ knockout than wild-type mice. This increase in ENaC activity is associated with a substantial increase in the blood pressure of knockout mice (18), reinforcing the idea that MARCKS and MLP-1 are relevant physiologically. ...
Article
Epithelial Na ⁺ channels (ENaCs) are members of a family of cation channels that function as sensors of the extracellular environment. ENaCs are activated by specific proteases in the biosynthetic pathway and at the cell surface that remove imbedded inhibitory tracts, which allows channels to transition to higher open probability states. Resolved structures of ENaC and an acid sensing ion channel revealed highly organized extracellular regions. Within the periphery of ENaC subunits are unique domains formed by anti-parallel ß strands containing the inhibitory tracts and protease cleavage sites. ENaCs are inhibited by Na ⁺ binding to specific extracellular site(s) that promotes channel transition to a lower open probability state. Specific inositol phospholipids and channel modification by cys-palmitoylation enhance channel open probability. How these regulatory factors interact in a concerted manner to influence channel open probability is an important question that has not been resolved. These various factors are reviewed, and the impact of specific factors on human disorders is discussed.
... It must be noted that other mechanisms for ENaC inhibition by SARS-CoV2 have been proposed. For instance, it has been proposed that activation of Protein Kinase C (PKC), which is a known inhibitor of ENaC [33][34][35][36], may be responsible for channel inhibition by the S Fig 4. Schematic representation of the proposed mechanism by which S protein expression may affect ENaC activity. In healthy alveoli (left), ENaC in type II pneumocytes contributes to fluid reabsorption from the alveoli to allow adequate conditions for gas exchange to occur. ...
Article
Full-text available
Severe cases of COVID-19 are characterized by development of acute respiratory distress syndrome (ARDS). Water accumulation in the lungs is thought to occur as consequence of an exaggerated inflammatory response. A possible mechanism could involve decreased activity of the epithelial Na ⁺ channel, ENaC, expressed in type II pneumocytes. Reduced transepithelial Na ⁺ reabsorption could contribute to lung edema due to reduced alveolar fluid clearance. This hypothesis is based on the observation of the presence of a novel furin cleavage site in the S protein of SARS-CoV-2 that is identical to the furin cleavage site present in the alpha subunit of ENaC. Proteolytic processing of αENaC by furin-like proteases is essential for channel activity. Thus, competition between S protein and αENaC for furin-mediated cleavage in SARS-CoV-2-infected cells may negatively affect channel activity. Here we present experimental evidence showing that coexpression of the S protein with ENaC in a cellular model reduces channel activity. In addition, we show that bidirectional competition for cleavage by furin-like proteases occurs between 〈ENaC and S protein. In transgenic mice sensitive to lethal SARS-CoV-2, however, a significant decrease in gamma ENaC expression was not observed by immunostaining of lungs infected as shown by SARS-CoV2 nucleoprotein staining.
... Here, we show various forms of DAG (Figure 4) as well as PKC activity ( Figure 5) is augmented in hAoEC treated with the NPRC agonist cANF compared to vehicle. PKC is known to inhibit renal ENaC activity [39], and presumably, EnNaC is inhibited by a similar mechanism in hAoEC. Here, we show pharmacological inhibition of PKC augments EnNaC protein levels in human aortic endothelial cells and this effect is partially blocked by the NPRC agonist cANF ( Figure 6). ...
Article
Full-text available
The C-type natriuretic peptide receptor (NPRC) is expressed in many cell types and binds all natriuretic peptides with high affinity. Ligand binding results in the activation or inhibition of various intracellular signaling pathways. Although NPRC ligand binding has been shown to regulate various ion channels, the regulation of endothelial sodium channel (EnNaC) activity by NPRC activation has not been studied. The objective of this study was to investigate mechanisms of EnNaC regulation associated with NPRC activation in human aortic endothelial cells (hAoEC). EnNaC protein expression and activity was attenuated after treating hAoEC with the NPRC agonist cANF compared to vehicle, as demonstrated by Western blotting and patch clamping studies, respectively. NPRC knockdown studies using siRNA’s corroborated the specificity of EnNaC regulation by NPRC activation mediated by ligand binding. The concentration of multiple diacylglycerols (DAG) and the activity of protein kinase C (PKC) was augmented after treating hAoEC with cANF compared to vehicle, suggesting EnNaC activity is down-regulated upon NPRC ligand binding in a DAG-PKC dependent manner. The reciprocal cross-talk between NPRC activation and EnNaC inhibition represents a feedback mechanism that presumably is involved in the regulation of endothelial function and aortic stiffness.
