ArticlePDF Available

Nonmyocyte ERK1/2 signaling contributes to load-induced cardiomyopathy in Marfan mice

Authors:

Abstract and Figures

Among children with the most severe presentation of Marfan syndrome (MFS), an inherited disorder of connective tissue caused by a deficiency of extracellular fibrillin-1, heart failure is the leading cause of death. Here, we show that, while MFS mice (Fbn1C1039G/+ mice) typically have normal cardiac function, pressure overload (PO) induces an acute and severe dilated cardiomyopathy in association with fibrosis and myocyte enlargement. Failing MFS hearts show high expression of TGF-β ligands, with increased TGF-β signaling in both nonmyocytes and myocytes; pathologic ERK activation is restricted to the nonmyocyte compartment. Informatively, TGF-β, angiotensin II type 1 receptor (AT1R), or ERK antagonism (with neutralizing antibody, losartan, or MEK inhibitor, respectively) prevents load-induced cardiac decompensation in MFS mice, despite persistent PO. In situ analyses revealed an unanticipated axis of activation in nonmyocytes, with AT1R-dependent ERK activation driving TGF-β ligand expression that culminates in both autocrine and paracrine overdrive of TGF-β signaling. The full compensation seen in wild-type mice exposed to mild PO correlates with enhanced deposition of extracellular fibrillin-1. Taken together, these data suggest that fibrillin-1 contributes to cardiac reserve in the face of hemodynamic stress, critically implicate nonmyocytes in disease pathogenesis, and validate ERK as a therapeutic target in MFS-related cardiac decompensation.
Content may be subject to copyright.
1insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
Authorship note: E.G. MacFarlane
and E. Takimoto contributed equally
to this work.
Conflict of interest: The authors have
declared that no conflict of interest
exists.
Submitted: November 3, 2016
Accepted: June 29, 2017
Published: August 3, 2017
Reference information:
JCI Insight. 2017;2(15):e91588.
https://doi.org/10.1172/jci.
insight.91588.
Nonmyocyte ERK1/2 signaling contributes
to load-induced cardiomyopathy in Marfan
mice
Rosanne Rouf,1 Elena Gallo MacFarlane,2 Eiki Takimoto,1 Rahul Chaudhary,1 Varun Nagpal,2
Peter P. Rainer,1 Julia G. Bindman,2 Elizabeth E. Gerber,2 Djahida Bedja,1 Christopher Schiefer,1
Karen L. Miller,1 Guangshuo Zhu,1 Loretha Myers,2 Nuria Amat-Alarcon,1 Dong I. Lee,1
Norimichi Koitabashi,1 Daniel P. Judge,1 David A. Kass,1 and Harry C. Dietz2,3
1Division of Cardiology, Department of Medicine, and 2McKusick-Nathans Institute of Genetic Medicine, Johns Hopkins
University School of Medicine, Baltimore, Maryland, USA. 3Howard Hughes Medical Institute, Bethesda, Maryland, USA.
Introduction
Heterozygous mutations in the gene encoding fibrillin-1 (FBN1) cause Marfan syndrome (MFS), a system-
ic connective tissue disorder characterized by manifestations in the ocular, skeletal, and cardiovascular
systems, including aortic root aneurysm (1). Heart failure is the leading cause of death in young children
with MFS (2, 3). While the prevailing view is that cardiac decompensation is a consequence of hemody-
namic load imposed by associated mitral valvular regurgitation, some clinical studies of MFS patients
have reported ventricular enlargement or cardiac dysfunction out of proportion to the severity of valvu-
lar disease (4–7), suggesting either a concomitant primary dilated cardiomyopathy (DCM) or myocardial
vulnerability to mechanical stress (8). Elucidation of the mechanism of cardiac dysfunction in MFS has
important implications for the determination of surgical thresholds for valve repair and for development of
novel medical therapies for heart failure.
Fibrillin-1–rich microfibrils are a critical component of the extracellular matrix (ECM) in many tis-
sues (9). The Fbn1C1039G/+ mouse model of MFS harbors a heterozygous cysteine substitution in an epider-
mal growth factor–like domain in fibrillin-1 representative of the most common class of mutation in MFS
patients (1). Such mutations result in normal mRNA expression but decreased fibrillin-1 deposition in
the ECM (1). Studies in Fbn1C1039G/+ mice have suggested that failed matrix sequestration of latent TGF-β,
with consequent increased cytokine activity, drives many manifestations of MFS, including emphysema,
skeletal muscle myopathy, aortic aneurysm, and myxomatous valvular disease (10–13). All of these man-
ifestations are attenuated in Fbn1C1039G/+ mice by pharmacological antagonism of the canonical TGF-β
signaling (Smad2/3) pathway with either TGF-β–neutralizing antibody (NAb) (10–14) or the angiotensin
Among children with the most severe presentation of Marfan syndrome (MFS), an inherited
disorder of connective tissue caused by a deficiency of extracellular fibrillin-1, heart failure is the
leading cause of death. Here, we show that, while MFS mice (Fbn1C1039G/+ mice) typically have normal
cardiac function, pressure overload (PO) induces an acute and severe dilated cardiomyopathy in
association with fibrosis and myocyte enlargement. Failing MFS hearts show high expression of
TGF-β ligands, with increased TGF-β signaling in both nonmyocytes and myocytes; pathologic
ERK activation is restricted to the nonmyocyte compartment. Informatively, TGF-β, angiotensin II
type 1 receptor (AT1R), or ERK antagonism (with neutralizing antibody, losartan, or MEK inhibitor,
respectively) prevents load-induced cardiac decompensation in MFS mice, despite persistent PO. In
situ analyses revealed an unanticipated axis of activation in nonmyocytes, with AT1R-dependent
ERK activation driving TGF-β ligand expression that culminates in both autocrine and paracrine
overdrive of TGF-β signaling. The full compensation seen in wild-type mice exposed to mild PO
correlates with enhanced deposition of extracellular fibrillin-1. Taken together, these data suggest
that fibrillin-1 contributes to cardiac reserve in the face of hemodynamic stress, critically implicate
nonmyocytes in disease pathogenesis, and validate ERK as a therapeutic target in MFS-related
cardiac decompensation.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
2
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
II type 1 receptor (AT1R) blocker (ARB), which suppresses expression of TGF-β receptors and ligands
(12, 14). AT1R-dependent signaling cascades also converge on the ERK1/2 pathway (15), which has also
been shown to be activated in MFS aortas; selective inhibition of ERK1/2 activation using RDEA119
(refametinib) rescues aortic growth and wall architecture in MFS mice (16). Intriguingly, tissue analysis
in non-MFS rodent and human heart failure reveals evidence for increased TGF-β signaling (17–20) as
well as increased deposition of collagens type I and III (21, 22), which are encoded by TGF-β–responsive
genes. Furthermore, in myocardial cells, TGF-β expression and secretion can be induced by angiotensin
II (AngII) signaling (23) and mechanical stretch (24), respectively, and have been shown to drive the tran-
scriptional changes that accompany the transition from compensated load-induced hypertrophy to heart
failure (25–27).
