ArticlePDF Available

Endothelial actions of atrial natriuretic peptide prevent pulmonary hypertension in mice

Authors:
  • Institute for Lung Health, Justus-Liebig-Universität Gießen

Abstract and Figures

The cardiac hormone atrial natriuretic peptide (ANP) regulates systemic and pulmonary arterial blood pressure by activation of its cyclic GMP-producing guanylyl cyclase-A (GC-A) receptor. In the lung, these hypotensive effects were mainly attributed to smooth muscle-mediated vasodilatation. It is unknown whether pulmonary endothelial cells participate in the homeostatic actions of ANP. Therefore, we analyzed GC-A/cGMP signalling in lung endothelial cells and the cause and functional impact of lung endothelial GC-A dysfunction. Western blot and cGMP determinations showed that cultured human and murine pulmonary endothelial cells exhibit prominent GC-A expression and activity which were markedly blunted by hypoxia, a condition known to trigger pulmonary hypertension (PH). To elucidate the consequences of impaired endothelial ANP signalling, we studied mice with genetic endothelial cell-restricted ablation of the GC-A receptor (EC GC-A KO). Notably, EC GC-A KO mice exhibit PH already under resting, normoxic conditions, with enhanced muscularization of small arteries and perivascular infiltration of inflammatory cells. These alterations were aggravated on exposure of mice to chronic hypoxia. Lung endothelial GC-A dysfunction was associated with enhanced expression of angiotensin converting enzyme (ACE) and increased pulmonary levels of Angiotensin II. Angiotensin II/AT1-blockade with losartan reversed pulmonary vascular remodelling and perivascular inflammation of EC GC-A KO mice, and prevented their increment by chronic hypoxia. This experimental study indicates that endothelial effects of ANP are critical to prevent pulmonary vascular remodelling and PH. Chronic endothelial ANP/GC-A dysfunction, e.g. provoked by hypoxia, is associated with activation of the ACE–angiotensin pathway in the lung and PH.
This content is subject to copyright. Terms and conditions apply.
ORIGINAL CONTRIBUTION
Endothelial actions of atrial natriuretic peptide prevent
pulmonary hypertension in mice
Franziska Werner
1
Baktybek Kojonazarov
2,3
Birgit Gaßner
1
Marco Abeßer
1
Kai Schuh
1
Katharina Vo
¨lker
1
Hideo A. Baba
4
Bhola K. Dahal
2,3
Ralph T. Schermuly
2,3
Michaela Kuhn
1
Received: 30 June 2015 / Accepted: 16 February 2016 / Published online: 24 February 2016
ÓThe Author(s) 2016. This article is published with open access at Springerlink.com, corrected publication 2022
Abstract The cardiac hormone atrial natriuretic peptide
(ANP) regulates systemic and pulmonary arterial blood
pressure by activation of its cyclic GMP-producing
guanylyl cyclase-A (GC-A) receptor. In the lung, these
hypotensive effects were mainly attributed to smooth
muscle-mediated vasodilatation. It is unknown whether
pulmonary endothelial cells participate in the homeostatic
actions of ANP. Therefore, we analyzed GC-A/cGMP
signalling in lung endothelial cells and the cause and
functional impact of lung endothelial GC-A dysfunction.
Western blot and cGMP determinations showed that cul-
tured human and murine pulmonary endothelial cells
exhibit prominent GC-A expression and activity which
were markedly blunted by hypoxia, a condition known to
trigger pulmonary hypertension (PH). To elucidate the
consequences of impaired endothelial ANP signalling, we
studied mice with genetic endothelial cell-restricted abla-
tion of the GC-A receptor (EC GC-A KO). Notably, EC
GC-A KO mice exhibit PH already under resting, normoxic
conditions, with enhanced muscularization of small arteries
and perivascular infiltration of inflammatory cells. These
alterations were aggravated on exposure of mice to chronic
hypoxia. Lung endothelial GC-A dysfunction was
associated with enhanced expression of angiotensin con-
verting enzyme (ACE) and increased pulmonary levels of
Angiotensin II. Angiotensin II/AT
1
-blockade with losartan
reversed pulmonary vascular remodelling and perivascular
inflammation of EC GC-A KO mice, and prevented their
increment by chronic hypoxia. This experimental study
indicates that endothelial effects of ANP are critical to
prevent pulmonary vascular remodelling and PH. Chronic
endothelial ANP/GC-A dysfunction, e.g. provoked by
hypoxia, is associated with activation of the ACE–an-
giotensin pathway in the lung and PH.
Keywords Atrial natriuretic peptide Endothelium
Guanylyl cyclase-A Cyclic GMP Pulmonary
hypertension
Introduction
Pulmonary hypertension (PH) is a complex and multifac-
torial disease which leads to overload of the right ventricle
(RV) and right heart failure. Pulmonary vasoconstriction,
endothelial cell (EC) dysfunction, vascular thickening,
inflammation and thrombosis contribute to disease pro-
gression in idiopathic and other forms of PH [3,21].
The cardiac hormone atrial natriuretic peptide (ANP),
via its cyclic GMP (cGMP)-synthesizing transmembrane
guanylyl cyclase A (GC-A) receptor, has critical functions
in the maintenance of systemic arterial blood pressure [6,
42] and also regulates pulmonary arterial blood pressure.
Hence, global inactivation of the genes encoding ANP or
GC-A increased resting pulmonary arterial pressure in
mice [28,29] or the susceptibility to hypoxia-induced PH
[58]. Conversely, infusion of synthetic ANP attenuated
hypoxia-induced experimental PH [56] and lowered
&Michaela Kuhn
michaela.kuhn@mail.uni-wuerzburg.de
1
Physiologisches Institut der Universita
¨tWu
¨rzburg,
Ro
¨ntgenring 9, 97070 Wu
¨rzburg, Germany
2
Department of Internal Medicine, University of Gießen and
Marburg Lung Center (UGMLC), Justus-Liebig University
Gießen, Giessen, Germany
3
German Center for Lung Research, Heidelberg, Germany
4
Institute of Pathology, University Hospital of Essen,
University of Duisburg-Essen, Essen, Germany
123
Basic Res Cardiol (2016) 111:22
DOI 10.1007/s00395-016-0541-x
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
pulmonary pressure in patients with high-altitude disease
[32]. Together, these experimental and clinical studies
indicate that endogenous ANP plays a physiological role in
maintaining pulmonary arterial pressure homeostasis. And,
furthermore, that enhancement of endogenous ANP/GC-A/
cGMP signalling, for instance with drugs inhibiting ANP
or cGMP degradation, may have therapeutical implications
[57].
Pulmonary arterial remodelling in PH involves multiple
vascular (EC and smooth muscle cells (SMC), adventitial
fibroblasts) and nonvascular cell types (leucocytes, mast
cells, platelets) [3,21]. With the exception of platelets and
leucocytes, all these cell types express the GC-A receptor
[30]. Because synthetic ANP prevented acute hypoxia-in-
duced pulmonary vasoconstriction [26,58] and exerted
direct cGMP-mediated anti-proliferative effects in cultured
pulmonary arterial SMCs [24], the protective role of the
ANP/GC-A/cGMP pathway in the lung circulation has
mainly been attributed to its effects on pulmonary SMC.
However, as shown in the present study, the GC-A receptor
is also expressed at high levels in lung EC. Whereas
endothelial dysfunction is central to all forms of PH [3,21],
it is unknown whether this involves impaired ANP/GC-A/
cGMP signalling and how this could contribute to the
progression of this disease. Therefore, the goals of this
study were (1) to analyze the expression and activity of
GC-A in lung endothelial cells and the impact of hypoxia;
(2) to dissect the role of endothelial cells in mediating the
effect of ANP in the chronic regulation of pulmonary
arterial pressure by studying mice with selective disruption
of the GC-A-encoding gene (Npr1) in endothelial cells; and
(3) to elucidate the impact of endothelial ANP/GC-A
dysfunction on EC inflammatory activation as well as the
pulmonary levels of endothelin-1 (ET-1) and Angiotensin
II (Ang II). It is known that these hormones are activated
and contribute to cardiopulmonary remodelling in patients
with PH [21]. On the other hand, it was shown that ANP/
GC-A signalling diminishes endothelial ET-1 secretion
[55] and the (inter)actions of ET-1 and Ang II in the heart
and systemic circulation [19]. However, the relevance of
this functional antagonism between ANP and ET-1/Ang II
expression and action in the pulmonary circulation is
unknown.
Materials and methods
Genetic mouse models
Mice with global (GC-A
-/-
) or endothelial cell-restricted
deletion of the GC-A receptor (GC-A
fl/fl
;Tie2Cre
?/-
:EC
GC-A KO) and their respective control littermates (GC-A
?/?
or GC-A
fl/fl
, with unaltered GC-A expression levels) were
generated and genotyped as described [33,46]. The EC GC-
A KO mice have an unaltered median life span and do not
manifest clinically apparent, macroscopic changes
throughout life (mice were observed until the age of
15 months). All present studies were performed with 2- to
4-month-old mice. The experiments were conducted under
the guidelines on humane use and care of laboratory animals
for biomedical research published by NIH (No. 85-23,
revised 1996 [41]) and they were approved by the local
governmental animal care committee.
Hypoxia-induced pulmonary hypertension in mice
and losartan treatment
Experimental pulmonary hypertension (PH) was induced
by exposure to normobaric hypoxia. EC GC-A KO mice
and littermate controls were placed into a partially venti-
lated plexiglass chamber (Biospherix, New York, USA),
and exposed to chronic hypoxia (F
I
O
2
10 %, 90 % nitro-
gen) for 21 days under normobaric conditions [15]. Age-
matched mice of both genotypes were maintained in room
air and served as normoxic controls. For pharmacological
blockade of the Ang II AT
1
-receptor losartan was admin-
istered via the drinking water (10 mg/kg BW/day) during
3 weeks in normoxia or hypoxia. The concentration of the
drug in water was adjusted for body weight and daily water
intake.
Assessment of right ventricular pressures,
pulmonary vascular remodelling and perivascular
inflammation
Closed-chest right ventricular (RV) pressures were mea-
sured in anesthetized freely breathing mice (0.8–1 %
isoflurane) by insertion of a 1.4 F high-fidelity pressure
catheter (Millar Instruments, Houston, TX, USA) via the
external jugular vein. After these invasive hemodynamic
measurements, the lungs of isoflurane (1 %)-anesthetized
mice were fixed with a 1 % PFA solution through the
trachea at a constant pressure of 20 cmH
2
O. The trachea
was ligated, and the lungs and hearts were immersed in
fixative overnight. After paraffin embedding, 4 lm sections
were taken along the longitudinal lung axis (ten sections
per organ) and immunostained with antibodies against a-
smooth muscle actin (aSMA; Sigma, Munich, Germany;
dilution 1:900) and CD45 (Novus Biological, USA; dilu-
tion 1:20) to analyze the number and wall thickness of
muscularized distal arteries and perivascular leucocyte
infiltration [15,47]. Vessels of 20–70 lm external diameter
were classified as fully muscularized (actin staining [75 %
of the circumference), partially muscularized (actin stain-
ing 25–75 % of the circumference), or nonmuscularized
(\25 %). In each section, the percentage of fully or
22 Page 2 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
partially muscularized arteries was calculated [15].
Perivascular inflammation was assessed in the tissue sec-
tion after staining for CD45 [41]. Images were captured at
409using a Leica DM6000B microscope (Leica Instru-
ments, Nussloch, Germany) fitted with a Leica DFC310FX
digital camera. All blood vessels within lung section
ranging from 20 to 70 lm were analyzed using Leica
QWin software. Positively stained CD45 cells surrounding
the vessels were counted [47].
Morphometric analyses of cardiac hypertrophy
The heart was dissected to separate RV from LV plus sep-
tum (S). RV and LV ?S weights were normalized to tibia
lengths. Formaldehyde-fixed right (RV) and left ventricles
(LV) were embedded in paraffin, and 5 lmsectionswere
stained with hematoxylin eosin, periodic acid Schiff (PAS,
to discriminate cardiomyocyte cell borders) or picrosirius
red for quantification of interstitial collagen fractions. The
mean cross-sectional myocyte diameters were calculated by
measuring 50 (RV) to 100 (LV) longitudinally oriented
myocytes with a centrally located nucleus per specimen [18,
46]. Photomicrographs were evaluated using a computer-
assisted image analysis system (Olympus, Hamburg, Ger-
many), using the analySIS software (SIS), the investigator
being blinded to the genotypes [18,46].
Measurements of systemic arterial blood pressure
and left ventricular hemodynamics
Systemic arterial blood pressure was measured by tail cuff in
awake mice [33,46]. Left ventricular (LV) function was
evaluated in isoflurane-anesthetized by LV catheterization
[18]. A 1.4-F combined micromanometer-tipped conduc-
tance catheter (SPR-839, Millar) was retrogradely advanced
via the right carotid artery, and simultaneous recordings of
LV pressure and volume were performed [18].
Effects of ANP on cyclic GMP content
of microvascular lung endothelial cells (MLEC)
Human microvascular lung (ML) EC were purchased from
Promocell (Heidelberg, Germany). The cells were main-
tained in complete EC growth medium MV2 (Promocell)
and studied at passage 4 and 5. The isolation and culture of
murine MLEC has been described before [46]. Immuno-
cytochemistry with antibodies against the endothelial
marker VE-cadherin demonstrated that after the second
selection, more than 95 % of cultured cells were
endothelial. For the experiments the cells were seeded in
gelatine-coated 6-well (for western blot) or 24-well plates
(for cGMP determinations) and cultured for 48 h before
synchronization in medium containing reduced serum
(1 %) concentration for 24 h. The cells were thereafter
exposed to 24 h hypoxia in a humidified 37 °C chamber
(BioSpherix). The concentration of oxygen was reduced to
1 % by replacement with N
2
, keeping CO
2
constant at 5 %.