... Cell-attached patch clamp was used to assess ENaC activity in isolated, split-open rat CCDs, as previously described (Bao et al., 2014;Wu et al., 2019). Principal cells were identified by their characteristic morphology in the split-open tubule. ...
Article
Full-text available
The use of cyclosporine A (CsA) in transplant recipients is limited due to its side effects of causing severe hypertension. We have previously shown that CsA increases the activity of the epithelial sodium channel (ENaC) in cultured distal nephron cells. However, it remains unknown whether ENaC mediates CsA-induced hypertension and how we could prevent hypertension. Our data show that the open probability of ENaC in principal cells of split-open cortical collecting ducts was significantly increased after treatment of rats with CsA; the increase was attenuated by lovastatin. Moreover, CsA also elevated the levels of intracellular cholesterol (Cho), intracellular reactive oxygen species (ROS) via activation of NADPH oxidase p47 phox , serum- and glucocorticoid-induced kinase isoform 1 (Sgk1), and phosphorylated neural precursor cell–expressed developmentally downregulated protein 4–2 ( p -Nedd4-2) in the kidney cortex. Lovastatin also abolished CsA-induced elevation of α-, ß -, and γ-ENaC expressions. CsA elevated systolic blood pressure in rats; the elevation was completely reversed by lovastatin (an inhibitor of cholesterol synthesis), NaHS (a donor of H 2 S which ameliorated CsA-induced elevation of reactive oxygen species), or amiloride (a potent ENaC blocker). These results suggest that CsA elevates blood pressure by increasing ENaC activity via a signaling cascade associated with elevation of intracellular ROS, activation of Sgk1, and inactivation of Nedd4-2 in an intracellular cholesterol-dependent manner. Our data also show that NaHS ameliorates CsA-induced hypertension by inhibition of oxidative stress.
... This suggests that Cx30 may not be fully blocked in the presence of physiological concentrations of divalent cations on the extracellular side, but may be slightly permeable for Ca 2+ and therefore stimulate Ca 2+ signaling pathways. Previous studies provide evidence that an increased intracellular Ca 2+ concentration downregulates ENaC by complex mechanisms (79,(88)(89)(90)(91)(92)(93). However, as mentioned above, there was an inverse relationship between the extracellular Ca 2+ concentration and the inhibitory effect of Cx30 on ENaC. ...
Article
Full-text available
Mice lacking connexin 30 (Cx30) display increased epithelial sodium channel (ENaC) activity in the distal nephron and develop salt-sensitive hypertension. This indicates a functional link between Cx30 and ENaC which remains incompletely understood. Here, we explore the effect of Cx30 on ENaC function using the Xenopus laevis oocyte expression system. Co-expression of human Cx30 with human αβγENaC significantly reduced ENaC-mediated whole-cell currents. The size of the inhibitory effect on ENaC depended on the expression level of Cx30 and required Cx30 ion channel activity. ENaC inhibition by Cx30 was mainly due to reduced cell surface ENaC expression resulting from enhanced ENaC retrieval without discernible effects on proteolytic channel activation and single-channel properties. ENaC retrieval from the cell surface involves the interaction of the ubiquitin ligase Nedd4-2 with PPPxY-motifs in the C-termini of ENaC. Truncating the C-termini of β- or γENaC significantly reduced the inhibitory effect of Cx30 on ENaC. In contrast, mutating the prolines belonging to the PPPxY-motif in γENaC or co-expressing a dominant-negative Xenopus Nedd4 (xNedd4-CS) did not significantly alter ENaC inhibition by Cx30. Importantly, the inhibitory effect of Cx30 on ENaC was significantly reduced by Pitstop-2, an inhibitor of clathrin-mediated endocytosis, or by mutating putative clathrin adaptor protein 2 (AP-2) recognition motifs (YxxФ) in the C-termini of β- or γ-ENaC. In conclusion, our findings suggest that Cx30 inhibits ENaC by promoting channel retrieval from the plasma membrane via clathrin-dependent endocytosis. Lack of this inhibition may contribute to increased ENaC activity and salt-sensitive hypertension in mice with Cx30 deficiency.