In the myocardium, fibrillin-1 is a predominantly nonmyocyte-derived ECM protein (28, 29) that is
localized in perivascular and perimysial spaces, closely paralleling the spatial arrangement of collagen
(20, 30). Recent studies in Fbn1C1039G/+ mice have shown late-onset development of DCM, which uniformly
associates with valvular regurgitation and increased myocardial ERK1/2 signaling (31); however, these
studies do not segregate the effects of perturbed hemodynamic load from intrinsic defects in the myocar-
dium. Similarly, Fbn1mgR/mgR mice, in which a homozygous hypomorphic allele of Fbn1 causes severe defi-
ciency of fibrillin-1 throughout development, develop DCM associated with severe valvular regurgitation
(32). Conditional deletion of Fbn1, using homozygous Fbn1lox/lox mice in combination with a cardiomyo-
cyte-specific αMHC-Cre-recombinase driver (Fbn1αMHC–/–), also associates with structural and functional
changes consistent with DCM (32). Although this latter study seems to point to an intrinsic defect in
cardiomyocytes as the primary driver of DCM in the context of profound deficiency of fibrillin-1 present
throughout development, the relevance of these findings to the broad clinical spectrum of MFS patients
with heterozygous mutations that impose a relative and postnatally progressive deficiency of extracellular
fibrillin-1 remains unclear.
To address these issues, we studied young Fbn1C1039G/+ mice (prior to valvular or myocardial dysfunc-
tion) in combination with a conditional provocation (transverse aortic constriction [TAC]) to impose acute
hemodynamic load. Comprehensive assessment of cardiac function, gross morphology, and cellular sig-
naling events in tight temporal sequence with TAC allowed a precise definition of gene-by-environment
interactions in the pathogenesis of MFS-associated DCM.
Results
Fbn1C1039G/+ mice are predisposed to load-induced heart failure. We performed serial echocardiography and pres-
sure-volume (PV) loop analysis in Fbn1C1039G/+ mice. Echocardiography of Fbn1C1039G/+ mice showed that the
presence of valvular regurgitation (mitral and/or aortic) associated directly with increased end-diastolic
(r = 0.82, P < 0.001) and end-systolic (r = 0.84, P < 0.0005) diameters and inversely with fractional short-
ening (r = –0.7792, P < 0.005) (Supplemental Figure 1A; supplemental material available online with this
article; https://doi.org/10.1172/jci.insight.91588DS1). However, in mice that did not manifest valvular
regurgitation, as assessed by color flow Doppler, cardiac dimensions and function were indistinguishable
between WT and Fbn1C1039G/+ mice (Supplemental Figure 1A). PV loop analysis, which isolates assessment
of cardiac function from loading conditions on the ventricle (33), also showed that systolic and diastolic
parameters were identical between WT and Fbn1C1039G/+ hearts in the absence of valvular regurgitation at
4 or 12 months of age (Supplemental Figure 1, B and C). Finally, up to 12 months of age, heart weights
(Supplemental Figure 1D) and left ventricular activation of Smad2 (phosphorylated [pSmad2]) or ERK1/2
(phosphorylated [pERK1/2]) was unaltered in Fbn1C1039G/+ mice (Supplemental Figure 2, A–C), as were
expression levels of prototypical TGF-β target genes Ctgf and Serpine1 (Supplemental Figure 2D). Impor-
tantly, these findings suggest that, in the unstressed heart, cardiac structure and function is normal in the
Fbn1C1039G/+ mouse model of MFS.
To study whether the cardiac response to increased hemodynamic load is altered in Fbn1C1039G/+ hearts, we
exposed WT and littermate Fbn1C1039G/+ mice without valvular regurgitation to TAC (referred to as WT:TAC
and Fbn1C1039G/+:TAC mice), using a magnitude of aortic constriction titrated to cause a mild pressure over-
load (PO). Baseline cardiac dimensions and function were indistinguishable between Fbn1C1039G/+ and WT
control mice (Supplemental Figure 3A). However, after 1 week of TAC, Fbn1C1039G/+ mice uniquely showed
marked ventricular dilatation and dysfunction, which progressed in the ensuing 3 weeks (Figure 1, A and B).
Rare Fbn1C1039G/+ and WT mice that had developed valvular regurgitation were excluded from all analyses.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
3
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
Postmortem analysis after 4 weeks of TAC revealed a profound increase in heart size and mass in Fbn1C1039G/+
mice compared with WT mice subjected to equivalent TAC (Figure 1C). Marked lung congestion (wet lung
weight) (Figure 1C) occurred as an expected correlate with left heart failure in Fbn1C1039G/+:TAC mice. More-
over, histological analysis disclosed a marked increase in cardiomyocyte cross-sectional area and interstitial
fibrosis in Fbn1C1039G/+:TAC hearts when compared with WT animals subjected to TAC (Figure 1, D and E,
and Supplemental Figure 3, B and C). The expression of myocyte-specific hypertrophy-related genes, atrial
Figure 1. Fbn1C1039G/+ mice develop marked heart failure acutely from pressure overload. (A) Weekly temporal changes in cardiac dimensions and function
after TAC. EDD, end-diastolic diameter; ESD, end-systolic diameter; FS, fractional shortening; LV mass, left ventricular mass; BL, baseline. Comparison
pairs shown in brackets on right side of panels. n = 4–8 per group. (B) Representative M-mode echocardiogram 4 weeks after TAC. White arrow, end-di-
astolic diameter; yellow arrow, end-systolic diameter. (C) Representative whole hearts and summary results for heart weight normalized to tibia length
(HW/TL) in Fbn1+/+ and Fbn1C1039G/+ mice subjected to 4 weeks of TAC. Wet lung weight (LW) was normalized to TL. n ≥ 5 per group. *P < 0.05, **P < 0.01,
***P < 0.001; ##P < 0.01 vs. Fbn1+/+:sham; ††P < 0.01 vs. Fbn1C1039G/+:sham, 2-way ANOVA, Bonferroni’s correction. (D and E) Representative hematoxylin
and eosin and Masson’s trichrome staining of formalin-fixed heart sections from Fbn1+/+ and Fbn1C1039G/+ mice subjected to 4 weeks of TAC. Scale bars: 75
μm (D); 150 μm (E). White dotted line (in D) outlines myocyte circumference. Blue stain (in E) indicates fibrosis.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
4
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
natriuretic peptide (Nppa), brain natriuretic peptide (Nppb), and β–myosin heavy chain (Myh7) (26, 34), was
significantly increased in Fbn1C1039G/+ hearts when compared with WT hearts subjected to TAC (Supplemen-
tal Figure 3D) but was comparable between sham groups. We found that though cardiac expression of Fbn1
mRNA was significantly increased in Fbn1C1039G/+ mice and WT controls in response to TAC (Figure 2A),
deposition of fibrillin-1 protein in the interstitial pericapillary and perivascular spaces adjacent to cardiac
Figure 2. Fibrillin-1 protein fails to deposit in myocardium in load-induced heart failure in Fbn1C1039G/+ mice. (A) mRNA expression of Fbn1 normalized to
Gapdh and then to Fbn1+/+:sham data, assessed by real-time RT-PCR. n ≥ 4 per group. (B) Representative images of immunofluorescent fibrillin-1 staining
(red) of heart sections from Fbn1+/+ and Fbn1C1039G/+ mice subjected to 4 weeks of TAC. Red, fibrillin-1 (interstitial space); blue, DAPI (nuclei); green, lipofus-
cin (myocytes). Scale bar: 20 μm. (C) Quantitative representation of fibrillin-1 normalized to unit area. n = 4–7 per group. (D and E) Representative Western
blot and summary quantification for phosphorylated/total (p/t) Smad2 and ERK1/2 using left ventricular tissue lysates. Lanes were run on the same gel
but were noncontiguous. n ≥ 4 per group. (F) mRNA expression of TGF-β targets, normalized to 18S and then to Fbn1+/+ data, assessed by real-time RT-PCR
from left ventricular tissue. Col1a1, collagen type 1 a1; Fn1, fibronectin; Serpine1, plasminogen activator inhibitor, type 1; Postn, periostin; Ctgf, connective
tissue growth factor; Spp1, osteopontin. n = 3–6 per group. *P < 0.05, **P < 0.01, ***P < 0.001, 1-way ANOVA, Tukey’s correction. In box-and-whisker plots,
the lower and upper margins of each box define the 25th and 75th percentiles, respectively; the internal line defines the median, and the whiskers define
the range. Values outside 1.5 times the interquartile distance are shown.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
5
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
nonmyocytes was increased in the compensated WT heart but not in decompensated Fbn1C1039G/+ hearts
exposed to TAC (Figure 2, B and C, and Supplemental Figure 4). Thus, while MFS hearts have no demon-
strable cardiac dysfunction at baseline, they do display an abnormal response to mechanical load, revealing
that fibrillin-1 plays a critical role in the adaptive response to hemodynamic stress.