Control was defined as 95 % air and 5 % CO
2
. Thereafter
the cells were immediately used for the extraction of
membrane proteins (cell fractionation kit; Nanotools,
Teningen, Germany) and for determination of cGMP
responses to ANP. These steps were performed under
normoxic conditions. For cGMP determinations, MLEC
were pretreated with the phosphodiesterase inhibitor
3-isobutyl-1-methylxanthine (IBMX, 0.5 mmol/L, 15 min;
Sigma) and then incubated with ANP (0.1 nmol/L–1 lmol/
L; Bachem, Bubendorf, Switzerland) for additional 10 min.
The incubation media were rapidly removed and cellular
cGMP was extracted with ice-cold ethanol (70 %, v/v).
After centrifugation (30009g, 5 min, 4 °C), the super-
natants were dried in a speed vacuum concentrator,
resuspended in sodium acetate buffer (50 mmol/L, pH 6.0)
and acetylated, and the cGMP content was determined by
radioimmunoassay [31,48].
Determination of GC-A expression and activity
in murine lung cell membranes
ANP-dependent guanylyl cyclase activity in crude lung cell
membranes was determined as described [48]. Freshly
dissected lungs were homogenized using a Polytron
homogenizer in hepes buffer (HB) [25 mM HEPES (pH
7.4), 50 mM NaCl, 20 % glycerol and protease inhibitor
cocktail from Roche, Mannheim, Germany]. The suspen-
sions were pelleted by centrifugation at 45,000gfor 20 min
at 4 °C. Pellets were resuspended in HB and centrifuged
two more times. To initiate cyclase activity, 40 lg mem-
brane protein was incubated in assay buffer [25 mM/L
HEPES, 4 mM/L MgCl
2
, 1 mM/L IBMX, 2 mM/L ATP,
2 mM/L GTP, 30 mM/L phosphocreatine, 400 lg/mL
creatine phosphokinase (185 units/mg) and 0.5 mg/mL
BSA] at 37 °C, with or without ANP. At 10 min of incu-
bation, the reaction was stopped by addition of ice-cold
ethanol (final concentration 70 % v/v). cGMP content was
determined by radioimmunoassay as described above.
cGMP production was normalized to protein content
(40 lg/sample) and the increase in cGMP content in ANP-
treated samples was compared to parallel vehicle-treated
membrane preparations of the same lung.
Western blotting
Membrane proteins from whole lungs were extracted
(Thermo Scientific, Schwerte, Germany) and subjected to
SDS-PAGE and immunoblotting as described [18]. The
primary antibodies were against GC-A (generated in our
Basic Res Cardiol (2016) 111:22 Page 3 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
laboratory [48]) and b-tubulin or GAPDH (for loading
control; Cell Signaling, Frankfurt/Main, Germany). The
blots were developed using the ECL detection system
(Biozym Scientific GmbH, Hessisch-Oldendorf, Germany)
and results were quantitated by densitometry
(ImageQuant).
Quantitative RT-PCR analysis of angiotensin
converting enzyme (ACE), endothelin-1,
intercellular cell adhesion molecule 1 (ICAM-1),
vascular cell adhesion protein 1 (VCAM-1) and E-
selectin mRNA expression levels
Extraction of mRNA from murine MLEC or peripheral lung
tissue and reverse-transcription were performed as described
using TRIzol reagent (Life Technologies GmbH, Darmstadt,
Germany) and Transcriptor First Strand cDNA synthesis kit
(Roche) [18]. Messenger RNA expression levels were ana-
lyzed by Real Time quantitative PCR with LightCycler
Technology (LC-96; Roche) and FastStart Essential Probes
Master with the following primers and probes (all from
Roche): for ACE, sense: 50-GTGGGTATCCCACTGAAAC
C-30; antisense: 50-CAGAAGGCTCCTGTGTCTGA-30;
and probe 121 (REF: 04693558001); for E Selectin, sense:
50-TCCTCTGGAGAGTGGAGTGC-30; antisense: 50-GGT
GGGTCAAAGCTTCACAT-30; and probe 19 (REF: 04686
926001); ET-1, sense: 50-CTGCTGTTCGTGACTTTCCA-
30, antisense: 50-TCTGCACTCCATTCTCAGCTC-30, and
probe 50 (REF: 04688112001); ICAM-1, sense: 50-CGAAG
CTTCTTTTGCTCTGC-30; antisense: 50-GTCCAGCCGA
GGACCATA-30; and probe 10 (REF: 04685091001);
VCAM-1: sense: 50-TGGTGAAATGGAATCTGAACC-30;
antisense: 50-CCCAGATGGTGGTTTCCTT-30; and probe
34 (REF: 04687671001). 12S ribosomal RNA served as
reference gene [sense: 30-GAAGCTGCCAAGGCCTTAG
A-30; antisense: 50-AACTGCAACCAACCACCTTC-30;
FastStart Essential DNA Green Master (Roche)].
Measurement of lung immunoreactive ET-1
Samples were assayed for ET-1 immunoreactivity with a
specific RIA (Bachem) as described by Aguirre et al. [1].
The peptide was extracted from lung tissue by boiling in
109(wt/vol) 1 mol/L acetic acid for 10 min. The samples
were then chilled and centrifuged at 5000gfor 10 min at
4°C. Aliquots (0.1 mL) of supernatant were applied to
Sep-PakC
18
columns (Waters Corporation, Milford, USA).
The columns were activated by 80 % acetonitrile in 0.1 %
TFA followed by 0.1 % TFA. After the column was slowly
washed with 10 % acetonitrile in 0.1 % TFA, samples were
eluted from the column with 80 % acetonitrile in 0.1 %
TFA into polypropylene tubes and evaporated to dryness in
a centrifugal concentrator. The samples were reconstituted
in RIA buffer and subjected to ET-1 radioimmunoassay
(Bachem) according to the manufacturer’s instructions.
Measurement of lung immunoreactive Angiotensin
II
Ang II from murine lungs was extracted and measured with
a commercial Ang II ELISA (Enzo Life Sciences GmbH,
Lo
¨rrach, Germany) according to the manufacturer’s
instructions.
Measurement of pulmonary bradykinin-9 levels
Bradykinin was measured with an EIA Kit (Phoenix Europe,
Karlsruhe, Germany). Tissue extractions and measurements
were performed according to the manufacturer’s protocol.
Freshly dissected lung samples were boiled in 75 % acetic
acid for 20 min (1 mL/100 mg tissue), homogenized with an
ULTRA-TURRAX, centrifuged (15,000g,30min,4°C)
and the supernatants were extracted with Sep-PakC
18
col-
umns. The eluates were dried, reconstituted in assay buffer
and subjected to Bradykinin EIA. Bradykinin levels were
normalized to protein content (BCA assay).
Statistics
Results are presented as mean ±SEM. Group comparisons
were performed using either unpaired ttest or two-way
ANOVA followed by the multiple-comparison Bonferroni
ttest to assess differences between groups. Pvalues of less
than 0.05 were considered statistically significant. The
individual sample sizes for each set of data (n) are provided
in the figure legends.
Results
GC-A is expressed in lung endothelial cells and is
downregulated by hypoxia
To assess the pulmonary endothelial role of GC-A, first
we tested the effects of ANP on cGMP levels of cultured
human and murine microvascular lung endothelial cells
(MLEC). Treatment with ANP (0.1 nmol/L–1 lmol/L,
10 min) provoked similar concentration-dependent
cGMP increases in both species (Fig. 1a). Accordingly,
western blot analyses revealed high lung GC-A levels in
control mice (Fig. 1b). As also shown, the immunoreac-
tive protein was not detected in lungs from mice with
global GC-A deletion (GC-A
-/-
), demonstrating the
specificity of our antibody [48]. Exposure of murine
22 Page 4 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
MLEC to hypoxia (1 % O
2
, 24 h) significantly attenuated
GC-A expression (Fig. 1c) and the cGMP-responses to
ANP (Fig. 1d). To study whether hypoxia-induced
downregulation of lung GC-A occurs in vivo, we exposed
mice to normobaric hypoxia (F
i
10 % O
2
) for 21 days
[15]. Figure 1e, f shows that pulmonary cell membrane
GC-A expression and activity were significantly impaired
by chronic hypoxia.
Endothelial cells are a main expression site of GC-A
in the lung
To dissect the role of endothelial cells in mediating the
homeostatic effect of ANP on pulmonary arterial pressure,
we studied mice with conditional, endothelial-restricted
disruption of GC-A (EC GC-A KO) and control littermates
[46]. As shown in Fig. 2a, in cultured MLEC isolated from
hypoxianormoxia
GC-A
GAPDH
normoxia hypoxia
0.0
0.2
0.4
0.6
0.8
1.0
1.2
lung GC-A / GAPDH
X-fold vs normoxia
*
F
E
ANP, nM
cGMP synthesis by lung
membranes, x-fold vs PBS
0
5
10
15
20
25
30 normoxia
hypoxia
*
*
D
intracellular cGMP
pmol / 100000 MLEC
normoxia
hypoxia
*
*
*
ANP, nM
0.0
0.2
0.4
0.6
0.8
1.0
A
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4 murine MLEC
intracellular cGMP
pmol/ 100000 MLEC
human MLEC
PBS 10 1000
PBS 0.1 1 10 100
PBS 0.1 1 10 100 1000
PBS 1 10 100
*
**
*
*
*
*
ANP, nM ANP, nM
CTR GC-A
-/-
HEK-293
lung membranes
GC-A
B
*
C
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
MLEC GC-A / -Tubulin
X-fold vs normoxia
GC-A
-Tubulin
hypoxianormoxia
normoxia hypoxia
β
β
Fig. 1 Pulmonary endothelial ANP/GC-A/cGMP signalling is atten-
uated by hypoxia. aEffect of ANP on cGMP content of cultured
human (6 wells per condition; 2 independent experiments) and murine
(15 dishes per condition; 5 experiments) microvascular lung endothe-
lial cells (MLEC, 10 min incubation). bRepresentative immunoblot:
strong GC-A expression (apparent MW is *130 kDa) in cell
membranes prepared from wildtype (CTR) lungs (loading 80 lg/
lane). The immunoreactive signal is abolished in membranes prepared
from mice with global GC-A deletion (GC-A
-/-
). Protein extracts
from GC-A-expressing HEK-293 cells were used as positive control.
c,dIn murine MLEC, hypoxia (1 % O
2
, 24 h downregulates GC-A
expression (c; western blots, 40 lg protein/lane) and ANP-induced
intracellular cGMP synthesis (d) (6 wells from 3 independent
experiments). e,f, In mice, chronic hypoxia (normobaric F
i
O
2
of
10 % during 3 weeks) downregulates pulmonary membrane GC-A
expression (e; western with 40 lg protein/lane) and ANP-stimulated
lung cell membrane GC-A/cGMP activity (n=6). *P\0.05 vs.
normoxia
Basic Res Cardiol (2016) 111:22 Page 5 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
the KO mice GC-A expression and cGMP responses to
ANP were fully abolished, demonstrating efficient
endothelial GC-A deletion. Western blot analyses of
whole-lung protein extracts revealed &60 % reduction of
pulmonary GC-A protein levels in EC GC-A KO mice
(Fig. 2b). Even more, ANP-stimulated GC-A activity in
lung cell membranes was reduced by more than 80 %
(Fig. 2c). As already mentioned, different cell types in the
lung express the GC-A receptor. Considering that EC make
up *30 % of the lung cells [13], our studies of control and
EC GC-A KO mice indicate that endothelia are one main
expression site of GC-A in the lung.
Genetic deletion of endothelial GC-A in mice causes
PH and pulmonary vascular remodelling
To study the impact of endothelial GC-A dysfunction on
pulmonary arterial pressure we compared RV pressures in
anesthetized EC GC-A KO and control littermates. RV
catheterization revealed that EC GC-A KO mice have
increased RV systolic pressures (RVSP; Fig. 3a). This was
accompanied by RV hypertrophy, with enhanced RV
weight/tibia length ratios (Fig. 3b) and greater RV myocyte
diameters (Fig. 3c depicts the mean cross-sectional diam-
eters of RV myocytes with a centrally located nucleus).
Picrosirius red stainings did not reveal signs of RV inter-
stitial fibrosis (Fig. 3d). Together these observations indi-
cate that EC GC-A KO mice have mild but consistent PH
already under normoxic conditions. This phenotype was
independent of age (2- to 8-month-old mice were studied)
and gender. Notably, peak RVSP values in EC GC-A KO
mice nearly reached the levels of mice with global, sys-
temic GC-A deletion (GC-A
-/-
mice [33], see Fig. 3e),
whereas RV hypertrophy was more pronounced in the later
genotype (Fig. 3f). Exposure to chronic hypoxia induced
PH and RV hypertrophy in EC GC-A KO and control lit-
termates, again with greater RVSP and RV hypertrophy in
the former, without significant RV fibrosis (Fig. 3a–d).
However, the absolute increase in mean RVSP in response
to hypoxia was not greater in EC GC-A KO mice than that
in controls (?7.5 vs. 6.6 mmHg, respectively). Hypoxia-
induced hematocrite raises did not differ between geno-
types (controls 0.47 ±0.01 % (normoxia) vs.
0.6 ±0.01 %* (hypoxia); EC GC-A KO 0.46 ±0.01 vs.
0.57 ±0.08 %*; *P\0.05 vs. normoxia).
To investigate the effect of endothelial GC-A ablation
on pulmonary vascular remodelling, the degree of muscu-
larization of peripheral arterioles was analyzed by
immunostainings with anti-a-SMA antibodies [15]. Mor-
phometrical analyses showed an increase in the relative
number of fully and partially muscularized vessels and a
concomitant decrease of nonmuscularized vessels in EC
GC-A KO as compared with control lungs (Fig. 4a). In
addition, immunostainings with anti-CD45 antibodies [47]
revealed mild perivascular leucocyte infiltration (Fig. 4b).