Article
Full-text available
We examined the interaction of a membrane-associated protein, MARCKS-like Protein-1 (MLP-1), and an ion channel, Epithelial Sodium Channel (ENaC), with the anionic lipid, phosphatidylinositol 4, 5-bisphosphate (PIP2). We found that PIP2 strongly activates ENaC in excised, inside-out patches with a half-activating concentration of 21 ± 1.17 µM. We have identified 2 PIP2 binding sites in the N-terminus of ENaC β and γ with a high concentration of basic residues. Normal channel activity requires MLP-1’s strongly positively charged effector domain to electrostatically sequester most of the membrane PIP2 and increase the local concentration of PIP2. Our previous data showed that ENaC covalently binds MLP-1 so PIP2 bound to MLP-1 would be near PIP2 binding sites on the cytosolic N terminal regions of ENaC. We have modified the charge structure of the PIP2 –binding domains of MLP-1 and ENaC and showed that the changes affect membrane localization and ENaC activity in a way consistent with electrostatic theory.
Article
Polyuria is found in patients with spinal cord injury (SCI). However, the underlying cellular and molecular mechanism is unknown. Here, we show that mice had elevated urine following 7 days after T10 contusion. Using multiphoton confocal microscopy, we performed intravital imaging experiments to evaluate water reabsorption in kidney tubules by examining fluorescent intensity in the lumen of the distal tubule from live mice. The data show that SCI significantly reduced the concentrating function of kidney tubules. The reduced water reabsorption appears to be mediated by atrial natriuretic peptide (ANP) because SCI increased the expression levels of both ANP and natriuretic peptide receptor A (NPR-A) in the kidney cortex. Our patch-clamp single-channel recordings from split-open distal tubule show that SCI decreased the activity of the epithelial sodium channel (ENaC). Western blot combined with confocal microscopy data show that the levels of 70 kD γ-ENaC, which is an active isoform due to proteolytic cleavage, were significantly reduced in distal tubule principal cells. An NPR-A inhibitor (A71915) given intravenously eliminated the effects of SCI on ENaC and polyuria. These data together with previous studies suggest that SCI causes polyuria probably by reducing ENaC activity through elevating ANP and NPR-A. Further investigation of the signal transduction pathways may provide useful information for discovering an efficient drug to treat SCI-induced polyuria.
Chapter
Amiloride-sensitive epithelial sodium (Na⁺) channels (ENaC) play a crucial role in Na⁺ transport and fluid reabsorption in the kidney, lung, colon, vasculature, and immune system. ENaC is made up of three homologous subunits, α, β, and γ, in a trimeric structure. The magnitude of ENaC-mediated Na⁺ transport in epithelial cells depends on the average open probability of the channels and the number of channels on the apical surface of epithelial cells. The open probability of ENaC depends upon the association of ENaC with inositol lipid phosphates and association with a collection of chaperone proteins including calmodulin and MARCKS (myristoylated alanine-rich C kinase substrate). MARCKS, when associated with the membrane, sequesters PIP2 and increases open probability and channel stability. The number of channels in the apical membrane, in turn, depends upon a balance between the rate of ENaC insertion and the rate of removal from the apical membrane. To this end, ENaC is associated with an additional collection of proteins, the ENaC regulatory complex (ERC) that contains proteins including SGK1, GILZ1, ERK, and Nedd4-2 that regulate channel stability in the membrane. The ERC under basal conditions contains several molecules including ERK and Nedd4-2 that inhibit ENaC activity. The C-terminal domain of all three subunits is intracellular and contains a proline-rich motif (PPxY) to which one isoform of the ubiquitin ligase, Nedd4-2 (neural-precursor-cell-expressed-developmentally-down-regulated-protein), binds to these domains in ENaC, thereby promoting ubiquitin conjugation of the channel subunits followed by internalization and degradation or recycling. In the presence of aldosterone, several proteins including SGK1, GILZ1, and MARCKS are recruited to the complex and stimulate ENaC activity. In addition, aldosterone promotes insertion of channels from intracellular pools and transcription/translation of new ENaC. Other agents that alter ENaC activity work by modulating PIP2 presentation to the channel or by altering the rate of retrieval, recycling, insertion, or transcription/translation of ENaC.