Canonical TGF-β signaling contributes to load-induced cardiac decompensation in Fbn1C1039G/+ mice. To gain insight
into the mechanism of load-induced heart failure in Fbn1C1039G/+ mice, we measured changes in the activation
of the Smad2 and ERK1/2 signaling pathways, which are known to mediate the progression of aortic aneu-
rysm in MFS mice (12, 16). Immunoblots showed increased levels of pSmad2 and pERK1/2 in Fbn1C1039G/+
Figure 3. Smad2 and ERK1/2 activation are dierentially increased in the myocyte and nonmyocyte compartments of failing Fbn1C1039G/+ hearts. (A)
Representative sections of pSmad2 (red) immunostaining. Scale bar: 20 μm. White arrowheads point to examples of positively stained myocyte nuclei.
(B) Summary quantification of total intensity normalized to Fbn1+/+:sham and labeling index (percentage positive cells) of pSmad2 in the nonmyocyte and
myocyte compartments. (C and D) Representative sections and summary quantification of pERK1/2 (red) immunostaining. Scale bar: 20 μm. n = 6 fields
per group. **P < 0.01, ***P < 0.001, 1-way ANOVA, Tukey’s correction.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
6
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
hearts when compared with WT controls after 4 weeks of TAC (Figure 2, D and E). Increased expression of
prototypical TGF-β–driven gene products, such as Col1a1, Postn, Serpin1, Ctgf, Spp1, and Fn1 (Figure 2F), was
also observed in Fbn1C1039G/+:TAC hearts when compared with WT:TAC controls. In situ analyses of heart
sections by immunofluorescence showed increased pSmad2 and pERK1/2 in Fbn1C1039G/+:TAC hearts when
compared with WT:TAC controls (Figure 3). Furthermore, in Fbn1C1039G/+:TAC hearts, pSmad2 colocalized
Figure 4. Treatment with TGF-β NAb, losartan, or MEKi improves load-induced cardiac decompensation in Fbn1C1039G/+ hearts, despite persistent
load. (A) Temporal changes of cardiac dimensions and function of Fbn1C1039G/+ hearts after TAC, with treatment arms. Fbn1+/+, blue circle with solid line;
Fbn1C1039G/+, red square with solid line; Fbn1C1039G/+ with TGF-β NAb treatment, orange triangle with dotted line; Fbn1C1039G/+ with losartan (Los) treatment,
green diamond with dotted line; Fbn1C1039G/+ with MEK1/2 inhibitor (MEKi) treatment, purple circle with dotted line. Early time point, 1 week after TAC; mid
time point, 2–4 weeks after TAC; end time point, 3–5 weeks after TAC. EDD, end-diastolic diameter; ESD, end-systolic diameter; FS, fractional shortening;
LV mass, left ventricular mass; BL, baseline. Comparison pairs shown in brackets on right side of respective panel. n ≥ 5 per group. (B) Representative
M-mode echocardiograms at end time point, Fbn1C1039G/+ sham vs. TAC groups, with and without treatment. (C) Summary quantification of heart weight
normalized to tibia length (HW/TL). n ≥ 5 per group. *P < 0.05, **P < 0.01, ***P < 0.001, 1-way ANOVA, Tukey’s correction.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
7
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
with both nonmyocyte and myocyte compartments (Figure 3, A and B; Supplemental Figure 5, A and B, top;
and Supplemental Figure 5, C and D, left), whereas enhanced pERK1/2 colocalized predominantly with the
nonmyocyte compartment (Figure 3, C and D; Supplemental Figure 5, A and B, bottom; and Supplemental
Figure 5, C and D, right). Further characterization of pSmad2-positive and pERK1/2-positive nonmyocytes
showed positive costaining of vimentin in a subset of cells (Supplemental Figure 5D), suggesting nonmyo-
cytes to be of fibroblast, endothelial, or myeloid origin (35).
Figure 5. Functional improvement of mechanically loaded Fbn1C1039G/+ hearts correlates with reduced fibrosis. (A) Representative Masson’s trichrome–
stained heart sections of Fbn1C1039G/+ sham vs. TAC groups, with and without treatment, showing fibrosis stained in blue. Scale bars: 750 μm (top); 100 μm
(bottom). (B) Quantification of fibrosis, quantified by total number of blue pixels normalized to whole tissue area and then to Fbn1+/+:sham data. n = 4–7
per group. (C) mRNA expression of TGF-β–responsive fibrogenic genes, normalized to Gapdh and then to Fbn1+/+:sham data, assessed by real-time RT-PCR.
Col1a1, collagen type 1 a1; Spp1, osteopontin. n = 4–6 per group. *P < 0.05, **P < 0.01, ***P < 0.001, 1-way ANOVA, Tukey’s correction. In box-and-whisker
plots, the lower and upper margins of each box define the 25th and 75th percentiles, respectively; the internal line defines the median, and the whiskers
define the range. Values outside 1.5 times the interquartile distance are shown.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
8
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
We next assessed for contributions of TGF-β, AT1R, and ERK1/2 signaling by pretreating Fbn1C1039G/+
mice systemically with a pan-specific TGF-β–neutralizing polyclonal antibody (NAb) (12), an ARB (losar-
tan) (12), or a selective MEK1/2 inhibitor (MEKi, PD98059 or RDEA119) (16, 36), respectively. Cultured
cardiac fibroblasts were used to show that MEKi antagonizes phosphorylation of ERK1/2 but not Smad2
(Supplemental Figure 6) as expected (16, 37, 38).
Serial echocardiography showed that NAb, losartan, or MEKi treatment substantially improved end-di-
astolic and end-systolic dimensions, fractional shortening, and left ventricular mass in Fbn1C1039G/+:TAC
hearts when compared with vehicle-treated counterparts (Figure 4). Therapeutic effects correlated with
a significant reduction in fibrotic area (Figure 5, A and B). Furthermore, TAC-induced upregulation of
TGF-β–responsive fibrogenic genes, collagen (Col1a1) and osteopontin (Spp1) (39–41), was attenuated by
NAb treatment and abrogated by either losartan or MEKi (Figure 5C). Curiously, myocyte size, which
is increased in Fbn1C1039G/+:TAC hearts, was normalized upon treatment with losartan or MEKi but not
with NAb (Supplemental Figure 7, A–C). Paralleling these observations, the hypertrophic gene program
induced by TAC in Fbn1C1039G/+ mice was unaltered by NAb treatment but was significantly suppressed by
either losartan or MEKi (Supplemental Figure 7D). Thus, the comparable therapeutic efficacy of NAb,
losartan, and MEKi associate with a reduction of fibrosis but not with myocyte hypertrophy.