Hypoxia provoked lung vascular remodelling and
perivascular inflammation in control and, significantly
more, in EC GC-A KO mice (Fig. 4a, b). Again, the rela-
tive changes (as compared to normoxia) were similar in
both genotypes.
Pulmonary hypertension in EC GC-A KO mice is
not secondary to left heart disease
In agreement with our previous report [40], EC GC-A KO
mice used in the present study had mild systemic hyper-
tension and subtle LV hypertrophy without fibrosis
(Table 1). The degree of LV hypertrophy was not changed
after hypoxia (Fig. 5). Pressure–volume relationships
(studied by LV catheterization) demonstrated that LV
contractile and relaxation functions of EC GC-A KO mice
were unaltered (Fig. 5). LV end-systolic pressures were
slightly greater, consistent with the mildly enhanced
afterload. As also shown in Fig. 5, hypoxia did not
Control mice
EC GC-A KO mice
cGMP synthesis
x-fold vs PBS
0
5
10
15
20
25
**
PBS 10 1000
ANP, nM
GC-A
β
β
-Tubulin
control KO
0.0
0.4
0.8
1.2
lung GC-A / -tubulin
X-fold vs CTR
*
CB
ANP, nM ANP, nM
A
0
0.4
0.8
1.2
1.6
2.0
intracellular cGMP
pmol/100000 MLEC
GC-A
CTR KO
GAPDH
PBS 0.1 1 10 100 1000 PBS 0.1 1 10 100 1000
*
*
*
*
Fig. 2 Inactivation of GC-A in lung endothelial cells of EC GC-A
KO mice. aEffects of ANP on intracellular cGMP content of MLEC
prepared from EC GC-A KO and control littermates (10 min
incubation; n=6 per genotype). Inset Representative western blot
of GC-A expression in MLEC. bImmunoblot analyses of GC-A
expression levels in whole lung protein extracts prepared from EC
GC-A KO and control mice (n=5). cGuanylyl cyclase activity
assays: ANP-dependent cGMP synthesis by lung cell membranes
prepared from EC GC-A KO and control mice (n=6). *P\0.05 vs.
controls
22 Page 6 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Control mice
EC GC-A KO mice
A
0
10
20
30
40
RVSP
(mm Hg)
*
*
B
100 μm
Controls EC GC-A KO
normoxia
hypoxia
C
RV myocyte diameter
(m)
*
*
normoxia hypoxia
0
4
8
12
16
RV weight / tibia length
(mg / cm)
0
0.4
0.8
1.2
1.6
normoxia hypoxia
**
FE
0
10
20
30
40 *
RVSP
(mm Hg)
GC-A-/-
CTR
RV weight / tibia length
(mg / cm)
*
GC-A-/-
CTR
0
0.4
0.8
1.2
1.6
2.0
D
0.00
RV collagen fraction (%)
normoxia hypoxia
0.02
0.04
0.06
0.08
0.10
50 μm
normoxia
hypoxia
Controls EC GC-A KO
normoxia hypoxia
μ
Fig. 3 Genetic deletion of the endothelial GC-A receptor in mice
causes pulmonary hypertension and right ventricular (RV) hypertro-
phy under normoxic conditions and, more, after chronic hypoxia
(F
i
O
2
10 % during 3 weeks). aElevated RV systolic pressures (SP) in
EC GC-A KO mice compared to respective controls under normoxia
and after hypoxia. b,cRatios of RV weight to tibia length and RV
myocyte diameters (indicated by white lines in longitudinal PAS
stained sections) were increased in EC GC-A KO mice. Hypoxia
further enhanced RV hypertrophy of EC GC-A KO mice. dPicrosirius
red stainings revealed that RV interstitial collagen fractions were not
different between genotypes and conditions (n=8 mice per group);
e,fIncreased RVSP and enhanced RVW/tibia length ratios in mice
with global, systemic GC-A deletion (GC-A
-/-
) compared to
respective controls (CTR) (n=6 mice per genotype studied under
normoxia). *P\0.01 vs. controls;
P\0.01 vs. normoxia
Basic Res Cardiol (2016) 111:22 Page 7 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
influence LV function in control or EC GC-A KO mice. In
addition the lung wet-to-dry weight ratios were equal for
EC GC-A KO and controls (Table 1). Together, these data
indicate that the PH of EC GC-A KO mice is not secondary
to LV dysfunction.
Pulmonary levels of immunoreactive endothelin-1
are not altered in EC GC-A KO mice
To elucidate the mechanism(s) contributing to PH in EC
GC-A KO mice we studied specific ANP-modulated
pathways known to be altered in clinical PH. In particular,
ET-1 levels are upregulated in patients with PH and
endothelin receptor antagonists are used in its treatment
[45]. Synthetic ANP inhibits ET-1 release from cultured
human umbilical venous endothelial cells [55]. Therefrom,
we hypothesized that endothelial GC-A dysfunction leads
to increased lung ET-1 levels which, via the vasocon-
strictory and SMC proliferative actions of this peptide,
could contribute to PH in EC GC-A KO mice. However,
qRT-PCR did not reveal significant differences of the ET-1
mRNA levels in GC-A-deficient MLEC and in lungs from
EC GC-A KO mice in comparison to controls (Fig. 6a).
Even more, pulmonary ET-1 levels did not differ between
genotypes (Fig. 6b).
fully
partially
non
0
20
40
60
80
F P N F P N F P N F P N
Degree of muscularization (%)
Arterioles of 20-70 μm ø
CTR EC GC-A KO
normoxia
*
**
*
*
*
CTR EC GC-A KO
hypoxia
A
Controls KO Controls KO
normoxia hypoxia
100 μm
Control mice
EC GC-A KO mice
0.0
0.2
0.4
0.6
0.8
1.0
1.2
CD45 positive cells
perivascular 20-70 m)
normoxia hypoxia
*
*
B
Controls KO Controls KO
normoxia hypoxia
100 m
μ
μ
Fig. 4 Genetic deletion of the endothelial GC-A receptor in mice
causes pulmonary vascular remodelling together with mild perivas-
cular inflammation under normoxic conditions and, more, after
chronic hypoxia. Lung sections were immunostained for SMC a-actin
or for lymphocyte common antigen (CD45). aQuantification of the
relative numbers of fully (F), partially (P) and non (N) muscularized
arterioles and bof perivascular CD45-positive cells per field
demonstrated enhanced pulmonary vascular remodelling and perivas-
cular inflammatory infiltration in EC GC-A KO mice under normoxia
and after hypoxia (n=8 mice per group). *P\0.01 vs. controls;
P\0.01 vs. normoxia
Table 1 EC GC-A KO mice have subtle systemic arterial hyper-
tension and mild left ventricular (LV) hypertrophy without fibrosis
Controls EC GC-A KO
SBP (mmHg) 118 ±2 133 ±3*
DBP (mmHg) 75 ±382±3*
HR (bpm) 589 ±18 564 ±13
Body weight (g) 25 ±1.5 25 ±1.2
Heart weight (mg) 118 ±4.5 145 ±6*
LV weight/tibia length (mg/cm) 4.78 ±0.19 5.9 ±0.2*
LV myocyte diameter (lm) 12 ±0.4 14.8 ±0.8*
Collagen fraction (%) 0.1 ±0.02 0.11 ±0.03
Lung wet/dry weight 4.4 ±0.04 4.5 ±0.04
Hematocrite (%) 43 ±1.7 42 ±2.4
Systemic systolic (SBP) and diastolic (DBP) arterial blood pressure,
heart rate (HR) (determined by tail cuff), hematocrite and LV mor-
phology (necropsy and histology) of EC GC-A KO and control mice.
n=8, * P\0.05 vs. control littermates
22 Page 8 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Enhanced expression levels of ACE
and of endothelial adhesion molecules in EC GC-A
KO lungs
Experimental and clinical studies indicate that the renin–
angiotensin–aldosterone system (RAAS) is involved in
the pathophysiology of PAH [12,14,35,37,38,43].
Synthetic, exogenous ANP attenuates the expression of
angiotensin converting enzyme (ACE) and counterregu-
lates the cardiovascular effects of Ang II [17,19,27,52].
Thus, we evaluated whether the ACE/Ang II pathway
participates in PH of EC GC-A KO mice. Indeed, qRT-
PCR revealed increased ACE expression in GC-A-defi-
cient MLEC and lungs from EC GC-A KO mice (Fig. 6c).
As also shown, ACE mRNA expression was unaltered in
other tissues from the KO mice such as heart. The direct
effect of the dipeptidyl peptidase ACE is to increase
levels of Ang II and decrease levels of bradykinin. To
follow the hypothesis that increased Ang II together with
decreased local bradykinin levels contribute to PH of EC
GC-A KO mice we determined the lung levels of these
peptides. Indeed, levels of immunoreactive Ang II were
greater in EC GC-A KO lungs (Fig. 6d). Concomitantly,
the levels of bradykinin-9 were attenuated although, due
to high variability, the difference to control lungs did not
reach statistical significance (P=0.08; Fig. 6e). Lastly,
qRT-PCR revealed increased pulmonary expression of the
EC adhesion molecules VCAM-1 and ICAM-1 and mild
not-significant increases of E-selectin in EC GC-A KO
mice (Fig. 6f, g).
Enhanced ACE/Angiotensin II signalling contributes
to PH of EC GC-A KO mice
To study whether increased lung ACE/Ang II levels
contribute to the PH of EC GC-A KO mice, we compared
)nim/lm(tuptuocaidraC)mc/gm(htgnelaibit/thgiewVL
Stroke work (mmHg/μl)
End-systolic pressure (mmHg) End-diastolic pressure (mmHg)
Heart rate (beats/min)
tmax
Ejection fraction (%) dP/d (mmHg/s) -dP/dtmin (mmHg/s)
0
2
4
6
8
10
12
0
500
1000
1500
2000
0
2000
4000
6000
8000
10000
12000
14000
0
2500
5000
7500
10000
12500
15000
0
10
20
30
40
50
60
0
100
200
300
400
500
600
0
2
4
6
8
0
1
2
3
4
5
6
7
normoxia hypoxia
**
0
20
40
60
80
100
120
140
normoxia hypoxia normoxia hypoxia
normoxia hypoxianormoxia hypoxianormoxia hypoxia
normoxia hypoxia normoxia hypoxia
normoxia hypoxia
**
Control mice
EC GC-A KO mice
Fig. 5 Left ventricular (LV) weight and function of anesthetized
control and EC GC-A KO mice determined by pressure–volume
analyses after normoxia or chronic hypoxia. Ratios of LV weight to
tibia length and LV systolic pressures were similarly increased in EC
GC-A KO mice under normoxic and after hypoxic conditions. All
other parameters of LV contraction and relaxation were not different
between genotypes and conditions. n=6 mice per genotype and
condition; *P\0.05 vs. controls
Basic Res Cardiol (2016) 111:22 Page 9 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
theeffectsofchronicblockadeoftheAngII/AT
1
-receptor
in both genotypes. Figure 7illustrates the impact of
losartan treatment (10 mg/kg/day, 3 weeks) on RVSP
(Fig. 7a) and on the ratios of RV weight/tibia length
(Fig. 7b) of mice maintained under normoxia or hypoxia.
As illustrated, losartan did not affect these parameters in
normoxic controls. The drug partly prevented the
increases in RVSP of control mice subjected to hypoxia
(Fig. 7a); however, this did not ameliorate either the
hypertrophy of the RV (Fig. 7b) or the thickening of the
distal pulmonary arteries. The percentage (%) of fully
muscularized distal arteries was: 0.96 ±0.44 in control
mice under normoxia; 6.84 ±1.6* in controls after
hypoxia; and 5.1 ±1.2* in controls treated with losartan
during hypoxia (n=6 mice per group; *P\0.05 vs.
normoxia).
Notably, while losartan had no appreciable effects in
normoxic control mice, it almost reversed the baseline
pulmonary hypertension of EC GC-A KO mice. This is
indicated by the decreases of RVSP (Fig. 7a), RV hyper-
trophy (Fig. 7b), pulmonary vascular remodelling and
perivascular inflammation (Fig. 7c, d). In addition,
administration of losartan during hypoxia partly but sig-
nificantly prevented the hypoxia-driven augmentation of
these cardiovascular changes (Fig. 7a–d). Lastly losartan
also reversed the mild (hypoxia-independent) LV hyper-
trophy of EC GC-A KO mice, as indicated by the following
LV weight/tibia length ratios (in mg/cm): 5.9 ±0.2 (KO,
ET-1 mRNA / S12
x-fold vs. controls
Lung
0
0.2
0.4
0.6
0.8
1.0
1.2
1.4 B
0
0.5
1.0
1.5
Endothelin-1
X-fold vs. controls
A
Control mice
EC GC-A KO mice
EDC
GF
mRNA /S12
x-fold vs. controls
*
VCAM-1 ICAM-1
*
0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
mRNA /S12s
x-fold vs. control
ET-1 mRNA /S12
x-fold vs. controls
MLEC
0
1.0
1.5
2.0
2.5
0.5
E-Selectin
1.8
0
1.0
1.2
1.4
1.6
0.8
0.2
0.4
0.6
Bradykinin
x-fold vs. controls
p = 0.0
6
0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Angiotensin II
x-fold vs. controls
*
0
0.5
1.0
1.5
2.5
2.0
ACE mRNA /S12
x-fold vs. controls
tissues
RV LVLung
*
*
MLEC
0
1.0
2.0
2.5
3.0
0.5
1.5
*
Fig. 6 Unaltered endothelin-1 but altered levels of angiotensin
converting enzyme (ACE), Ang II, bradykinin and EC adhesion
molecules in cultured microvascular lung endothelial cells (MLEC)
and/or in lungs of EC GC-A KO mice. a,bReal-time RT-PCR and
radioimmunoassay (RIA) showed that the endothelial and pulmonary
mRNA and peptide levels of ET-1 were not significantly different
between genotypes. cACE mRNA expression was increased in GC-
A-deficient MLEC and in lungs from EC GC-A KO mice (n=5). d,
ePulmonary levels of immunoreactive Ang II were significantly
greater in EC GC-A KO mice whereas the pulmonary levels of
bradykinin were diminished (P=0.08). fVCAM-1, ICAM-1 and
E-Selectin mRNA levels were increased in lungs from EC GC-A KO
mice. The mRNA levels of all target genes were normalized to the
levels of 12S ribosomal RNA as reference gene. All data are
illustrated as x-fold changes in EC GC-A KO vs. control mice. n=8
per genotype; *P\0.05 vs. controls
22 Page 10 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
normoxia, vehicle); 3.9 ±0.11* (KO, normoxia, losartan);
5.8 ±0.14 (KO, hypoxia, vehicle); 3.8 ±0.27* (KO,
hypoxia, losartan) (n=9 mice per group; *P\0.05 vs.
vehicle). Together these observations indicate that AT
1
receptor signalling has a significant role in the cardiac and
pulmonary remodelling changes of EC GC-A KO mice.