Article
ENaC gating is regulated by phosphatidylinositol 4,5-bisphosphate (PIP 2 ). The PIP 2 -dependent regulation of ENaC is mediated by the myristoylated alanine-rich protein kinase C substrate-like protein-1 (MLP-1). MLP-1 binds to PIP 2 at the plasma membrane. We examined MLP-1 regulation of ENaC in DCT-15 cells. MLP-1 consists of a positively charged effector domain that sequesters PIP 2 and contains serines that are a PKC targets and controls MLP-1 membrane association; a myristoylation domain promoting membrane association, and a multiple homology 2 domain of unknown function. We constructed several MLP-1 mutants: WT: a full length wild-type protein; S3A: 3 substitutions in the effector domain to prevent phosphorylation; S3D: replaced three serines with aspartates to mimic constitutive phosphorylation; GA: replaced the myristoylation site glycine with alanine so GA could not be myristoylated. Mutants were tagged with N-terminal 3XFLAG or C-terminal m-Cherry or V5. Transfection with MLP mutants modified ENaC activity: S3A activity was highest and S3D lowest; the activity of both was significantly different from WT. In Western blots, when transfected with 3XFLAG tagged MLP-1 mutants, expression of full length MLP-1 at 52 KDa increased in S3A-MLP-1 transfected cells and decreased in S3D-MLP-1 transfected cells. Lower molecular weight bands were detected that correspond to potential protease cleavage products. Confocal imaging shows that mutants localize in different sub-cellular compartments consistent with their preferred location in the membrane or cytosol. Protein kinase C increases phosphorylation of endogenous MLP-1 and reduces ENaC activity. Our results suggest a complicated role for proteolytic processing in MLP-1 regulation of ENaC.
Article
Background: Aldosterone activates the intercalated cell mineralocorticoid receptor, which is enhanced with hypokalemia. Whether this receptor directly regulates the intercalated cell chloride/bicarbonate exchanger pendrin is unclear, as are potassium's role in this response and the receptor's effect on intercalated and principal cell function in the cortical collecting duct (CCD). Methods: We measured CCD chloride absorption, transepithelial voltage, epithelial sodium channel activity, and pendrin abundance and subcellular distribution in wild-type and intercalated cell-specific mineralocorticoid receptor knockout mice. To determine if the receptor directly regulates pendrin, as well as the effect of serum aldosterone and potassium on this response, we measured pendrin label intensity and subcellular distribution in wild-type mice, knockout mice, and receptor-positive and receptor-negative intercalated cells from the same knockout mice. Results: Ablation of the intercalated cell mineralocorticoid receptor in CCDs from aldosterone-treated mice reduced chloride absorption and epithelial sodium channel activity, despite principal cell mineralocorticoid receptor expression in the knockout mice. With high circulating aldosterone, intercalated cell mineralocorticoid receptor gene ablation directly reduced pendrin's relative abundance in the apical membrane region and pendrin abundance per cell whether serum potassium was high or low. Intercalated cell mineralocorticoid receptor ablation blunted, but did not eliminate, aldosterone's effect on pendrin total and apical abundance and subcellular distribution. Conclusions: With high circulating aldosterone, intercalated cell mineralocorticoid receptor ablation reduces chloride absorption in the CCD and indirectly reduces principal cell epithelial sodium channel abundance and function. This receptor directly regulates pendrin's total abundance and its relative abundance in the apical membrane region over a wide range in serum potassium concentration. Aldosterone regulates pendrin through mechanisms both dependent and independent of the IC MR receptor.