We next examined Smad2 and ERK1/2 signaling in Fbn1C1039G/+ hearts subjected to TAC and treated
with NAb, losartan, or MEKi in an attempt to define the pathogenic sequence that culminates in cardiac
decompensation. The therapeutic efficacies of NAb, losartan, and MEKi correlated with reduced levels
of pSmad2 (Figure 6, A–D), supporting the notion that proximal events in TGF-β signaling contribute to
load-induced heart failure. However, with NAb treatment, pERK1/2 levels were unaltered (Figure 6, A and
B), suggesting that ERK1/2 activation occurs independently or upstream of TGF-β signaling. Informa-
tively, MEKi treatment normalized both pERK1/2 and pSmad2 levels, suggesting that ERK1/2 signaling
contributes directly or indirectly to induction of TGF-β canonical signaling. Losartan also normalized both
pERK1/2 and pSmad2 levels, suggesting an AT1R:ERK1/2:Smad2 axis of activation (Figure 6, C–E). In
further support of an AT1R:ERK1/2:Smad2 axis of activation, we show that AngII treatment of murine
cardiac fibroblasts after 24 hours results in increased activation of ERK1/2 but not Smad2 (Supplemental
Figure 8). Whether or not these events represent cell-autonomous relationships remained to be determined.
Losartan or MEKi potently suppresses Smad2 activation in both nonmyocyte and myocyte compartments. To
determine relationships of signaling pathways in nonmyocyte and myocyte compartments, we performed
in situ analysis of pSmad2 and pERK1/2 in Fbn1C1039G/+:TAC hearts treated with NAb, losartan, or MEKi.
NAb treatment achieved a marked decrease in nonmyocyte Smad2 activation, whereas myocyte Smad2
activation largely persisted (Figure 7A and Supplemental Figure 9A). In contrast, treatment with both
losartan and MEKi resulted in full normalization of pSmad2 in both cellular compartments. Losartan or
MEKi achieved a marked decrease in nonmyocyte ERK1/2 activation; however, NAb had no effect (Figure
7B and Supplemental Figure 9B). All three treatments reduced vimentin reactivity (Figure 7) and prevented
TAC-induced expansion of the nonmyocyte compartment in Fbn1C1039G/+ mice (Supplemental Figure 9C).
The failure of NAb to suppress myocyte Smad2 signaling likely reflects a bioavailability issue as previously
discussed (42). Taken together, these data are consistent with a model that invokes a primary increase in
AT1R-mediated ERK1/2 activation in nonmyocytes that is required for subsequent Smad2 activation in
both nonmyocytes and myocytes.
ERK1/2 activation is pivotal in the progression of load-induced heart failure and upregulates TGF-β ligand expres-
sion. We postulated that ERK1/2 signaling may induce nonmyocyte TGF-β ligand production, thereby driv-
ing canonical TGF-β signaling in both an autocrine and paracrine manner. In keeping with this hypothesis,
we found that TAC increased expression of Tgfb2 and Tgfb3 in Fbn1C1039G/+ hearts that was normalized with
losartan or MEKi but not NAb (Figure 8). Furthermore, in situ analysis demonstrated that the expression
of TGF-β3 mRNA localized predominantly to the nonmyocyte compartment (Supplemental Figure 10, A
and B), similarly to that of vimentin, a mesenchymal cell marker that is upregulated in fibroblast-to-myofi-
broblast transformation (43) (Supplemental Figure 10C). Cultured cardiac fibroblasts demonstrate a 12-fold
higher expression of Tgfb3 compared with cultured cardiomyocytes (Supplemental Figure 10D). AngII treat-
ment of cardiac fibroblasts after 12 hours resulted in ERK1/2 activation but not direct Smad2 activation,
while TGF-β3 treatment showed intact Smad2 activation (Supplemental Figure 11). Furthermore, cardiac
fibroblasts cultured from Fbn1C1039G/+ hearts demonstrated even greater sensitivity to ERK1/2 activation in
response to AngII treatment than WT hearts (Supplemental Figure 11). Taken together, these data identify a
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
9
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
mechanism by which ERK1/2 activation may promote the deleterious effects of TGF-β signaling. This may
be particularly important in the fibrillin-1–deficient myocardium of the hemodynamically loaded MFS heart.
Discussion
While prior reports have suggested a tendency for cardiac decompensation in MFS, it has been difficult
to determine with confidence if this relates to an intrinsic myocardial predisposition, the influence of
valve dysfunction, or a combination of the two. This ambiguity relates to the high frequency of mitral
Figure 6. ERK1/2 activation is upstream of Smad2 activation in Fbn1C1039G/+ hearts. (A–D) Representative Western blot and summary quantification for
Smad2 and ERK1/2 phosphorylation from left ventricular tissue lysates. n = 4–6 per group. *P < 0.05, **P < 0.01, ***P < 0.001, 1-way ANOVA, Tukey’s
correction. (E) Diagram illustrating the pathogenic sequence of load-induced heart failure in the fibrillin-1–deficient myocardium of Fbn1C1039G/+ mice. NM,
nonmyocytes; M, myocytes. In box-and-whisker plots, the lower and upper margins of each box define the 25th and 75th percentiles, respectively; the
internal line defines the median, and the whiskers define the range. Values outside 1.5 times the interquartile distance are shown.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
10
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
and/or aortic valve regurgitation in people and mice with MFS that can be difficult to quantitate and the
use of mouse models with targeted disruption of both Fbn1 alleles that show accelerated cardiovascular
disease (32) but uncertain mechanistic relevance to Marfan patients with heterozygous mutations and a
deficiency of extracellular fibrillin-1 that progresses postnatally. In keeping with a critical contribution
of hemodynamic stress, heart failure and infantile death in children with the most severe presentation of
MFS uniformly associate with significant volume overload (44, 45) and can be delayed or even prevented
by aggressive surgical intervention for valve dysfunction (46–48). In this study, we utilized a standardized
PO (via TAC) as a conditional provocation in young mice heterozygous for a typical MFS-associated
missense mutation in fibrillin-1. Importantly, these animals had no abnormality of heart structure or
function at the onset of the study, and only mice that did not have any valvular regurgitation during
Figure 7. Inhibition of ERK1/2 activation suppresses Smad2 activation in both nonmyocyte and myocyte compartments. Representative coimmu-
nostaining of (A) pSmad2 (red, row 1) and vimentin (pink, row 2) and of (B) pERK1/2 (red, row 1) and vimentin (pink, row 2) of myocyte-enriched sec-
tions of Fbn1C1039G/+ hearts subjected to sham vs. TAC, with treatment groups. Scale bars: 20 μm; zoom inset: ×2 of nonmyocyte-enriched areas. Blue,
DAPI (nuclei); green, lipofuscin (myocytes).
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
11
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
the entire follow-up period were analyzed. MFS mice
exposed to TAC showed acute and profound DCM and
heart failure, while their WT counterparts remained ful-
ly compensated. Furthermore, decompensation in MFS
associated with aberrant activation of cellular signal-
ing in the myocardium, including both the TGF-β and
ERK1/2 pathways. Pharmacologic attenuation of aber-
rant signaling with TGF-β, AT1R, or ERK1/2 antago-
nists preserved heart structure and function, despite a
fixed PO, suggesting that predisposition does not simply
relate to a chronic deficiency in the structure and hence
mechanical integrity of the ECM, but rather reflects
dynamic but modifiable compensatory cellular signaling
events that attend myocardial remodeling. These data
suggest a number of conclusions with strong potential
for clinical relevance. First, the concept that the myocar-
dium is sensitized to acute hemodynamic stress in MFS
supports consideration of more frequent cardiovascular
surveillance for individuals with volume overload and
aggressive avoidance of hypertension for indications
beyond suppression of aneurysm progression. Second,
the observation that mild cardiovascular stress selec-
tively tips the MFS heart into profound failure supports
the earlier application of methods to more accurately
quantify valvular dysfunction (such as cardiac MRI) in
patients with MFS. The thresholds used to determine
when valve repair or replacement should be performed
should be reconsidered in this context. Finally, as dis-
cussed below, a refined understanding of the molecular
pathogenesis of heart failure in MFS may allow the test-
ing and implementation of novel medical therapies for
this and related disorders.