Fig. 7 Blockade of the Ang II/AT
1
-receptor reversed the pulmonary
vascular changes in EC GC-A KO mice. aTreatment of control mice
with losartan (10 mg/kg BW/day during 3 weeks) had no effect on
baseline RVSP (normoxia) but attenuated the increment by chronic
hypoxia. In EC GC-A KO littermates losartan decreased elevated
RVSP under normoxic conditions and attenuated the increment by
chronic hypoxia. bIn control mice losartan did not prevent hypoxia-
induced RV enlargement. However, losartan reversed baseline RV
hypertrophy (under normoxia) in EC GC-A KO littermates and
prevented the increase by hypoxia. cThe number of fully muscular-
ized lung arterioles, and dsurrounding infiltration by CD45-positive
leucocytes in EC GC-A KO mice under normoxia and after hypoxia
were significantly decreased by losartan. n=6–9 mice per group;
*P\0.05 vs. vehicle;
P\0.05 vs. normoxia
Basic Res Cardiol (2016) 111:22 Page 11 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Discussion
Together with previous reports [32,5658], our experi-
mental studies demonstrate that ANP, via its GC-A
receptor, plays an important physiological role in the
moderation of pulmonary arterial pressure and lung vas-
cular remodelling under normoxic and hypoxic conditions.
The major novel findings are (1) EC are a major expression
site of the GC-A receptor in the lung; (2) hypoxia impairs
pulmonary endothelial GC-A expression and signaling; (3)
genetic inactivation of the endothelial GC-A receptor in
mice (EC GC-A KO) provokes PH, pulmonary vascular
remodeling and subtle perivascular inflammatory infiltra-
tion already under normoxic conditions; (4) peak RVSP
values in EC GC-A KO mice were similar to the levels of
mice with deletion of GC-A in all cell types (GC-A
-/-
),
indicating that the endothelial effects of ANP are critically
involved in the chronic moderation of pulmonary arterial
pressure and vascular homeostasis by this hormone, at least
in the murine system; and (5) enhanced local ACE/Ang II
signaling contributes to the pulmonary vascular alterations
in mice with endothelial GC-A dysfunction.
The increases in RVSP and the extent of pulmonary
vascular remodeling in mice with global ANP or GC-A
inactivation [28,29], or endothelial-restricted GC-A abla-
tion are very consistent. In fact, less pronounced and more
variable changes were observed in other disease-relevant
genetic mouse models. For instance, wide ranges of RVSP
were observed in mice with endothelial deletion of the
BMPR2 gene (20.7–56.3 mmHg; median, 27 mmHg)
compared with control mice (19.9–26.7 mmHg; median
23 mmHg), and only a subset of BMPR2-deficient mice
with RVSP [30 mmHg exhibited RV hypertrophy and
pulmonary vascular remodeling [23]. Even more, exposure
of wild type rats or mice to chronic hypoxia (as accepted
experimental model of PH) increases RVSP by
7–10 mmHg [15,28,29]. Hence, in general the functional
and morphological pulmonary alterations in experimental
PH are much less pronounced as in the clinical setting,
emphasizing that patients have a multifactorial disease
whereas experimental studies attempt to dissect the con-
tribution of specific genes or mechanisms. The present
experimental study suggests that endothelial ANP/GC-A
dysfunction could be one aspect of the complex neurohu-
moral imbalance accompanying and aggravating PH, in
particular hypoxia-induced PH in chronic high-altitude
disease. Our observations may stimulate clinical studies to
follow this possibility.
Experimental and clinical studies showed that during
chronic hypoxia, right heart ANP and BNP synthesis and
circulating NP levels increase, possibly in response to the
RV pressure overload provoked by pulmonary
vasoconstriction [911,44]. Because synthetic ANP
counterregulates hypoxic pulmonary vasoconstriction [8,
22,26] and limits the interaction of endothelial and
inflammatory cells [25,39] and the proliferation of cultured
vascular SMC [24], it was proposed that enhanced
endogenous ANP/BNP release helps to mitigate the
development of hypoxic PH [9,10]. However, as shown
here, hypoxia can decrease lung GC-A levels and
endothelial GC-A/cGMP responses to ANP, which will
attenuate these protective ANP (and BNP) effects. The
inhibition of ANP/GC-A signaling by hypoxia has also
been observed in coronary EC [2] but the molecular
mechanism is presently unknown and requires further
study.
Endothelial GC-A dysfunction might cause PH in mice
by provoking chronic increases in pulmonary arteriolar
tone and/or vascular remodelling. Hence, we hypothesized
that ANP physiologically regulates the endothelial release
or (in) activation of factors locally modulating these pro-
cesses, such as ET-1, Ang II or bradykinin. And, con-
versely, that this effect of ANP is abolished in EC GC-A
KO mice. Interestingly, whereas ET-1 mRNA and protein
levels were unaltered, ACE mRNA levels were increased
in GC-A-deficient MLEC and in lungs from EC GC-A KO
mice. Concomitantly, pulmonary Ang II levels were
greater in the mutants whereas bradykinin levels tended to
be diminished. It is well known that Ang II, via AT
1
sig-
nalling, not only causes vasoconstriction, but also migra-
tion and proliferation of SMC as well as recruitment of
inflammatory cells [16,54]. Specifically, inhibition of ACE
decreased the cellular inflammatory response in experi-
mental models of lung inflammation [4]. Indeed, in the
present study AT
1
-receptor blockade with losartan largely
reversed PH, pulmonary vascular remodelling and inflam-
mation in normoxic EC GC-A KO mice. Even more,
losartan significantly attenuated the exacerbation of these
cardiovascular changes in response to hypoxia. Together
these observations indicate that enhanced ACE-dependent
local Ang II formation contributes to these phenotypical
alterations. In line with our results, several experimental
and clinical studies have implicated the involvement of the
RAAS in the pathogenesis of PH [35]. All components,
including renin, angiotensinogen, ACE and both subtypes
of Ang II receptors, are expressed in the lung [38,43].
Increased ACE expression and activity in the endothelium
of peripheral pulmonary arteries have been found in animal
models of PH and, importantly, in patients with various
forms of PAH [38,43]. However, the pathophysiological
mechanism(s) remain(s) unclear. Our studies add a novel
piece of information showing that pulmonary endothelial
ANP/GC-A/cGMP-dysfunction is associated with
enhanced ACE expression and activity. The inhibition of
22 Page 12 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
ACE levels by ANP was also observed by others [52] and
we will try to clarify the mechanism in our future
investigations.
Notably, in the present study losartan did not clearly
ameliorate hypoxic pulmonary hypertension in control mice.
The increase in RVSP was only partly prevented, whereas
RV hypertrophy and lung vascular remodelling were not at
all attenuated by the drug. Hence, mechanisms independent
of the AT
1
receptor seem to predominate. In line with our
observations, blockade of the AT
1
receptor by olmesartan
[49] or genetic deletion of ACE [53] also failed to amelio-
rate hypoxic PH and RV hypertrophy in mice. In contrast,
AT
1
antagonists (GR138950C, olmesartan) reversed
hypoxia-induced cardiopulmonary remodelling in rats [40,
57]. The discrepancy between these results remains unex-
plained; species differences might be involved.
Beside increased Ang II diminished bradykinin levels
may contribute to PH and lung perivascular inflammation
of EC GC-A KO mice. The small nine amino-acid
vasoactive peptide bradykinin has dual roles by exerting
pathophysiological as well as beneficial physiological
effects, mainly by stimulation of bradykinin B2 receptors.
Specifically in the lung, inhibition of bradykinin metabolic
breakdown by ACE inhibitors or exogenous administration
of B2 receptor agonists exerted protective effects, reducing
pulmonary arterial pressure in experimental hypertension
[50] and neutrophil recruitment by lipopolysaccharide [4].
These protective effects of bradykinin involve the
endothelial release of NO, prostacyclin and tissue-type
plasminogen activator [4]. Hence, we hypothesize that PH
and perivascular inflammation in EC GC-A KO mice is
mediated through both a local increase in Ang II and a
decrease in bradykinin mediated signalling.
In general, experimental and clinical studies emphasize
that a compromised endothelial barrier plays a central role
in the pathogenesis of PH [3,45]. In fact, both acute and
chronic hypoxia in mice and rats induce subtle but signif-
icant inflammation in the lung prior to the onset of struc-
tural changes in the vessel wall [34,36]. On the other hand,
numerous studies in vitro/in vivo indicated that ANP exerts
pulmonary endothelial barrier-protecting actions. Synthetic
ANP reduced hypoxia, TNF-a, thrombin, or bacterial
endotoxin (PepG)-induced paracellular hyperpermeability
of pulmonary microvascular and macrovascular endothelial
cells cultured on permeable supports and acute PepG-in-
duced lung injury in mice [30,51]. Conversely, enhanced
PepG-induced lung injury, ICAM-1/VCAM-1 expression
and vascular leak were observed in ANP
-/-
mice [51]. We
did not observe macroscopic signs of pulmonary edema in
EC GC-A KO mice under normoxic or hypoxic conditions.
However, we found increased pulmonary levels of the
endothelial adhesion molecules ICAM-1 and VCAM-1.
Together with the imbalance between Ang II and
bradykinin signalling these changes possibly contribute to
enhanced pulmonary neutrophil infiltration and PH in EC
GC-A KO mice.
Study limitations
One limitation of the EC GC-A KO mice is that the GC-A
receptor is absent not only in pulmonary but also in sys-
temic endothelia. Unfortunately, a selective disruption of
target genes within the pulmonary circulation is technically
impossible so far and therefore this limitation is shared by
other disease relevant genetic mouse models [20]. Hence,
because EC GC-A KO mice have mild systemic arterial
hypertension and subtle LV hypertrophy, it is possible that
PH was secondary to the systemic phenotype. However,
invasive haemodynamic studies clearly demonstrated that
cardiac output and LV function of EC GC-A KO mice are
unaltered, also after chronic hypoxia. In addition there are
no signs of pulmonary edema, corroborating that the PH of
these mice is not secondary to left ventricular failure. Even
more, we did not observe vascular thickening or inflam-
mation in other tissues of EC GC-A KO mice. Together
these observations indicate that the pulmonary vascular
alterations of EC GC-A KO mice are not secondary to
systemic changes.
Concordant to the mice with systemic ANP or GC-A
deletion ([28,29] and present study), mice with EC-re-
stricted GC-A ablation have mild PH already under base-
line, normoxic conditions, which was aggravated by
chronic hypoxia. However, the absolute increases in mean
RVSP and in vascular remodelling in response to CH were
similar in EC GC-A KO mice and in controls. Again this is
consistent with previous observations in mice with global
ANP or GC-A inactivation [28,29]. Hence, it remains
impossible to definitively determine whether ANP/GC-A
dysfunction aggravates hypoxic PH or merely produces
normoxic PH that is then amplified by hypoxia.
Conclusion
In summary, endothelial effects of ANP play a critical
physiological role in the chronic maintenance of pul-
monary vascular homeostasis. Our observations in vitro
and in EC GC-A KO mice suggest that ANP moderates the
endothelial expression (VCAM-1, ICAM-1) or formation
of local factors (ACE/Ang II, possibly bradykinin) regu-
lating SMC proliferation and the interaction of EC and
inflammatory cells. Our experimental observations in a
monogenetic mouse model suggest that chronic endothelial
ANP/GC-A dysfunction, e.g. provoked by hypoxia, might
contribute to lung endothelial barrier impairment and
vascular remodelling, and thereby to PH.
Basic Res Cardiol (2016) 111:22 Page 13 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Acknowledgments This study was supported by the Deutsche
Forschungsgemeinschaft (DFG KU 1037/6-1, to M.K.) and by the
Excellence Cluster Cardio Pulmonary System (ECCPS, to R.S.).
Compliance with ethical standards
Conflict of interest The authors declare that there are no conflicts
of interest.
Open Access This article is distributed under the terms of the
Creative Commons Attribution 4.0 International License (http://crea
tivecommons.org/licenses/by/4.0/), which permits unrestricted use,
distribution, and reproduction in any medium, provided you give
appropriate credit to the original author(s) and the source, provide a
link to the Creative Commons license, and indicate if changes were
made.