Article
Full-text available
Liddle's syndrome (pseudoaldosteronism) is an autosomal dominant form of human hypertension characterized by a constellation of findings suggesting constitutive activation of the amiloride-sensitive distal renal epithelial sodium channel. We demonstrate complete linkage of the gene encoding the β subunit of the epithelial sodium channel to Liddle's syndrome in Liddle's original kindred. Analysis of this gene reveals a premature stop codon that truncates the cytoplasmic carboxyl terminus of the encoded protein in affected subjects. Analysis of subjects with Liddle's syndrome from four additional kindreds demonstrates either premature termination or frameshift mutations in this same carboxy-terminal domain in all four. These findings demonstrate that Liddle's syndrome is caused by mutations in the β subunit of the epithelial sodium channel and have implications for the regulation of this epithelial ion channel as well as blood pressure homeostasis.
Article
Full-text available
Pseudohypoaldosteronism type 1 (PHA1, OMIM 264350) is a rare Mendelian disorder characterised by end-organ unresponsiveness to mineralocorticoids. Most steroid hormone insensitivity syndromes arise from mutations in the corresponding receptor, but available genetic evidence is against involvement of the mineralocorticoid receptor gene, MLR, in PHA1. A complete genome scan for PHA1 genes was under-taken using homozygosity mapping in 11 consan-guineous families. Conclusive evidence of linkage with heterogeneity was obtained with a maximum two-locus admixture lod score of 9.9. The disease locus mapped to chromosome 16p12.2–13.11 in six families and to 12p13.1-pter in the other five families. The two chromosomal regions harbour genes for subunits of the amiloride-sensitive epithelial sodium channel: SCNN1B and SCNN1G on 16p and SCNN1A on 12p. Liddle's syndrome of hypertension and pseudoaldos-teronism has been shown to arise from mutations in SCNN1B and SCNN1G. These results strongly suggest that PHA1 and Liddle's syndrome are allelic variants caused by mutations in genes encoding subunits of this sodium channel. These genes are of broad biological interest both in relation to sodium and water home-ostasis in mammals and by virtue of their homology to the mec genes of Caenorhabditis elegans involved in mechanosensitivity and neuronal degeneration.
Article
Full-text available
Phosphatidylinositol phosphates (PIPs) are known to regulate epithelial sodium channels (ENaC). Lipid binding assays and coimmunoprecipitation showed that the amino-terminal domain of the β- and γ-subunits of Xenopus ENaC can directly bind to phosphatidylinositol 4,5-bisphosphate (PIP(2)), phosphatidylinositol 3,4,5-trisphosphate (PIP(3)), and phosphatidic acid (PA). Similar assays demonstrated various PIPs can bind strongly to a native myristoylated alanine-rich C-kinase substrate (MARCKS), but weakly or not at all to a mutant form of MARCKS. Confocal microscopy demonstrated colocalization between MARCKS and PIP(2). Confocal microscopy also showed that MARCKS redistributes from the apical membrane to the cytoplasm after PMA-induced MARCKS phosphorylation or ionomycin-induced intracellular calcium increases. Fluorescence resonance energy transfer studies revealed ENaC and MARCKS in close proximity in 2F3 cells when PKC activity and intracellular calcium concentrations are low. Transepithelial current measurements from Xenopus 2F3 cells treated with PMA and single-channel patch-clamp studies of Xenopus 2F3 cells treated with a PKC inhibitor altered Xenopus ENaC activity, which suggest an essential role for MARCKS in the regulation of Xenopus ENaC activity.