Prior work performed using a mouse model that is
homozygous for a hypomorphic Fbn1 allele that expresses
normal but reduced levels of Fbn1, diminishing the total
amount of produced fibrillin to about 15% of normal
from the time of conception, attributed the relatively mild
cardiac dysfunction that was observed to altered mecha-
nosensing and ultimately enhanced ERK1/2 signaling by
cardiomyocytes (32). The focus on this cell type stemmed from the observation of DCM in a mouse line in
which both copies of a conditional (flanked by loxP) Fbn1 allele were deleted through expression of Cre-re-
combinase using a cardiac myosin heavy chain promoter element (αMHC-Cre). These experiments were
difficult to interpret with certainty; however, the use of this Cre driver was also associated with a substantial
(~50%) reduction in fibrillin-1 expression in cardiac fibroblasts that was equivalent to that observed in mice
Figure 8. Inhibition of ERK1/2 activation suppresses expres-
sion of all three TGF-β isoforms. mRNA expression of Tgfb1
(TGF-β1), Tgfb2 (TGF-β2), and Tgfb3 (TGF-β3) isoforms, nor-
malized to Gapdh and then to Fbn1+/+:sham data, assessed by
real-time RT-PCR. n = 4–6 per group. *P < 0.05, **P < 0.01, ***P
< 0.001, 1-way ANOVA, Tukey’s correction. In box-and-whisker
plots, the lower and upper margins of each box define the 25th
and 75th percentiles, respectively; the internal line defines the
median, and the whiskers define the range. Values outside 1.5
times the interquartile distance are shown.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
12
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
with germline haploinsufficiency for Fbn1 (32). Furthermore, isolated expression of Cre-recombinase in the
myocardium has previously been associated with predisposition for cardiomyopathy (49, 50). While control
experiments by Cook and colleagues suggested that neither isolated Fbn1 haploinsufficiency nor Cre-recom-
binase expression was sufficient to induce cardiomyopathy in their hands, the potential pathogenic influence
of these events in combination was not tested (32). Our study now documents overt pathogenic significance
for ERK1/2 signaling in load-induced heart failure in Marfan mice, given that specific antagonism with
a MEKi prevented all perturbations of myocardial architecture and function. A number of observations,
however, focus attention on the nonmyocyte compartment. First, application of PO in WT animals was
associated with pronounced upregulation of fibrillin-1 expression, with increased deposition in the nonmyo-
cyte compartment; failure of this event in MFS mice correlated with adverse cardiac remodeling and failure.
Second, our application of in situ–based methods revealed that excessive ERK1/2 activation is specific to
nonmyocytes in the failing hearts of MFS mice. Finally, the attenuation of disease seen upon treatment with
TGF-β NAb was associated with inhibition of canonical Smad signaling in nonmyocytes but not myocytes,
presumably due to limited bioavailability in the latter, as has previously been reported (42).
The pathogenic sequence for load-induced heart failure in nonmyocytes of MFS mice appears to ini-
tiate with increased ERK1/2 activation through AT1R signaling. In theory, this could relate to increased
expression of factors along the angiotensin-converting enzyme/AngII/AT1R axis in response to the com-
bined influence of an altered matricellular environment and mechanical stress. What is both clear and
surprising is that ERK1/2 signaling drives TGF-β ligand production (largely TGF-2 and TGF-β3) by
inference in nonmyocytes and increases canonical TGF-β signaling in both the nonmyocyte and myocyte
compartments. This is in contrast to previous studies in the aorta of MFS mouse models that have shown
that excessive ERK1/2 activation is, at least in part, dependent upon and therefore downstream of TGF-β
(14, 16, 51). Of greater importance, this work shows that, similar to findings in mice with hypertrophic
cardiomyopathy (52), load-induced nonmyocyte expansion, fibrosis, and heart failure in MFS mice are
greatly attenuated by specific antagonism of TGF-β using NAb; the substantial residual Smad2 activa-
tion in myocytes but not nonmyocytes (likely on the basis of differential bioavailability of NAb) infers
specific pathogenic significance in the progression of heart failure for the latter. Prevention of myocyte
hypertrophy parallels the attenuation of myocyte Smad2 signaling seen with treatment with either ARB
or MEKi (but not with TGF-β NAb) and suggests that AT1R- and ERK1/2-dependent nonmyocyte secre-
tion of TGF-β ligands drives hypertrophy via paracrine mechanisms, while driving fibrosis and failure in
an autocrine fashion. These data support a primary role for extracellular fibrillin-1 in compensating for
ventricular injury and stress and help to establish the nonmyocyte microenvironment as a critical factor
in the loaded MFS heart. These findings may also inform mechanisms relevant to other presentations of
heart failure beyond MFS.
Methods
Mice. Fbn1C1039G/+ mice were developed as previously described (1) and bred on a pure C57BL/6 background
(backcrossed for >9 generations).
TAC and drug treatments. PO was performed by surgical placement of suture around the transverse
aorta as described previously (53). A 7-0 ligature was placed around the vessel using a 26-gauge needle to
ensure consistent occlusion. This needle size was chosen to elicit a mild response, as initial studies using
our standard TAC model (27-gauge needle) led to increased mortality in Fbn1C1039G/+ mice. Size-, age-,
and sex-matched (male) Fbn1C1039G/+ mice and littermate Fbn1+/+ mice were randomized to receive TAC
versus sham surgery (same procedure but without constriction).
For drug treatment studies, size-, age-, and sex-matched (male) Fbn1C1039G/+ mice were randomized to
receive drug treatment versus vehicle as follows: rabbit polyclonal TGF-β NAb (pan-specific against TGF-β1,
TGF-β2, and TGF-β3) and control IgG (R&D Systems) were reconstituted in PBS and administered via i.p.
injection at a dose of 10 mg/kg every 10 days. The first dose was given 2 days prior to TAC. The MEKi,
PD98059 (Calbiochem), was reconstituted in 25% DMSO/50% ethanol in PBS and administered daily i.p.
at a dose of 3 mg/kg, or placebo (DMSO) by was administered by i.p. injection, 5 days a week (36). The
first dose was given 1 day prior to TAC. RDEA119 (NIH/National Human Genome Research Institute,
Selleck Chemicals) was reconstituted in 10% 2-hydroxypropyl-β-cyclodextrin (Sigma-Aldrich) in PBS and
administered twice daily by oral gavage at a dose of 25 mg/kg. The first dose was given 1 day prior to TAC.
Losartan was administered daily in drinking water at a dose of 0.6 g/l as previously described (12). Mice
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
13
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
were euthanized at 3 weeks (TGF-β NAb, PD98059, and losartan) or 5 weeks (RDEA119) for tissue analy-
sis. The studies and analyses were performed blinded as to experimental group.
Echocardiography. All transthoracic echocardiography was performed in conscious mice. To assess the
presence of valvular regurgitation, high-frame rate color Doppler imaging was performed before TAC and
at the study end (Vevo 2100, 40-MHz mechanical transducer, VisualSonics). Rare mice that developed
valvular regurgitation after TAC were excluded from all after TAC analyses. M-mode echocardiography
followed reported methods (42).
In vivo hemodynamics. Cardiac function and arterial loading were assessed by PV analyses as described
previously (33, 53). Briefly, mice were anesthetized, intubated, and mechanically ventilated. The LV apex
was exposed between the seventh and eighth ribs, and a 1.4-Fr PV catheter (SPR 839; Millar Instruments
Inc.) was advanced through the apex to lie along the longitudinal axis. Absolute volume was determined by
saline calibration and aortic ultrasound, and data were assessed at steady state and during preload reduc-
tion. Data were analyzed with custom software.