References
1. Aguirre JI, Morrell NW, Long L, Clift P, Upton PD, Polak JM,
Wilkins MR (2000) Vascular remodeling and ET-1 expression in
rat strains with different responses to chronic hypoxia. Am J
Physiol Lung Cell Mol Physiol 278:L981–L987
2. Agullo
´L, Garcia-Dorado D, Escalona N, Inserte J, Ruiz-Meana
M, Barrabe
´s JA, Mirabet M, Pina P, Soler-Soler J (2002) Hypoxia
and acidosis impair cGMP synthesis in microvascular coronary
endothelial cells. Am J Physiol 283:H917–H925. doi:10.1152/
ajpheart.01067.2001
3. Archer SL, Weir EK, Wilkins MR (2010) Basic science of pul-
monary arterial hypertension for clinicians: new concepts and
experimental therapies. Circulation 121:2045–2066. doi:10.1161/
CIRCULATIONAHA.108.847707
4. Arndt PG, Young SK, Poch KR, Nick JA, Falk S, Schrier RW,
Worthen GS (2006) Systemic inhibition of the angiotensin-con-
verting enzyme limits lipopolysaccharide-induced lung neu-
trophil recruitment through both bradykinin and angiotensin II-
regulated pathways. J Immunol 177:7233–7241. doi:10.4049/
jimmunol.177.10.7233
5. Baliga RS, Zhao L, Madhani M, Lopez-Torondel B, Visintin C,
Selwood D, Wilkins MR, MacAllister RJ, Hobbs AJ (2008) Syn-
ergy between natriuretic peptides and phosphodiesterase 5 inhibi-
tors ameliorates pulmonary arterial hypertension. Am J Respir Crit
Care Med 178:861–869. doi:10.1164/rccm.200801-121OC
6. Baxter GF (2004) The natriuretic peptides. Basic Res Cardiol
99:71–75. doi:10.1007/s00395-004-0457-8
7. Bubb KJ, Trinder SL, Baliga RS, Patel J, Clapp LH, MacAllister
RJ, Hobbs AJ (2014) Inhibition of phosphodiesterase 2 augments
cGMP and cAMP signaling to ameliorate pulmonary hyperten-
sion. Circulation 130:496–507. doi:10.1161/CIRCULATIO
NAHA.114.009751
8. Cargill RI, Lipworth BJ (1995) Acute effects of ANP and BNP on
hypoxic pulmonary vasoconstriction in humans. Br J Clin Phar-
macol 40:585–590. doi:10.1111/j.1365-2125.1995.tb05803.x
9. Casserly B, Klinger JR (2009) Brain natriuretic peptide in pul-
monary arterial hypertension: biomarker and potential therapeutic
agent. Drug Des Dev Ther 3:269–287. doi:10.2147/DDDT.S4805
10. Casserly B, Pietras L, Schuyler J, Wang R, Hill NS, Klinger JR
(2010) Cardiac atria are the primary source of ANP release in
hypoxia-adapted rats. Life Sci 87:382–389. doi:10.1016/j.lfs.
2010.07.013
11. Chun YS, Hyun JY, Kwak YG, Kim IS, Kim CH, Choi E, Kim
MS, Park JW (2003) Hypoxic activation of the atrial natriuretic
peptide gene promoter through direct and indirect actions of
hypoxia-inducible factor-1. Biochem J 370:149–157. doi:10.
1042/BJ20021087
12. Clozel JP, Saunier C, Hartemann D, Fischli W (1991) Effects of
cilazapril, a novel angiotensin converting enzyme inhibitor, on
the structure of pulmonary arteries of rats exposed to chronic
hypoxia. J Cardiovasc Pharmacol 17:36–40
13. Crapo JD, Barry BE, Gehr P, Bachofen M, Weibel ER (1982)
Cell number and cell characteristics of the normal human lung.
Am Rev Respir Dis 126:332–337
14. de Man FS, Tu L, Handoko ML, Rain S, Ruiter G, Franc¸ois C,
Schalij I, Dorfmu
¨ller P, Simonneau G, Fadel E, Perros F,
Boonstra A, Postmus PE, van der Velden J, Vonk-Noordegraaf A,
Humbert M, Eddahibi S, Guignabert C (2012) Dysregulated
renin–angiotensin–aldosterone system contributes to pulmonary
arterial hypertension. Am J Respir Crit Care Med 186:780–789.
doi:10.1164/rccm.201203-0411OC
15. Dumitrascu R, Weissmann N, Ghofrani HA, Dony E, Beuerlein
K, Schmidt H, Stasch JP, Gnoth MJ, Seeger W, Grimminger F,
Schermuly RT (2006) Activation of soluble guanylate cyclase
reverses experimental pulmonary hypertension and vascular
remodeling. Circulation 113:286–295. doi:10.1161/CIRCULA
TIONAHA.105.581405
16. Ferrario CM, Strawn WB (2006) Role of the renin–angiotensin–
aldosterone system and proinflammatory mediators in cardio-
vascular disease. Am J Cardiol 98:121–128. doi:10.1016/j.amj
card.2006.01.059
17. Fraccarollo D, Galuppo P, Motschenbacher S, Ruetten H, Scha
¨fer
A, Bauersachs J (2014) Soluble guanylyl cyclase activation
improves progressive cardiac remodeling and failure after
myocardial infarction. Cardioprotection over ACE inhibition.
Basic Res Cardiol 109:421. doi:10.1007/s00395-014-0421-1
18. Frantz S, Klaiber M, Baba HA,Oberwinkler H, Vo
¨lker K, Gaßner B,
Bayer B, Abeßer M, Schuh K, Feil R, Hofmann F, Kuhn M (2013)
Stress-dependent dilated cardiomyopathy in mice with cardiomy-
ocyte-restricted inactivation of cyclic GMP-dependent protein
kinase I. Eur Heart J 34:1233–1244. doi:10.1093/eurheartj/ehr445
19. Fujita S, Shimojo N, Terasaki F, Otsuka K, Hosotani N, Kohda Y,
Tanaka T, Nishioka T, Yoshida T, Hiroe M, Kitaura Y, Ishizaka
N, Imanaka-Yoshida K (2013) Atrial natriuretic peptide exerts
protective action against angiotensin II-induced cardiac remod-
eling by attenuating inflammation via endothelin-1/endothelin
receptor A cascade. Heart Vessels 28:646–657. doi:10.1007/
s00380-012-0311-0
20. Guignabert C, Alvira CM, Alastalo TP, Sawada H, Hansmann G,
Zhao M, Wang L, El-Bizri N, Rabinovitch M (2009) Tie2-me-
diated loss of peroxisome proliferator-activated receptor-gamma
in mice causes PDGF receptor-beta-dependent pulmonary arterial
muscularization. Am J Physiol 297:L1082–L1090. doi:10.1152/
ajplung.00199.2009
21. Guignabert C, Tu L, Girerd B, Ricard N, Huertas A, Montani D,
Humbert M (2015) New molecular targets of pulmonary vascular
remodeling in pulmonary arterial hypertension: importance of
endothelial communication. Chest 147:529–537. doi:10.1378/
chest.14-0862
22. Ho
¨hne C, Drzimalla M, Krebs MO, Boemke W, Kaczmarczyk G
(2003) Atrial natriuretic peptide ameliorates hypoxic pulmonary
vasoconstriction without influencing systemic circulation.
J Physiol Pharmacol 54:497–510
23. Hong KH, Lee YJ, Lee E, Park SO, Han C, Beppu H, Li E,
Raizada MK, Bloch KD, Oh SP (2008) Genetic ablation of the
BMPR2 gene in pulmonary endothelium is sufficient to predis-
pose to pulmonary arterial hypertension. Circulation
118:722–730. doi:10.1161/CIRCULATIONAHA.107.736801
24. Hutchinson HG, Trindade PT, Cunanan DB, Wu CF, Pratt RE
(1997) Mechanisms of natriuretic peptide-induced growth
22 Page 14 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
inhibition of vascular smooth muscle cells. Cardiovasc Res
35:158–167. doi:10.1016/S0008-6363(97)00086-2
25. Irwin DC, Tissot van Patot MC, Tucker A, Bowen R (2005)
Direct ANP inhibition of hypoxia-induced inflammatory path-
ways in pulmonary microvascular and macrovascular endothelial
monolayers. Am J Physiol 288:L849–L859. doi:10.1152/ajplung.
00294.2004
26. Jin H, Yang RH, Chen YF, Jackson RM (1991) Atrial natriuretic
peptide in acute hypoxia-induced pulmonary hypertension in rats.
J Appl Physiol 71:807–814
27. Kilic
´A, Bubikat A, Gassner B, Baba HA, Kuhn M (2007) Local
actions of atrial natriuretic peptide counteract angiotensin II
stimulated cardiac remodeling. Endocrinology 148:4162–4169.
doi:10.1210/en.2007-0182
28. Klinger JR, Warburton RR, Pietras LA, Smithies O, Swift R, Hill
NS (1999) Genetic disruption of atrial natriuretic peptide causes
pulmonary hypertension in normoxic and hypoxic mice. Am J
Physiol 276:L868–L874
29. Klinger JR, Warburton RR, Pietras L, Oliver P, Fox J, Smithies
O, Hill NS (2002) Targeted disruption of the gene for natriuretic
peptide receptor-A worsens hypoxia-induced cardiac hypertro-
phy. Am J Physiol 282:H58–H65
30. Kuhn M (2012) Endothelial actions of atrial and B-type natri-
uretic peptides. Br J Pharmacol 166:522–531. doi:10.1111/j.
1476-5381.2012.01827.x
31. Kuhn M, Vo
¨lker K, Schwarz K, Carbajo-Lozoya J, Flo
¨gel U,
Jacoby C, Stypmann J, van Eickels M, Gambaryan S, Hartmann
M, Werner M, Wieland T, Schrader J, Baba HA (2009) The
natriuretic peptide/guanylyl cyclase-A system functions as a
stress-responsive regulator of angiogenesis in mice. J Clin Invest
119:2019–2030. doi:10.1172/JCI37430
32. Liu LS, Cheng HY, Chin WJ, Jin HK, Oparil S (1989) Atrial
natriuretic peptide lowers pulmonary arterial pressure in patients
with high altitude disease. Am J Med Sci 298:397–401
33. Lopez MJ, Wong SK, Kishimoto I, Dubois S, Mach V, Friesen J,
Garbers DL, Beuve A (1995) Salt-resistant hypertension in mice
lacking the guanylyl cyclase-A receptor for atrial natriuretic
peptide. Nature 378:65–68. doi:10.1038/378065a0
34. Madjdpour C, Jewell UR, Kneller S, Ziegler U, Schwendener R,
Booy C, Kla
¨usli L, Pasch T, Schimmer RC, Beck-Schimmer B
(2003) Decreased alveolar oxygen induces lung inflammation.
Am J Physiol Lung Cell Mol Physiol 284:L360–L367. doi:10.
1152/ajplung.00158.2002
35. Maron BA, Leopold JA (2014) The role of the renin–angiotensin–
aldosterone system in the pathobiology of pulmonary arterial
hypertension (2013 Grover Conference series). Pulm Circ
4:200–210. doi:10.1086/675984
36. Minamino T, Christou H, Hsieh CM, Liu Y, Dhawan V, Abraham
NG, Perrella MA, Mitsialis SA, Kourembanas S (2001) Targeted
expression of heme oxygenase-1 prevents the pulmonary
inflammatory and vascular responses to hypoxia. Proc Natl Acad
Sci USA 98:8798–8803. doi:10.1073/pnas.161272598
37. Morrell NW, Atochina EN, Morris KG, Danilov SM, Stenmark
KR (1995) Angiotensin converting enzyme expression is
increased in small pulmonary arteries of rats with hypoxia-in-
duced pulmonary hypertension. J Clin Invest 96:1823–1833.
doi:10.1172/JCI118228
38. Morrell NW, Morris KG, Stenmark KR (1995) Role of angio-
tensin-converting enzyme and angiotensin II in development of
hypoxic pulmonary hypertension. Am J Physiol 269:H1186–
H1194
39. el Mtairag M, Houard X, Rais S, Pasquier C, Oudghiri M, Jacob
MP, Meilhac O, Michel JB (2002) Pharmacological potentiation
of natriuretic peptide limits polymorphonuclear neutrophil–vas-
cular cell interactions. Arterioscler Thromb Vasc Biol
22:1824–1831. doi:10.1161/01.ATV.0000037102.31086.F4
40. Nakamoto T, Harasawa H, Akimoto K, Hirata H, Kaneko H,
Kaneko N, Sorimachi K (2005) Effects of olmesartan medoxomil
as an angiotensin II-receptor blocker in chronic hypoxic rats. Eur
J Pharmacol 528:43–51. doi:10.1016/j.ejphar.2005.10.063
41. National Institutes of Health. Guide for the care and use of lab-
oratory animals. NIH publication, 85-231996
42. Newton-Cheh C, Larson MG, Vasan RS, Levy D, Bloch KD,
Surti A, Guiducci C, Kathiresan S, Benjamin EJ, Struck J,
Morgenthaler NG, Bergmann A, Blankenberg S, Kee F, Nilsson
P, Yin X, Peltonen L, Vartiainen E, Salomaa V, Hirschhorn JN,
Melander O, Wang TJ (2009) Association of common variants in
NPPA and NPPB with circulating natriuretic peptides and blood
pressure. Nat Genet 41:348–353. doi:10.1038/ng.328
43. Orte C, Polak JM, Haworth SG, Yacoub MH, Morrell NW (2000)
Expression of pulmonary vascular angiotensin-converting
enzyme in primary and secondary plexiform pulmonary hyper-
tension. J Pathol 192:379–384. doi:10.1002/1096-9896
44. Pfeifer M, Wolf K, Blumberg FC, Elsner D, Muders F, Holmer
SR, Riegger GA, Kurtz A (1997) ANP gene expression in rat
hearts during hypoxia. Pflugers Arch 434:63–69
45. Rabinovitch M (2008) Molecular pathogenesis of pulmonary
arterial hypertension. J Clin Invest 118:2372–2379. doi:10.1172/
JCI33452
46. Sabrane K, Kruse MN, Fabritz L, Zetsche B, Mitko D, Skryabin
BV, Zwiener M, Baba HA, Yanagisawa M, Kuhn M (2005)
Vascular endothelium is critically involved in the hypotensive
and hypovolemic actions of atrial natriuretic peptide. J Clin
Invest 115:1666–1674. doi:10.1172/JCI23360
47. Savai R, Pullamsetti SS, Kolbe J, Bieniek E, Voswinckel R, Fink
L, Scheed A, Ritter C, Dahal BK, Vater A, Klussmann S, Gho-
frani HA, Weissmann N, Klepetko W, Banat GA, Seeger W,
Grimminger F, Schermuly RT (2012) Immune and inflammatory
cell involvement in the pathology of idiopathic pulmonary arte-
rial hypertension. Am J Respir Crit Care Med 186:897–908.
doi:10.1164/rccm.201202-0335OC
48. Schro
¨ter J, Zahedi RP, Hartmann M, Gassner B, Gazinski A,
Waschke J, Sickmann A, Kuhn M (2010) Homologous desensi-
tization of guanylyl cyclase A, the receptor for atrial natriuretic
peptide, is associated with a complex phosphorylation pattern.