Article
Full-text available
The epithelial sodium channel (ENaC) plays an important role in regulating sodium balance, extracellular volume, and blood pressure. Evidence suggests the α and γ subunits of ENaC are cleaved during assembly before they are inserted into the apical membranes of epithelial cells, and maximal activity of ENaC depends on cleavage of the extracellular loops of α and γ subunits. Here, we report that Xenopus 2F3 cells apically express the cysteine protease cathepsin B, as indicated by two-dimensional gel electrophoresis and mass spectrometry analysis. Recombinant GST ENaC α, β, and γ subunit fusion proteins were expressed in Escherichia coli and then purified and recovered from bacterial inclusion bodies. In vitro cleavage studies revealed the full-length ENaC α subunit fusion protein was cleaved by active cathepsin B but not the full-length β or γ subunit fusion proteins. Both single channel patch clamp studies and short circuit current experiments show ENaC activity decreases with the application of a cathepsin B inhibitor directly onto the apical side of 2F3 cells. We suggest a role for the proteolytic cleavage of ENaC by cathepsin B, and we suggest two possible mechanisms by which cathepsin B could regulate ENaC. Cathepsin B may cleave ENaC extracellularly after being secreted or intracellularly, while ENaC is present in the Golgi or in recycling endosomes.
Article
Full-text available
For the past 25 years NIH Image and ImageJ software have been pioneers as open tools for the analysis of scientific images. We discuss the origins, challenges and solutions of these two programs, and how their history can serve to advise and inform other software projects.
Article
Bipolar disorder is one of the most severely debilitating of all medical illnesses. For a large number of patients, outcomes are quite poor. The illness results in tremendous suffering for patients and their families and commonly impairs functioning and workplace productivity. Risks of increased morbidity and mortality, unfortunately, are frequent occurrences as well. Until recently, little has been known about the specific molecular and cellular underpinnings of bipolar disorder. Such knowledge is crucial for the prospect of developing specific targeted therapies that are more effective and that have a more rapid onset of action than currently available treatments. Exciting recent data suggest that regulation of certain signalling pathways may be involved in the aetiology of bipolar disorder and that these pathways may be profitably targeted to treat the disorder. In particular, mania is associated with overactive protein kinase C (PKC) intracellular signalling, and recent genome-wide association studies of bipolar disorder have implicated an enzyme that reduces the activation of PKC. Importantly, the current mainstays in the treatment of mania, lithium (a monovalent cation) and valproate (a small fatty acid) indirectly inhibit PKC. In addition, recent clinical studies with the relatively selective PKC inhibitor tamoxifen add support to the relevance of the PKC target in bipolar disorder. Overall, a growing body of work both on a preclinical and clinical level indicates that PKC signalling may play an important role in the pathophysiology and treatment of bipolar disorder. The development of CNS-penetrant PKC inhibitors may have considerable benefit for this devastating illness.
Article
SUMMARYA method to measure the degree of co‐localization of objects in confocal dual‐colour images has been developed. This image analysis produced two coefficients that represent the fraction of co‐localizing objects in each component of a dual‐channel image. The generation of test objects with a Gaussian intensity distribution, at well‐defined positions in both components of dual‐channel images, allowed an accurate investigation of the reliability of the procedure. To do that, the co‐localization coefficients were determined before degrading the image with background, cross‐talk and Poisson noise. These synthesized sources of image deterioration represent sources of deterioration that must be dealt with in practical confocal imaging, namely dark current, non‐specific binding and cross‐reactivity of fluorescent probes, optical cross‐talk and photon noise. The degraded images were restored by filtering and cross‐talk correction. The co‐localization coefficients of the restored images were not significantly different from those of the original undegraded images. Finally, we tested the procedure on images of real biological specimens. The results of these tests correspond with data found in the literature. We conclude that the co‐localization coefficients can provide relevant quantitative information about the positional relation between biological objects or processes.