Molecular analyses. Protein electrophoresis and immunoblot assays followed standard procedures, using
cell lysis buffer (Cell Signaling) containing protease inhibitor and phosphatase inhibitor cocktail (Milli-
pore). Assays were run on 10% Bis-Tris gels (Bio-Rad), subjected to SDS-PAGE, and transferred to PVDF
membranes. Membranes were probed for pSmad2 (3101, Cell Signaling), pERK1/2 (4377, Cell Signaling),
GAPDH (ab9483, Abcam), tSmad2 (3103, Cell Signaling), and tERK1/2 (4695, Cell Signaling), followed
by labeling with anti-rabbit/goat secondary antibody and visualizing by either enhanced chemilumines-
cence (SuperSignal West Femto, Life Technologies) or fluorescence detection (LI-COR Odyssey). Blots
were quantified using ImageJ software (NIH) or the LI-COR Odyssey system.
Quantitative PCR using RT-PCR by amplification (Applied Biosystems) was used to assess RNA
expression using standard procedures. The following TaqMan probes (Applied Biosystems) were used:
Mm00514908_m1 (Fbn1), Mm00600555_m1 (Myh7), Mm01255747_g1 (Nppa), Mm00435304_g1 (Nppb),
Mm99999915_g1 (Gapdh), Mm01178820_m1 (Tgfb1), Mm00436955_m1 (Tgfb2), Mm00436960_m1
(Tgfb3), Mm00801666_g1 (Col1a1), Mm01256744_m1 (Fn1), Mm00436767_m1 (Spp1), Mm00435858_m1
(Serpine1), Mm01192933_g1 (Ctgf), and Mm00450111_m1 (Postn). Relative quantification for each tran-
script was obtained by normalizing against housekeeping control transcript.
Histology. Myocardium was fixed with 10% formaldehyde, paraffin embedded, and sectioned into
5-μm-thick slices. Masson’s trichrome staining was used to visualize fibrosis. Quantification of fibrosis
content was performed in whole heart sections using the pen tool in Aperio ImageScope (Leica Bio-
systems) to exclude large vessels, epicardium, or pericardium and automated morphometric analysis of
digital images and a previously validated modified positive pixel algorithm to quantify the amount of Tri-
chrome (blue) stain (54) (Aperio ImageScope). Quantification of myocyte cross-sectional area, when per-
formed in hematoxylin and eosin–stained sections, was performed with ImageJ. For wheat germ aggluti-
nin (WGA) staining, slides were deparaffinized, rehydrated, and subjected to citrate-based heat-mediated
antigen retrieval as previously described in (55). Sodium borohydride solution (1 mg/ml in PBS) was used
for quenching. Slides were incubated with Alexa Fluor 647–preconjugated WGA (W32466, Invitrogen)
overnight at 4°C and mounted using Prolong Gold mounting medium (P36934, Thermofisher). Image
acquisition was performed on an EVOS epifluorescence microscope (Life Technologies). Quantification
of myocyte cross-sectional area was performed using an automated algorithm with Image J, analyzing
1,040 ± 70 cells from 6–16 areas per mouse heart.
Immunohistochemistry. Slides with 5-μm-thick sections were deparaffinized with xylene and hydrated in
a graded alcohol series. Slides were immersed in boiling antigen retrieval solution (10 mM sodium citrate
buffer, pH 6.0) for 10 minutes and then set aside to cool in the same solution at room temperature for 30
minutes. Slides were incubated in freshly prepared sodium borohydride solution (10 mg/ml PBS) for 20
minutes, incubated in 100 mM glycine TBT (TBS containing 0.1% Triton-X 100) for 20 minutes, and then
treated with Fc block (NB309) and Background-buster (NB306) from Innovex Biosciences according to
the manufacturer’s instructions. Slides were incubated in a humid chamber overnight at 4°C in the fol-
lowing primary antibodies, as applicable: anti-fibrillin-1 (1:100, ab53803, Abcam), anti-pSmad2 (1:100,
4-953, Millipore), and anti-pERK1/2 (1:100, 4370, Cell Signaling) diluted in TBT. After washing, slides
were then incubated in a humid chamber with Alexa Fluor 594–conjugated donkey anti-rabbit IgG sec-
ondary antibody (1:500, R37119, Life Technologies) at room temperature for 30 minutes, washed, and
then mounted in VECTASHIELD HardSet Mounting Medium with DAPI (H-1500, Vector Laboratories).
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
14
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
When applicable, following staining with the secondary antibody, slides were further incubated with Alexa
Fluor 647–preconjugated anti-vimentin antibody (1:100, Cell Signaling, 9856). Image acquisition was per-
formed on a Zeiss AxioExaminer with 710NLO-Meta multiphoton confocal microscope at magnifications
of ×20 and ×40.
High-resolution images of immunostaining (red fluorescence) were quantified using ImageJ. After
background correction, fibrillin-1–positive staining was quantified by signal intensity (red) per unit area.
The quantification of pSmad2 and pERK1/2 (red fluorescence) staining in myocytes and nonmyocytes
was performed using ImageJ. Myocytes, known to exhibit robust lipofuscin autofluorescence and sarco-
meric Z-lines in sodium borohydride-treated formaldehyde-fixed sections (56), were defined by the fol-
lowing characteristics: (a) cells with circular or oval-shaped nuclei (blue channel) centrally located with-
in myocytes (green channel), (b) relatively large cell size, and (c) presence of sarcomeric Z-lines (56).
Nonmyocytes were otherwise defined as all remaining cells with identifiable nuclei located peripheral
to myocytes, with minimal autofluorescence, in addition to being located in the interstitial space. The
number of positive cells (staining red for pSmad2 or pERK1/2) per total cells was manually counted in
each compartment by two blinded observers in 10–15 areas per treatment condition.
In situ RNA hybridization. RNAscope technology (Advanced Cell Diagnostics) was used to perform in situ
mRNA hybridization on formalin-fixed paraffin-embedded slides following the manufacturer’s instructions.
Briefly, 5-μm-thick tissue sections were hybridized with Tgfb3 (406211) and Vim (457961) from RNAscope
probes (Advanced Cell Diagnostics). A negative control probe and a positive control probe served as techni-
cal quality controls. Multiple images were acquired in Z-stack mode using a Zeiss 780 Laser Scanning Con-
focal Microscope. The in situ hybridization images are maximum intensity projections of acquired images.
Isolation of cardiac fibroblasts. Neonatal ventricles from 3-day-old mice or rats were separated and minced
in ice-cold balanced salt solution, as described previously with minor modifications (57). To isolate cardiac
cells, the tissues were incubated in a balanced salt solution containing 0.2% collagenase type 2 (Worthing-
ton Biochemical) for 6 minutes at 37°C with agitation. The digestion buffer was replaced 7 times, at which
point the tissues were completely digested. The dispersed cells were incubated in T25 flasks for 120 minutes
to plate nonmyocytes. Nonmyocytes that attached to the dishes were cultured in DMEM supplemented
with 10% FBS and allowed to grow to confluence; then they were trypsinized and passaged at 1:4. This
procedure yielded cell cultures that were almost exclusively fibroblasts by the first passage. Experiments
were carried out after 1 additional passage.