FEBS J 277:2440–2453. doi:10.1111/j.1742-4658.2010.07658.x
49. Tanabe Y, Morikawa Y, Kato T, Kanai S, Watakabe T, Nishijima
A, Iwata H, Isobe K, Ishizaki M, Nakayama K (2006) Effects of
olmesartan, an AT
1
receptor antagonist, on hypoxia-induced
activation of ERK1/2 and pro-inflammatory signals in the mouse
lung. Naunyn Schmiedebergs Arch Pharmacol 374:235–248.
doi:10.1007/s00210-006-0110-1
50. Taraseviciene-Stewart L, Scerbavicius R, Stewart JM, Gera L,
Demura Y, Cool C, Kasper M, Voelkel NF (2005) Treatment of
severe pulmonary hypertension: a bradykinin receptor 2 agonist
B9972 causes reduction of pulmonary artery pressure and right
ventricular hypertrophy. Peptides 26:1292–1300. doi:10.1016/j.
peptides.2005.03.050
51. Tian Y, Mambetsariev I, Sarich N, Meng F, Birukova AA (2015)
Role of microtubules in attenuation of PepG-induced vascular
endothelial dysfunction by atrial natriuretic peptide. Biochim
Biophys Acta 1852:104–119. doi:10.1016/j.bbadis.2014.10.012
52. Tsuneyoshi H, Nishina T, Nomoto T, Kanemitsu H, Kawakami R,
Unimonh O, Nishimura K, Komeda M (2004) Atrial natriuretic
peptide helps prevent late remodeling after left ventricular
aneurysm repair. Circulation 110:174–179. doi:10.1161/01.CIR.
0000138348.77856.ef
53. van Suylen RJ, Aartsen WM, Smits JF, Daemen MJ (2013)
Dissociation of pulmonary vascular remodeling and right ven-
tricular pressure in tissue angiotensin-converting enzyme-defi-
cient mice under conditions of chronic alveolar hypoxia. Am J
Basic Res Cardiol (2016) 111:22 Page 15 of 16 22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Respir Crit Care Med 163:1241–1245. doi:10.1164/ajrccm.163.5.
2003144
54. Wong WT, Tian XY, Xu A, Ng CF, Lee HK, Chen ZY, Au CL,
Yao X, Huang Y (2010) Angiotensin II type 1 receptor-dependent
oxidative stress mediates endothelial dysfunction in type 2 dia-
betic mice. Antioxid Redox Signal 13:757–768. doi:10.1089/ars.
2009.2831
55. Yoshida H, Nakamura M, Makita S, Hiramori K (1994) Inhibi-
tory effect of atrial natriuretic peptide on accelerated endothelin
secretion from cultured human endothelial cells. J Atheroscler
Thromb 1:76–79
56. Zhao L, Winter RJ, Krausz T, Hughes JM (1991) Effects of
continuous infusion of atrial natriuretic peptide on the pulmonary
hypertension induced by chronic hypoxia in rats. Clin Sci (Lond)
81:379–385. doi:10.1042/cs0810379
57. Zhao L, al-Tubuly R, Sebkhi A, Owji AA, Nunez DJ, Wilkins
MR (1996) Angiotensin II receptor expression and inhibition in
the chronically hypoxic rat lung. Br J Pharmacol 119:1217–1222.
doi:10.1111/j.1476-5381.1996.tb16025.x
58. Zhao L, Long L, Morrell NW, Wilkins MR (1999) NPR-A-de-
ficient mice show increased susceptibility to hypoxia-induced
pulmonary hypertension. Circulation 99:605–607. doi:10.1161/
01.CIR.99.5.605
22 Page 16 of 16 Basic Res Cardiol (2016) 111:22
123
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... This raises the question of whether the endogenous endothelial hormone exerts local protective effects in pulmonary microcirculation and whether such effects are preserved in PAH. To approach this question, firstly, we studied lung CNP and GC-B expression levels in two experimental models of PH: Monocrotaline (MCT)-induced PH in rats 27 and milder, chronic hypoxia (HOX)-induced PH in mice 28 . Quantitative real-time RT-PCR (qRT PCR) revealed that the CNP expression levels were significantly reduced in lung samples from rats and mice with PH in comparison to respective controls ( Fig. 8a and b, left panels). ...
... Hence, CNP signaling prevails in pericytes and ANP signaling in ECs, which emphasizes the distinct and complementary functions of these hormones. In fact, our previous studies had already shown endothelial-dependent protective effects of ANP in experimental PH 28 . ...
... CNP, via GC-B/cGMP signaling, activates cGKI. It has been shown that cGKI elicits inactivating phosphorylations of RhoA at Ser 188 and of GSK3b at Ser 9 , thereby preventing their inhibitory phosphorylations of PTEN 27,28 . Activated PTEN dephosphorylates PIP3 and prevents AKT activation, resulting in an increase of nuclear FoxO3 and a concomitant reduction in pericyte proliferation. ...
Article
Full-text available
Pericyte dysfunction, with excessive migration, hyperproliferation, and differentiation into smooth muscle-like cells contributes to vascular remodeling in Pulmonary Arterial Hypertension (PAH). Augmented expression and action of growth factors trigger these pathological changes. Endogenous factors opposing such alterations are barely known. Here, we examine whether and how the endothelial hormone C-type natriuretic peptide (CNP), signaling through the cyclic guanosine monophosphate (cGMP) -producing guanylyl cyclase B (GC-B) receptor, attenuates the pericyte dysfunction observed in PAH. The results demonstrate that CNP/GC-B/cGMP signaling is preserved in lung pericytes from patients with PAH and prevents their growth factor-induced proliferation, migration, and transdifferentiation. The anti-proliferative effect of CNP is mediated by cGMP-dependent protein kinase I and inhibition of the Phosphoinositide 3-kinase (PI3K)/AKT pathway, ultimately leading to the nuclear stabilization and activation of the Forkhead Box O 3 (FoxO3) transcription factor. Augmentation of the CNP/GC-B/cGMP/FoxO3 signaling pathway might be a target for novel therapeutics in the field of PAH.
... The Npr3 receptor is highly expressed in lung endothelial cells and lacks an intracellular guanylate cyclase domain; binding of NPs to Npr3 leads to cellular internalization and lysosomal degradation of ANP (Kolb et al., 2015). Previous studies have demonstrated that increased circulating ANP in the chronic hypoxia model is driven largely by selective downregulation of lung Npr3 expression (Francis et al., 2011;Sun et al., 2000;Werner et al., 2016). ...
... The relationship between NP-dependent signaling and CH-PH is complex. Multiple previous studies have used knockout and transgenic mice to demonstrate a protective role for NPs in the CH-PH model (Chen et al., 2006;Klinger et al., 1999;Sun et al., 2000;Werner et al., 2016). ANP-knockout and ANP receptor (guanylyl cyclase-A; GC-A)-knockout mice demonstrated baseline (normoxic) increases in myocardial mass (RV and LV) and pulmonary vessel muscularization and more severe PH after exposure to 3-6 weeks of chronic hypoxia (Chen et al., 2006;Klinger et al., 1999;Sun et al., 2000;Werner et al., 2016). ...
... Multiple previous studies have used knockout and transgenic mice to demonstrate a protective role for NPs in the CH-PH model (Chen et al., 2006;Klinger et al., 1999;Sun et al., 2000;Werner et al., 2016). ANP-knockout and ANP receptor (guanylyl cyclase-A; GC-A)-knockout mice demonstrated baseline (normoxic) increases in myocardial mass (RV and LV) and pulmonary vessel muscularization and more severe PH after exposure to 3-6 weeks of chronic hypoxia (Chen et al., 2006;Klinger et al., 1999;Sun et al., 2000;Werner et al., 2016). Conversely, transgenic mice overexpressing ANP were protected from pulmonary vascular muscularization, RVH, and RVSP in the CH-PH model (Klinger et al., 1993). ...
Article
Full-text available
Inhibition of cyclic guanosine monophosphate (cGMP)‐specific phosphodiesterases (PDEs) is a cornerstone of pulmonary arterial hypertension (PAH)‐specific therapy. PDE9A, expressed in the heart and lung tissue, has the highest affinity for cGMP of all known PDEs. PDE9A deficiency protects mice against chronic left ventricular (LV) pressure overload via increased natriuretic peptide (NP)‐dependent cGMP signaling. Chronic‐hypoxic pulmonary hypertension (CH‐PH) is a model of chronic right ventricular (RV) pressure overload, and previous studies have demonstrated a protective role for NPs in the murine model. Therefore, we hypothesized that PDE9A deficiency would promote NP‐dependent cGMP signaling and prevent RV remodeling in the CH‐PH model, analogous to findings in the LV. We exposed wild‐type and PDE9A‐deficient (Pde9a−/−) C57BL/6 mice to CH‐PH for 3 weeks. We measured RV pressure, hypertrophy, and levels of lung and RV cGMP, PDE9A, PDE5A, and phosphorylation of the protein kinase G substrate VASP (vasodilatory‐stimulated phosphoprotein) after CH‐PH. In wild‐type mice, CH‐PH was associated with increased circulating ANP and lung PDE5A, but no increase in cGMP, PDE9A, or VASP phosphorylation. Downstream effectors of cGMP were not increased in Pde9a−/− mice exposed to CH‐PH compared with Pde9a+/+ littermates, and CH‐PH induced increases in RV pressure and hypertrophy were not attenuated in knockout mice. Taken together, these findings argue against a prominent role for PDE9A in the murine CH‐PH model. Phosphodiesterase (PDE) 9A deficiency protects mice from chronic left ventricular pressure overload by augmenting NP‐dependent signaling. Chronic‐hypoxic PH is a model of chronic right ventricular pressure overload, and NP overexpression has previously been shown to be protective. PDE 9A deficient mice did not have increased NP‐dependent signaling and were not protected from chronic‐hypoxic PH.
... MLEC cultures were described before. 6,23 Cortices were incubated in 30 U/mL papain and 40 μg/mL DNase I (70 minutes at 37°C). The tissue digest was pelleted, resuspended in endothelial growth medium (CC-3202; Lonza Walkersville Inc), and plated on collagen-coated wells. ...
... Intracellular cGMP was extracted with ice-cold 70% ethanol and determined by radioimmunoassay. 6,23 For live imaging of pericyte cAMP responses to CNP, cultured pericytes or whole retinas from tamoxifen-induced R26-STOP-Epac1-camps; PDGFRβ-Cre-ERT2 mice were subjected to FRET recordings as described. 17 ...
... Terminal simultaneous recordings of aortic (AoP) and central (jugular) venous pressures (CVP), heart rate, cardiac output (CO), and ejection fraction were performed under isoflurane (2%) anesthesia with Millar SPR-671 Mikro-Tip pressure and SPR-839 Mikro-Tip pressure-volume catheters. 23,27 Total peripheral resistance (TPR, mm Hg/mL×min) was derived from the equation ΔBP (AoP -CVP) = CO × TPR. 28 ...
Article
Background -Peripheral vascular resistance has a major impact on arterial blood pressure levels. Endothelial C-type natriuretic peptide (CNP) participates in the local regulation of vascular tone but the target cells remain controversial. The cGMP-producing guanylyl cyclase-B (GC-B) receptor for CNP is expressed in vascular smooth muscle cells (VSMC). However, whereas endothelial cell-specific CNP knockout mice are hypertensive, mice with deletion of GC-B in VSMC have unaltered blood pressure. Methods -We analyzed whether the vasodilating response to CNP changes along the vascular tree, i.e. whether the GC-B receptor is expressed in microvascular types of cells. Mice with a floxed GC-B (Npr2) gene were interbred withTie2-CreorPDGF-Rβ-CreERT2lines to develop mice lacking GC-B in endothelial cells or in precapillary arteriolar SMC and capillary pericytes. Intravital microscopy, (non)invasive hemodynamics, fluorescence energy transfer studies of pericyte's cAMP levelsin situand renal physiology were combined to dissect whether and how CNP/GC-B/cGMP signaling modulates microcirculatory tone and blood pressure. Results -Intravital microscopy studies revealed that the vasodilatatory effect of CNP increases towards small-diameter arterioles and capillaries. Consistently, CNP did not prevent endothelin-1-induced acute constrictions of proximal arterioles but fully reversed endothelin effects in precapillary arterioles and capillaries. Here, the GC-B receptor is expressed both in endothelial and mural cells, i.e. in pericytes. Notably, the vasodilatatory effects of CNP were preserved in mice with endothelial GC-B deletion but abolished in mice lacking GC-B in microcirculatory SMC and pericytes. CNP, via GC-B/cGMP signaling modulates two signaling cascades in pericytes: it activates cGMP-dependent protein kinase I to phosphorylate downstream targets such as the cytoskeleton-associated vasodilator activated phosphoprotein; and it inhibits phosphodiesterase 3A, thereby enhancing pericyte's cAMP levels. Ultimately these pathways prevent endothelin-induced increases of pericyte calcium levels and pericyte contraction. Mice with deletion of GC-B in microcirculatory SMC and pericytes have elevated peripheral resistance and chronic arterial hypertension without a change in renal function. Conclusions -Our studies indicate that endothelial CNP regulates distal arteriolar and capillary blood flow. CNP-induced GC-B/cGMP signaling in microvascular SMC and pericytes is essential for the maintenance of normal microvascular resistance and blood pressure.