Article
Recent studies of the distribution of PKC isoenzymes in the mouse kidney demonstrated that PKC-alpha, -beta(I), and -delta are expressed in intercalated cells. The purpose of this study was to identify the intercalated cell subtypes that express the different PKC isoenzymes and determine the location of the PKC isoenzymes within these cells. Adult C57BL/6 mice kidney tissues were processed for multiple-labeling immunohistochemistry. Antibodies against the vacuolar H(+)-ATPase and pendrin were used to identify intercalated cell subtypes, whereas antibodies against calbindin D(28K) and aquaporin-2 (AQP2) were used to identify connecting tubule cells and principal cells of the collecting duct, respectively. Within type A intercalated cells, PKC-delta was highly expressed in the apical part of the cells, whereas immunoreactivity for both PKC-alpha and PKC-beta(I) was weak. Type B intercalated cells exhibited strong expression of PKC-alpha, -beta(I), and -delta. PKC-alpha and -beta(I) were localized throughout the cytoplasm, whereas PKC-delta was restricted to the basal domain. Within non-A-non-B cells, immunoreactivity for both PKC-alpha and PKC-beta(I) was high in intensity and localized diffusely in the cytoplasm, whereas PKC-delta was localized in the apical part of the cells. None of the PKC isoenzymes (PKC-alpha, -beta(I), or -delta) were expressed in the calbindin D(28K)-positive connecting tubule cells. Within AQP2-positive principal cells of the collecting duct, PKC-alpha was expressed on the basolateral plasma membrane, but no significant staining was detected for PKC-beta(I) and -delta. In summary, this study demonstrates distinct and differential expression patterns of PKC-alpha, -beta(I), and -delta in the three subtypes of intercalated cells in the mouse kidney.
Article
To identify novel genes influencing blood pressure response to thiazide diuretic therapy for hypertension, we conducted genome-wide association meta-analyses of ≈1.1 million single-nucleotide polymorphisms in a combined sample of 424 European Americans with primary hypertension treated with hydrochlorothiazide from the Pharmacogenomic Evaluation of Antihypertensive Responses study (n=228) and the Genetic Epidemiology of Responses to Antihypertensive study (n=196). Polymorphisms associated with blood pressure response at P<10(-5) were tested for replication of the associations in independent samples of hydrochlorothiazide-treated European hypertensives. The rs16960228 polymorphism in protein kinase C, α replicated for same-direction association with diastolic blood pressure response in the Nordic Diltiazem study (n=420) and the Genetics of Drug Responsiveness in Essential Hypertension study (n=206), and the combined 4-study meta-analysis P value achieved genome-wide significance (P=3.3×10(-8)). Systolic or diastolic blood pressure responses were consistently greater in carriers of the rs16960228 A allele than in GG homozygotes (>4/4 mm Hg) across study samples. The rs2273359 polymorphism in the GNAS-EDN3 region also replicated for same-direction association with systolic blood pressure response in the Nordic Diltiazem study, and the combined 3-study meta-analysis P value approached genome-wide significance (P=5.5×10(-8)). The findings document clinically important effects of genetic variation at novel loci on blood pressure response to a thiazide diuretic, which may be a basis for individualization of antihypertensive drug therapy and identification of new drug targets.
Article
A serine-threonine protein kinase, WNK4, reduces Na(+) reabsorption and K(+) secretion in the distal convoluted tubule by reducing trafficking of thiazide-sensitive Na-Cl co-transporter to and enhancing renal outer medullary potassium channel retrieval from the apical membrane. Epithelial sodium channels (ENaC) in the distal nephron also play a role in regulating Na(+) reabsorption and are also regulated by WNK4, but the mechanism is unclear. In A6 distal nephron cells, transepithelial current measurement and single channel recording show that WNK4 inhibits ENaC activity. Analysis of the number of channel per patch shows that WNK4 reduces channel number but has no effect on channel open probability. Western blots of apical and total ENaC provide additional evidence that WNK4 reduces apical as well as total ENaC expression. WNK4 enhances ENaC internalization independent of Nedd4-2-mediated ENaC ubiquitination. WNK4 also reduced the amount of ENaC available for recycling, but has no effect on the rate of transepithelial current increase to forskolin. In contrast, Nedd4-2 not only reduced ENaC in the recycling pool but also decreased the rate of increase of current after forskolin. WNK4 associates with wild type as well as Liddle's mutated ENaC, and WNK4 reduces both wild type and mutated ENaC expressed in HEK293 cells.