Statistics. Data are presented as dot plots, with mean ± SEM, or as box-and-whisker plots, with the
lower and upper margins of each box defining the 25th and 75th percentiles, respectively; the internal line
defining the median; and the whiskers defining the range. Values outside 1.5 times the interquartile distance
are shown. Comparisons of multiple groups were performed using either 1-way or 2-way ANOVA, as
appropriate, followed by Tukey’s or Bonferroni’s post-hoc correction (GraphPad Prism). Two-group anal-
ysis used an unpaired 2-tailed Student’s t test. Statistically significant differences (defined as P ≤ 0.05) and
trends (defined as 0.05 < P < 0.1) between genotypes (WT vs. CH) and respective treatment groups (sham
vs. TAC vs. TAC plus drug treatment) are shown.
Study approval. All mice were cared for and protocols were performed under strict compliance with and
with approved by the Animal Care and Use Committee of the Johns Hopkins University School of Medicine.
Author contributions
RR designed, directed, performed, and analyzed experiments. ET and GZ performed expert TAC and PV
loop analysis. RR and DB acquired and analyzed all echocardiographic images. RR, EGM, RC, VN, PPR,
JGB, EEG, CS, KLM, LM, NAA, and DIL performed experiments and analyzed data. NK provided tis-
sue for analysis. DPJ, DAK, and HCD provided valuable guidance and expertise. RR and HCD wrote the
paper. All authors approved the content and submission of this manuscript.
Acknowledgments
We thank the Johns Hopkins Institute Biomedical Sciences Core Microscope Facility, funded by NIH
grants S10RR024550 and S10OD016374, for assistance in acquisition of immunofluorescence images.
This research was supported by the Sarnoff Cardiovascular Research Foundation (to RR), the William
S. Smilow Foundation for Marfan Syndrome Research (to HCD), and the Howard Hughes Medical
Institute (to HCD).
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
15
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
Address correspondence to: Rosanne Rouf, Division of Cardiology, Department of Medicine, Johns Hop-
kins University School of Medicine, 720 Rutland Avenue, Ross Building 809, Baltimore, Maryland 21224,
USA. Phone: 410.502.2857; Email: rrouf1@jhmi.edu.
Or to: Harry C. Dietz, Institute of Genetic Medicine, Johns Hopkins University School of Medicine, MRB
539, 733 N. Broadway, Baltimore, Maryland 21205, USA. Phone: 410.614.0701; Email: hdietz@jhmi.edu.
RC’s present address is: Department of Medicine, Sinai Hospital of Baltimore, Baltimore Maryland, USA.
PPR’s present address is: Division of Cardiology, Medical University of Graz, Graz, Austria.
NK’s present address is: Department of Cardiovascular Medicine, Gunma University Graduate School of
Medicine, Maebashi, Japan.
1. Judge DP, et al. Evidence for a critical contribution of haploinsufficiency in the complex pathogenesis of Marfan syndrome. J Clin
Invest. 2004;114(2):172–181.
2. Sisk HE, Zahka KG, Pyeritz RE. The Marfan syndrome in early childhood: analysis of 15 patients diagnosed at less than 4
years of age. Am J Cardiol. 1983;52(3):353–358.
3. Morse RP, et al. Diagnosis and management of infantile Marfan syndrome. Pediatrics. 1990;86(6):888–895.
4. Alpendurada F, et al. Evidence for Marfan cardiomyopathy. Eur J Heart Fail. 2010;12(10):1085–1091.
5. Kuwahara F, et al. Transforming growth factor-beta function blocking prevents myocardial fibrosis and diastolic dysfunction in
pressure-overloaded rats. Circulation. 2002;106(1):130–135.
6. Meijboom LJ, et al. Evaluation of left ventricular dimensions and function in Marfan’s syndrome without significant valvular
regurgitation. Am J Cardiol. 2005;95(6):795–797.
7. de Witte P, et al. Intrinsic biventricular dysfunction in Marfan syndrome. Heart. 2011;97(24):2063–2068.
8. Uretsky S, et al. Quantification of left ventricular remodeling in response to isolated aortic or mitral regurgitation. J Cardiovasc
Magn Reson. 2010;12:32.
9. Sakai LY, Keene DR, Engvall E. Fibrillin, a new 350-kD glycoprotein, is a component of extracellular microfibrils. J Cell Biol.
1986;103(6 Pt 1):2499–2509.
10. Neptune ER, et al. Dysregulation of TGF-beta activation contributes to pathogenesis in Marfan syndrome. Nat Genet.
2003;33(3):407–411.
11. Ng CM, et al. TGF-beta-dependent pathogenesis of mitral valve prolapse in a mouse model of Marfan syndrome. J Clin Invest.
2004;114(11):1586–1592.
12. Habashi JP, et al. Losartan, an AT1 antagonist, prevents aortic aneurysm in a mouse model of Marfan syndrome. Science.
2006;312(5770):117–121.
13. Cohn RD, et al. Angiotensin II type 1 receptor blockade attenuates TGF-beta-induced failure of muscle regeneration in multiple
myopathic states. Nat Med. 2007;13(2):204–210.
14. Cook JR, et al. Dimorphic effects of transforming growth factor-β signaling during aortic aneurysm progression in mice suggest
a combinatorial therapy for Marfan syndrome. Arterioscler Thromb Vasc Biol. 2015;35(4):911–917.
15. Eguchi S, Dempsey PJ, Frank GD, Motley ED, Inagami T. Activation of MAPKs by angiotensin II in vascular smooth muscle
cells. Metalloprotease-dependent EGF receptor activation is required for activation of ERK and p38 MAPK but not for JNK.
J Biol Chem. 2001;276(11):7957–7962.
16. Holm TM, et al. Noncanonical TGFβ signaling contributes to aortic aneurysm progression in Marfan syndrome mice. Science.
2011;332(6027):358–361.
17. Khan S, Joyce J, Margulies KB, Tsuda T. Enhanced bioactive myocardial transforming growth factor-β in advanced human
heart failure. Circ J. 2014;78(11):2711–2718.
18. Malebranche L, et al. Enhanced bioactive tgf- levels in the myocardium suggests early pathological transition in asymptomatic
severe mitral regurgitation. J Am Coll Cardiol. 2014;63(12):A922.
19. Fielitz J, et al. Activation of the cardiac renin-angiotensin system and increased myocardial collagen expression in human aortic
valve disease. J Am Coll Cardiol. 2001;37(5):1443–1449.
20. Bouzeghrane F, Reinhardt DP, Reudelhuber TL, Thibault G. Enhanced expression of fibrillin-1, a constituent of the myocardial
extracellular matrix in fibrosis. Am J Physiol Heart Circ Physiol. 2005;289(3):H982–H991.
21. Marijianowski MM, Teeling P, Mann J, Becker AE. Dilated cardiomyopathy is associated with an increase in the type I/type III
collagen ratio: a quantitative assessment. J Am Coll Cardiol. 1995;25(6):1263–1272.
22. Brooks A, Schinde V, Bateman AC, Gallagher PJ. Interstitial fibrosis in the dilated non-ischaemic myocardium. Heart.
2003;89(10):1255–1256.
23. Rosenkranz S. TGF-beta1 and angiotensin networking in cardiac remodeling. Cardiovasc Res. 2004;63(3):423–432.
24. van Wamel AJ, Ruwhof C, van der Valk-Kokshoom LE, Schrier PI, van der Laarse A. The role of angiotensin II, endothelin-1
and transforming growth factor-beta as autocrine/paracrine mediators of stretch-induced cardiomyocyte hypertrophy. Mol Cell
Biochem. 2001;218(1-2):113–124.
25. Boluyt MO, et al. Alterations in cardiac gene expression during the transition from stable hypertrophy to heart failure. Marked
upregulation of genes encoding extracellular matrix components. Circ Res. 1994;75(1):23–32.