... Our results showed that the ANP level was increased during hypoxia and this is due to partly to hypoxia effect per se and partly by increased load on right ventricle. It seems that its effect was blunted by a concomitant decreased in NO production; the action of ANP on vascular smooth muscle is through particulate guanylyl cyclase (GC-A) which is found at a high level in the mouse lung [27]. Hypoxic conditions markedly decrease the expression and activity of GC-A in pulmonary endothelial cells [27]. ...
... It seems that its effect was blunted by a concomitant decreased in NO production; the action of ANP on vascular smooth muscle is through particulate guanylyl cyclase (GC-A) which is found at a high level in the mouse lung [27]. Hypoxic conditions markedly decrease the expression and activity of GC-A in pulmonary endothelial cells [27]. The blunted effect of decreased NO production, and decreased activity and expression of GC-A, explains the overall hypoxic effect of an increase in pulmonary vascular resistance and thus more load on the right ventricle, leading to its hypertrophy. ...
Article
Full-text available
In this study, we investigated the effect of hypoxia and concomitant sildenafil treatment on MHC isoforms in hypoxia-induced hypertrophied right ventricles. Right ventricular hypertrophy was induced in mice by exposing them to hypoxic stimulus (11% ambient oxygen) in a normobaric chamber for 20 days. 45 mice were used in this study, distributed randomly into three groups: the first group served as a control (CO), the second group was exposed to hypoxia for 20 days without sildenafil treatment (HY), and the third group was given sildenafil orally at a dose of 30 mg.kg-1.day-1 plus exposure to hypoxia for 20 days (HS). Relative amounts of MHC isoforms were calculated using two ELISA kits containing antibodies against α and β MHC, and by SDS-PAGE. Compared with the CO group, the HY group showed a significant increase in right ventricle weight/left ventricle plus septum ratio (Fulton's ratio). The HS group showed a significant decrease in Fulton's ratio compared with the HY group, but not with the CO group. Expression of the MHC-β isoform was significantly increased in the HY group compared with the CO group. There was no significant difference in MHC-β between the HY group and the HS group. Plasma atrial natriuretic peptide level was significantly higher in HY group than HS group and did not return to normal after sildenafil treatment. Conclusion: sildenafil reversed the right ventricular hypertrophy induced by hypoxia but did not decrease the expression of MHC-β to normal levels.
... AMPE damages the vascular endothelium, induces platelet activation and increases the plasma concentration of thromboxane A2 (TXA2) 10 , which leads to prostaglandin I2 (PGI2) metabolic disorders and elevated plasma concentrations of endothelin-1 (ET-1) 11,12 . Elevated plasma TXA2, PGI2 and ET-1 concentrations may cause pulmonary thrombosis and hypertension, which may further aggravate pulmonary embolism and exacerbate the myocardial damage associated with pulmonary embolism [10][11][12] . ...
... AMPE damages the vascular endothelium, induces platelet activation and increases the plasma concentration of thromboxane A2 (TXA2) 10 , which leads to prostaglandin I2 (PGI2) metabolic disorders and elevated plasma concentrations of endothelin-1 (ET-1) 11,12 . Elevated plasma TXA2, PGI2 and ET-1 concentrations may cause pulmonary thrombosis and hypertension, which may further aggravate pulmonary embolism and exacerbate the myocardial damage associated with pulmonary embolism [10][11][12] . ...
Article
Full-text available
Purpose: To investigate changes in the plasma concentrations of cardiac troponin I (CTnI), thromboxane A2 (TXA2), prostaglandin I2 (PGI2) and endothelin-1 (ET-1) in rabbits with massive pulmonary embolism (AMPE) and the impact of nitric oxide inhalation (NOI) on these indices. Methods: A total of 30 Japanese rabbits were used to construct an MPE model and were divided into 3 groups equally (n=10), including an EXP group (undergoing modeling alone), an NOI group (receiving NOI 2 h post-modeling) and a CON group (receiving intravenous physiological saline). Results: In the model group, plasma concentration of CTnI peaked at 16 h following modeling (0.46±0.10 µg/ml) and significantly decreased following NOI. Plasma levels of TXB2, PGI2 and ET-1 peaked at 12, 16 and 8 h following modeling, respectively, and significantly decreased at different time points (0, 2, 4, 8, 12, 16, 20 and 24 h) following NOI. A significant correlation was observed between the peak plasma CTnI concentration and peak TXB2, 6-keto prostaglandin F1α and ET-1 concentrations in the model and NOI groups. Conclusion: Increases in plasma TXA2, PGI2 and ET-1 levels causes myocardial damage in a rabbit model of AMPE; however, NOI effectively down regulates the plasma concentration of these molecules to produce a myocardial-protective effect.
... Emerging evidence indicates that the cause of PE is associated with genetic factors since it was first expounded in the early 1960s [3,4]. Pregnant mice lacking catechol-O-methyltransferase, Corin, atrial natriuretic peptide, or Elabela exhibit a PE-like phenotype [5][6][7][8]. In humans, geneticists [9] suggested that a single gene may possibly be responsible for PE. ...
Article
Full-text available
Objective: Preeclampsia (PE) is a severe complication in pregnancy and a leading cause of maternal and infant mortality. However, the exact underlying etiology of PE remains unknown. Emerging evidence indicates that the cause of PE is associated with genetic factors. Therefore, the aim of this study is to identify susceptibility genes to PE. Materials and methods: Human Exome BeadChip assays were conducted using 370 cases and 482 controls and 21 loci were discovered. A further independent set of 958 cases and 1007 controls were recruited for genotyping to determine whether the genes of interest ROS1 and PTPRK are associated with PE. Immunohistochemistry was used for localization. Both qPCR and Western blotting were utilized to investigate the levels of PTPRK in placentas of 20 PE and 20 normal pregnancies. Results: The allele frequency of PTPRK rs3190930 differed significantly between PE and controls and was particularly significant in severe PE subgroup and early-onset PE subgroup. PTPRK is primarily localized in placental trophoblast cells. The mRNA and protein levels of PTPRK in PE were significantly higher than those in controls. Conclusion: These results suggest that PTPRK appears to be a previously unrecognized susceptibility gene for PE in Han Chinese women, and its expression is also associated with PE, while ROS1 rs9489124 has no apparent correlation with PE risk.
... The main function of ANP is to lower blood pressure by a number of actions on the kidney where it increases vascular permeability, induces vasorelaxation and causes natriuresis (27). Furthermore, ANP inhibits the production of aldosterone by actions on the adrenal glands and induces vasorelaxation of vascular smooth muscle cells in general (28)(29)(30)(31)(32)(33). ANP may also be secreted from cells in the lung (33,34) and some studies have shown a direct relaxant effect of ANP on isolated bronchi from guinea pigs and cows (35)(36)(37). ...
Article
Full-text available
Glucagon-like peptide-1 (GLP-1) is protective in lung disease models but the underlying mechanisms remain elusive. Since the hormone atrial natriuretic peptide (ANP) also has beneficial effects in lung disease, we hypothesized that GLP-1 effects may be mediated by ANP expression. To study this putative link, we used a mouse model of chronic obstructive pulmonary disease (COPD) and assessed lung function by unrestrained whole body plethysmography. In one study, we investigated the role of endogenous GLP-1 by genetic GLP-1R knockout (KO) and pharmaceutical blockade of the GLP-1R with the antagonist exendin 9-39 (EX-9). In another study the effects of exogenous GLP-1 were assessed. Lastly, we investigated bronchodilatory properties of ANP and a GLP-1R agonist on isolated bronchial sections from healthy and COPD mice. Lung function did not differ between mice receiving PBS and EX-9 or between GLP-1R KO mice and their WT littermates. The COPD mice receiving GLP-1R agonist improved pulmonary function (P<0.01) with less inflammation, but no less emphysema compared to PBS-treated mice. Compared with the PBS-treated mice, treatment with GLP-1 agonist increased ANP gene expression by 10-fold (P<0.01) and decreased endothelin-1 (ET-1) (P<0.01), a peptide associated with bronchoconstriction. ANP had moderate broncodilatory effects in isolated bronchial sections and GLP-1R agonist also showed bronchodilatory properties but less than ANP. Responses to both peptides were significantly increased in COPD mice (P<0.05, P<0.01). Taken together, our study suggests a link between GLP-1 and ANP in COPD.
... 1,2 Imbalance in endothelialderived vasoactive factors, excessive infiltration of proinflammatory cytokines along with dysregulated vascular cells proliferation drive the PAH pathobiology. [3][4][5] However, the precise molecular mechanisms underlying PAH is still poorly understood. 6,7 Recently, metabolic reprogramming had increasingly been reported in PAH, highlighting the role of metabolic pathways as an attractive concept for therapeutic intervention. ...
Article
Background: Pulmonary arterial hypertension (PAH) is a severe progressive disease with systemic metabolic dysregulation. Monocrotaline (MCT)-induced and hypoxia-induced pulmonary hypertension (PH) rodent models are the most widely used preclinical models, however, whether or not these preclinical models recapitulate metabolomic profiles of PAH patients remain unclear. Methods: In this study, a targeted metabolomics panel of 126 small molecule metabolites was conducted. We applied it to the plasma of the two preclinical rodent models of PH and 30 idiopathic pulmonary arterial hypertension (IPAH) patients as well as 30 healthy controls to comparatively assess the metabolomic profiles of PAH patients and rodent models. Results: Significantly different metabolomics profiling and pathways were shown among the two classical rodent models and IPAH patients. Pathway analysis demonstrated that methionine metabolism and urea cycle metabolism were the most significant pathway involved in the pathogenesis of hypoxia-induced PH model and MCT-induced model, respectively, and both of them were also observed in the dysregulated pathways in IPAH patients. Conclusions: These two models may develop PAH through different metabolomic pathways and each of the two classical PH model resembles IPAH patients in certain aspects.
... Both molecules have been reported to accelerate the development of hypoxia induce PAH [30,39]. The increase ANP lung and heart expression indicates an attempt of endogenous system to induce pulmonary vaso-relaxation and to modulate lung and heart vascular remodeling in response to hypoxia [56,57]. This was coordinated by hypoxia induced activation of acute phase related transcriptional factors: Nrf2 and NF-kB in both mouse strains. ...
Article
Pulmonary-artery-hypertension (PAH) is a life-threatening and highly invalidating chronic disorder. Chronic oxidation contributes to lung damage and disease progression. Peroxiredoxin-2 (Prx2) is a typical 2-cysteine (Cys) peroxiredoxin but its role on lung homestasis is yet to be fully defined. Here, we showed that Prx2-/- mice displayed chronic lung inflammatory disease associated with (i) abnormal pulmonary vascular dysfunction; and (ii) increased markers of extracellular-matrix remodeling. Hypoxia was used to induce PAH. We focused on the early phase PAH to dissect the role of Prx2 in generation of PAH. Hypoxic Prx2-/-mice showed (i) amplified inflammatory response combined with cytokine storm; (ii) vascular activation and dysfunction; (iii) increased PDGF-B lung levels, as marker of extracellular-matrix deposition and remodeling; and (iv) ER stress with activation of UPR system and autophagy. Rescue experiments with in vivo the administration of fused-recombinant-PEP-Prx2 show a reduction in pulmonary inflammatory vasculopathy and in ER stress with down-regulation of autophagy. Thus, we propose Prx2 plays a pivotal role in the early stage of PAH as multimodal cytoprotector, targeting oxidation, inflammatory vasculopathy and ER stress with inhibition of autophagy. Collectively, our data indicate that Prx2 is able to interrupt the hypoxia induced vicious cycle involving oxidation-inflammation-autophagy in the pathogenesis of PAH.
Article
Full-text available
Background Pulmonary hypertension (PH) is a common complication of end-stage renal disease which is associated with adverse outcomes including all-cause mortality and cardiovascular events. Recent studies have demonstrated that Sacubitril/Valsartan (Sac/Val) as an enkephalinase inhibitor and angiotensin II receptor blocker could reduce pulmonary artery systolic pressure (PASP) and improve the prognosis of patients with heart failure. However, whether Sac/Val is effective in hemodialysis (HD) patients with PH is essentially unknown. In this retrospective study, we aimed to evaluate the efficacy and safety of Sac/Val in the treatment of PH in HD patients. Methods A total of 122 HD patients with PH were divided into Sac/Val group ( n = 71) and ARBs group ( n = 51) based on the treatment regimen. The PASP, other cardiac parameters measured by echocardiography, and cardiac biomarkers including N-terminal fragment of BNP (NT-proBNP) and cardiac troponin I (cTnI) were observed at baseline and 3 months after treatment. Results There were no significant differences in the baseline characteristics between the two groups. PASP decreased significantly from 45(38, 54) to 28(21, 40) mmHg in Sac/Val group ( p < 0.001). PASP reduced from 41(37, 51) to 34(27, 44) mmHg in ARBs group ( p < 0.001), and the decrease was more pronounced in the Sac/Val group ( p < 0.001). In addition, improvements in the right atrial diameter (RAD), left ventricular diameter (LVD), left ventricular posterior wall thickness (LVPWT), left atrial diameter (LAD), pulmonary artery diameter (PAD), left ventricular end-diastolic volume (LVEDV), left ventricular end-systolic volume (LVESV), left ventricular ejection fraction (LVEF), and fractional shortening (FS) were found in Sac/Val group (p s < 0.05). After 3 months, LVD, LAD, LVEDV, LVESV, LVEF, SV, and PASP were significantly improved in Sac/Val group compared with ARBs group (p s <0.05). Significant reduction in NT-proBNP [35,000 (15,000, 70,000) pg/ml vs. 7,042 (3,126, 29,060) pg/ml, p < 0.001] and cTnI [0.056(0.031, 0.085) ng/ml vs. 0.036 (0.012, 0.056) ng/ml, p < 0.001) were observed in Sac/Val group. No significant differences were observed in adverse events between the two groups (p s > 0.05). Conclusion Sac/Val seems to be an efficacious regimen in PH with favorable safety and has huge prospects for treating PH in HD patients.
Article
Full-text available
Pulmonary arterial hypertension (PAH) is a disorder in which mechanical obstruction of the pulmonary vascular bed is largely responsible for the rise in mean pulmonary arterial pressure, resulting in a progressive functional decline despite current available therapeutic options. The fundamental pathogenetic mechanisms underlying this disorder include pulmonary vasoconstriction, in situ thrombosis, medial hypertrophy, and intimal proliferation, leading to occlusion of the small to mid-sized pulmonary arterioles and the formation of plexiform lesions. Several predisposing or promoting mechanisms that contribute to excessive pulmonary vascular remodeling in PAH have emerged, such as altered crosstalk between cells within the vascular wall, sustained inflammation and dysimmunity, inhibition of cell death, and excessive activation of signaling pathways, in addition to the impact of systemic hormones, local growth factors, cytokines, transcription factors, and germline mutations. Although the spectrum of therapeutic options for PAH has expanded in the last 20 years, available therapies remain essentially palliative. However, over the past decade, a better understanding of new key regulators of this irreversible pulmonary vascular remodeling has been obtained. This review examines the state-of-the-art potential new targets for innovative research in PAH, focusing on (1) the crosstalk between cells within the pulmonary vascular wall, with particular attention to the role played by dysfunctional endothelial cells; (2) aberrant inflammatory and immune responses; (3) the abnormal extracellular matrix function; and (4) altered BMPRII/KCNK3 signaling systems. A better understanding of novel pathways and therapeutic targets will help in the designing of new and more effective approaches for PAH treatment.
Article
Aims: Cardiac hypertrophy is a common and often lethal complication of arterial hypertension. Elevation of myocyte cyclic GMP levels by local actions of endogenous atrial natriuretic peptide (ANP) and C-type natriuretic peptide (CNP) or by pharmacological inhibition of phosphodiesterase-5 was shown to counter-regulate pathological hypertrophy. It was suggested that cGMP-dependent protein kinase I (cGKI) mediates this protective effect, although the role in vivo is under debate. Here, we investigated whether cGKI modulates myocyte growth and/or function in the intact organism. Methods and results: To circumvent the systemic phenotype associated with germline ablation of cGKI, we inactivated the murine cGKI gene selectively in cardiomyocytes by Cre/loxP-mediated recombination. Mice with cardiomyocyte-restricted cGKI deletion exhibited unaltered cardiac morphology and function under resting conditions. Also, cardiac hypertrophic and contractile responses to β-adrenoreceptor stimulation by isoprenaline (at 40 mg/kg/day during 1 week) were unaltered. However, angiotensin II (Ang II, at 1000 ng/kg/min for 2 weeks) or transverse aortic constriction (for 3 weeks) provoked dilated cardiomyopathy with marked deterioration of cardiac function. This was accompanied by diminished expression of the \([Ca^{2+}]_i\)-regulating proteins SERCA2a and phospholamban (PLB) and a reduction in PLB phosphorylation at Ser16, the specific target site for cGKI, resulting in altered myocyte \(Ca^{2+}_i\) homeostasis. In isolated adult myocytes, CNP, but not ANP, stimulated PLB phosphorylation, \(Ca^{2+}_i\)-handling, and contractility via cGKI. Conclusion: These results indicate that the loss of cGKI in cardiac myocytes compromises the hypertrophic program to pathological stimulation, rendering the heart more susceptible to dysfunction. In particular, cGKI mediates stimulatory effects of CNP on myocyte \(Ca^{2+}_i\) handling and contractility.
Article
Objective: While natriuretic peptides can inhibit growth of vascular smooth muscle cells (VSMC), controversy exists as to whether this effect is mediated via the guanylate cyclase-coupled receptors, NPR-A and NPR-B, or the clearance receptor, NPR-C. The original aim of this study was to examine the mechanism by which the NPR-C receptor regulates growth. Methods: Rat VSMC were characterized with regard to natriuretic peptide receptor expression by RT/PCR and radioligand binding studies. The effect on growth following addition of the peptides and the ligands for NPR-C was measured by [3H]thymidine incorporation. Cyclic guanosine monophosphate (cGMP) levels were determined by radioimmunoassay and mitogen activating protein kinase activity was based on the phosphorylation of myelin basic protein. Results: In rat VSMC, passages 4–12, both atrial natriuretic peptide (ANP) and C-type natriuretic peptide (CNP) dose-dependently inhibited serum and PDGF-induced VSMC growth. In contrast, NPR-C specific ligands alone had no effect on cell growth but enhanced growth inhibition when co-administered with ANP and CNP. ANP and CNP also decreased PDGF-BB-stimulated MAP kinase activity. Once again, NPR-C specific ligands alone had no effect but enhanced the effects of ANP. Furthermore, a cGMP specific phosphodiesterase inhibitor dose-dependently inhibited VSMC growth and markedly enhanced natriuretic-peptide-induced inhibition at low peptide concentrations. To examine a potential mechanism for the controversy concerning the NPR-C, we investigated the autocrine expression of ANP and CNP by VSMC and found that mRNA encoding both peptides could be detected by RT/PCR. Conclusion: Our findings indicate that the guanylyl-cyclase-linked receptors mediate the antiproliferative actions of the natriuretic peptides on vascular smooth muscle cell growth. Moreover, we hypothesize that the apparent inhibition of growth by NPR-C specific ligands reported by others may be due to stabilization of natriuretic peptides produced by the cultured VSMC and subsequent action of these peptides at guanylyl-cyclase-linked receptors.
Article
Objective— Activated polymorphonuclear neutrophils (PMNs) are the main source of circulating neutral endopeptidase (NEP). We tested the hypothesis that NEP inhibition could potentiate the effect of atrial natriuretic peptide (ANP) on PMN-vascular cell interactions in vitro. Methods and Results— ANP alone and its potentiation by retrothiorphan, the NEP inhibitor, significantly inhibited superoxide, lysozyme, and matrix metalloproteinase (MMP)-9 release by N -formyl-Met-Leu-Phe-stimulated PMNs. Activated PMNs degraded exogenous ANP, which was prevented by NEP inhibition. Hypoxia significantly increased the adhesion of PMNs to endothelial cells and their subsequent MMP-9 release by 60% and 150%, respectively ( P <0.01). ANP and its potentiation by retrothiorphan limited PMN adhesion to hypoxic endothelial cells and thus decreased their MMP-9 release ( P <0.01). Smooth muscle cells (SMCs) incubated with conditioned medium of N -formyl-Met-Leu-Phe-stimulated PMNs exhibited morphological and biochemical changes characteristic of apoptosis (terminal deoxynucleotidyl transferase-mediated dUTP nick end-labeling positivity, nuclear condensation/fragmentation, poly ADP-ribose polymerase cleavage, and DNA laddering). SMC detachment and subsequent apoptosis could be related to leukocyte elastase-induced pericellular proteolysis, inasmuch as both events are inhibited by elastase inhibitors. ANP and its potentiation by retrothiorphan were able to limit elastase release, fibronectin degradation, and SMC apoptosis. Conclusions— ANP potentiation by NEP inhibition could limit PMN activation and its consequences on vascular cells.
Article
Apart from control of circulating fluid, atrial natriuretic peptide (ANP) exhibits anti-inflammatory effects in the lung. However, molecular mechanisms of ANP anti-inflammatory effects are not well-understood. Peripheral microtubule (MT) dynamics is essential for agonist-induced regulation of vascular endothelial permeability. Here we studied the role of MT-dependent signaling in ANP protective effects against endothelial cell (EC) barrier dysfunction and acute lung injury induced by Staphylococcus aureus-derived peptidoglican-G (PepG). PepG-induced vascular endothelial dysfunction was accompanied by MT destabilization and disruption of MT network. ANP attenuated PepG-induced MT disassembly, NFkB signaling and activity of MT-associated Rho activator GEF-H1 leading to attenuation of EC inflammatory activation reflected by expression of adhesion molecules ICAM1 and VCAM1. ANP-induced EC barrier preservation and MT stabilization were linked to phosphorylation and inactivation of MT-depolymerizing protein stathmin. Expression of stathmin phosphorylation-deficient mutant abolished ANP protective effects against PepG-induced inflammation and EC permeability. In contrast, siRNA-mediated stathmin knockdown prevented PepG-induced peripheral MT disassembly and endothelial barrier dysfunction. ANP protective effects in a murine model of PepG-induced lung injury were associated with increased phosphorylation of stathmin, while exacerbated lung injury in the ANP knockout mice was accompanied by decreased pool of stable MT. Stathmin knockdown in vivo reversed exacerbation of lung injury in the ANP knockout mice. These results show a novel MT-mediated mechanism of endothelial barrier protection by ANP in pulmonary EC and animal model of PepG-induced lung injury via stathmin-dependent control of MT assembly. Copyright © 2014. Published by Elsevier B.V.
Article
Pulmonary arterial hypertension (PAH) is associated with aberrant pulmonary vascular remodeling that leads to increased pulmonary artery pressure, pulmonary vascular resistance, and right ventricular dysfunction. There is now accumulating evidence that the renin-angiotensin-aldosterone system is activated and contributes to cardiopulmonary remodeling that occurs in PAH. Increased plasma and lung tissue levels of angiotensin and aldosterone have been detected in experimental models of PAH and shown to correlate with cardiopulmonary hemodynamics and pulmonary vascular remodeling. These processes are abrogated by treatment with angiotensin receptor or mineralocorticoid receptor antagonists. At a cellular level, angiotensin and aldosterone activate oxidant stress signaling pathways that decrease levels of bioavailable nitric oxide, increase inflammation, and promote cell proliferation, migration, extracellular matrix remodeling, and fibrosis. Clinically, enhanced renin-angiotensin activity and elevated levels of aldosterone have been detected in patients with PAH, which suggests a role for angiotensin and mineralocorticoid receptor antagonists in the treatment of PAH. This review will examine the current evidence linking renin-angiotensin-aldosterone system activation to PAH with an emphasis on the cellular and molecular mechanisms that are modulated by aldosterone and may be of importance for the pathobiology of PAH.
Article
Impaired nitric oxide (NO)-soluble guanylate cyclase (sGC)-cGMP signaling is involved in the pathogenesis of ischemic heart diseases, yet the impact of long-term sGC activation on progressive cardiac remodeling and heart failure after myocardial infarction (MI) has not been explored. Moreover, it is unknown whether stimulating the NO/heme-independent sGC provides additional benefits to ACE inhibition in chronic ischemic heart failure. Starting 10 days after MI, rats were treated with placebo, the sGC activator ataciguat (10 mg/kg/twice daily), ramipril (1 mg/kg/day), or a combination of both for 9 weeks. Long-term ataciguat therapy reduced left ventricular (LV) diastolic filling pressure and pulmonary edema, improved the rightward shift of the pressure-volume curve, LV contractile function and diastolic stiffness, without lowering blood pressure. NO/heme-independent sGC activation provided protection over ACE inhibition against mitochondrial superoxide production and progressive fibrotic remodeling, ultimately leading to a further improvement of cardiac performance, hypertrophic growth and heart failure. We found that ataciguat stimulating sGC activity was potentiated in (myo)fibroblasts during hypoxia-induced oxidative stress and that NO/heme-independent sGC activation modulated fibroblast-cardiomyocyte crosstalk in the context of heart failure and hypoxia. In addition, ataciguat inhibited human cardiac fibroblast differentiation and extracellular matrix protein production in response to TGF-β1. Overall, long-term sGC activation targeting extracellular matrix homeostasis conferred cardioprotection against progressive cardiac dysfunction, pathological remodeling and heart failure after myocardial infarction. NO/heme-independent sGC activation may prove to be a useful therapeutic target in patients with chronic heart failure and ongoing fibrotic remodeling.
Article
Background: Pulmonary hypertension (PH) is a life-threatening disorder characterized by increased pulmonary artery pressure, remodeling of the pulmonary vasculature, and right ventricular failure. Loss of endothelium-derived nitric oxide (NO) and prostacyclin contributes to PH pathogenesis, and current therapies are targeted to restore these pathways. Phosphodiesterases (PDEs) are a family of enzymes that break down cGMP and cAMP, which underpin the bioactivity of NO and prostacyclin. PDE5 inhibitors (eg, sildenafil) are licensed for PH, but a role for PDE2 in lung physiology and disease has yet to be established. Herein, we investigated whether PDE2 inhibition modulates pulmonary cyclic nucleotide signaling and ameliorates experimental PH. Methods and results: The selective PDE2 inhibitor BAY 60-7550 augmented atrial natriuretic peptide- and treprostinil-evoked pulmonary vascular relaxation in isolated arteries from chronically hypoxic rats. BAY 60-7550 prevented the onset of both hypoxia- and bleomycin-induced PH and produced a significantly greater reduction in disease severity when given in combination with a neutral endopeptidase inhibitor (enhances endogenous natriuretic peptides), trepostinil, inorganic nitrate (NO donor), or a PDE5 inhibitor. Proliferation of pulmonary artery smooth muscle cells from patients with pulmonary arterial hypertension was reduced by BAY 60-7550, an effect further enhanced in the presence of atrial natriuretic peptide, NO, and treprostinil. Conclusions: PDE2 inhibition elicits pulmonary dilation, prevents pulmonary vascular remodeling, and reduces the right ventricular hypertrophy characteristic of PH. This favorable pharmacodynamic profile is dependent on natriuretic peptide bioactivity and is additive with prostacyclin analogues, PDE5 inhibitor, and NO. PDE2 inhibition represents a viable, orally active therapy for PH.