26. Parker TG, Packer SE, Schneider MD. Peptide growth factors can provoke “fetal” contractile protein gene expression in rat
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
16
insight.jci.org https://doi.org/10.1172/jci.insight.91588
RESEARCH ARTICLE
cardiac myocytes. J Clin Invest. 1990;85(2):507–514.
27. Villarreal FJ, Dillmann WH. Cardiac hypertrophy-induced changes in mRNA levels for TGF-beta 1, fibronectin, and collagen.
Am J Physiol. 1992;262(6 Pt 2):H1861–H1866.
28. Shimkets RA, et al. Gene expression analysis by transcript profiling coupled to a gene database query. Nat Biotechnol.
1999;17(8):798–803.
29. Zhang H, Hu W, Ramirez F. Developmental expression of fibrillin genes suggests heterogeneity of extracellular microfibrils.
J Cell Biol. 1995;129(4):1165–1176.
30. Vracko R, Thorning D, Frederickson RG. Spatial arrangements of microfibrils in myocardial scars: application of antibody to
fibrillin. J Mol Cell Cardiol. 1990;22(7):749–757.
31. Tae HJ, Petrashevskaya N, Marshall S, Krawczyk M, Talan M. Cardiac remodeling in the mouse model of Marfan syndrome
develops into two distinctive phenotypes. Am J Physiol Heart Circ Physiol. 2016;310(2):H290–H299.
32. Cook JR, et al. Abnormal muscle mechanosignaling triggers cardiomyopathy in mice with Marfan syndrome. J Clin Invest.
2014;124(3):1329–1339.
33. Pacher P, Nagayama T, Mukhopadhyay P, Bátkai S, Kass DA. Measurement of cardiac function using pressure-volume conduc-
tance catheter technique in mice and rats. Nat Protoc. 2008;3(9):1422–1434.
34. Kuwahara K, et al. NRSF regulates the fetal cardiac gene program and maintains normal cardiac structure and function. EMBO
J. 2003;22(23):6310–6321.
35. Moore-Morris T, et al. Resident fibroblast lineages mediate pressure overload-induced cardiac fibrosis. J Clin Invest.
2014;124(7):2921–2934.
36. Muchir A, Shan J, Bonne G, Lehnart SE, Worman HJ. Inhibition of extracellular signal-regulated kinase signaling to prevent
cardiomyopathy caused by mutation in the gene encoding A-type lamins. Hum Mol Genet. 2009;18(2):241–247.
37. Hung KY, Chen CT, Huang JW, Lee PH, Tsai TJ, Hsieh BS. Dipyridamole inhibits TGF-beta-induced collagen gene expression
in human peritoneal mesothelial cells. Kidney Int. 2001;60(4):1249–1257.
38. Iverson C, et al. RDEA119/BAY 869766: a potent, selective, allosteric inhibitor of MEK1/2 for the treatment of cancer. Cancer
Res. 2009;69(17):6839–6847.
39. Zahradka P. Novel role for osteopontin in cardiac fibrosis. Circ Res. 2008;102(3):270–272.
40. Ashizawa N, et al. Osteopontin is produced by rat cardiac fibroblasts and mediates A(II)-induced DNA synthesis and collagen
gel contraction. J Clin Invest. 1996;98(10):2218–2227.
41. Wang D, et al. Effects of pressure overload on extracellular matrix expression in the heart of the atrial natriuretic peptide-null
mouse. Hypertension. 2003;42(1):88–95.
42. Koitabashi N, et al. Pivotal role of cardiomyocyte TGF-β signaling in the murine pathological response to sustained pressure
overload. J Clin Invest. 2011;121(6):2301–2312.
43. Baum J, Duffy HS. Fibroblasts and myofibroblasts: what are we talking about? J Cardiovasc Pharmacol. 2011;57(4):376–379.
44. Booms P, et al. Novel exon skipping mutation in the fibrillin-1 gene: two ‘hot spots’ for the neonatal Marfan syndrome. Clin
Genet. 1999;55(2):110–117.
45. Geva T, Sanders SP, Diogenes MS, Rockenmacher S, Van Praagh R. Two-dimensional and Doppler echocardiographic and
pathologic characteristics of the infantile Marfan syndrome. Am J Cardiol. 1990;65(18):1230–1237.
46. Brito-Filho SL, Oporto V, Campos O, Alvares AB, Carvalho AC. A case of neonatal Marfan syndrome with good late fol-
low-up: is it possible to avoid an early unfavourable outcome? Cardiol Young. 2013;23(2):301–303.
47. Kitahara H, Aeba R, Takaki H, Shimizu H. Palliative mitral valve repair during infancy for neonatal Marfan syndrome. Ann
Thorac Surg. 2016;101(5):1987–1988.
48. Seo YJ, Lee KE, Kim GB, Kwon BS, Bae EJ, Noh CI. Infantile Marfan syndrome in a Korean tertiary referral center. Korean J
Pediatr. 2016;59(2):59–64.
49. Buerger A, et al. Dilated cardiomyopathy resulting from high-level myocardial expression of Cre-recombinase. J Card Fail.
2006;12(5):392–398.
50. Davis J, Maillet M, Miano JM, Molkentin JD. Lost in transgenesis: a user’s guide for genetically manipulating the mouse in
cardiac research. Circ Res. 2012;111(6):761–777.
51. Habashi JP, et al. Angiotensin II type 2 receptor signaling attenuates aortic aneurysm in mice through ERK antagonism. Science.
2011;332(6027):361–365.
52. Teekakirikul P, et al. Cardiac fibrosis in mice with hypertrophic cardiomyopathy is mediated by non-myocyte proliferation and
requires Tgf-β. J Clin Invest. 2010;120(10):3520–3529.
53. Takimoto E, et al. Chronic inhibition of cyclic GMP phosphodiesterase 5A prevents and reverses cardiac hypertrophy. Nat Med.
2005;11(2):214–222.
54. Krajewska M, et al. Image analysis algorithms for immunohistochemical assessment of cell death events and fibrosis in tissue
sections. J Histochem Cytochem. 2009;57(7):649–663.
55. Seo K, et al. Combined TRPC3 and TRPC6 blockade by selective small-molecule or genetic deletion inhibits pathological cardi-
ac hypertrophy. Proc Natl Acad Sci U S A. 2014;111(4):1551–1556.
56. Baschong W, Suetterlin R, Laeng RH. Control of autofluorescence of archival formaldehyde-fixed, paraffin-embedded tissue in
confocal laser scanning microscopy (CLSM). J Histochem Cytochem. 2001;49(12):1565–1572.
57. Komuro I, et al. Stretching cardiac myocytes stimulates protooncogene expression. J Biol Chem. 1990;265(7):3595–3598.
Downloaded from http://insight.jci.org on March 8, 2018. https://doi.org/10.1172/jci.insight.91588
... Moreover, MFS mice that are exposed to pressure overload develop an acute severe Erk1/2-Tgfb mediated cardiomyopathy that can be rescued by Losartan. (Rouf et al., 2017). In our small cohort of MFS patients, we see systolic and diastolic cardiac dysfunction as well as increased indexed LV volume and mass, which is in line with prior reports. ...
... In addition, BNP correlates strongly with both cfPWV (r = 0.68, p = 0.004) and CSP (r = 0.57, p = 0.023)-equivalent to pressure overload in the mouse model. (Rouf et al., 2017). As nonmyocytes play a critical role in pathogenesis of MFS cardiomyopathy, we suggest that a decompensated Frank-Starling mechanism due to LV dilation may trigger then LV dysfunction. ...
... As nonmyocytes play a critical role in pathogenesis of MFS cardiomyopathy, we suggest that a decompensated Frank-Starling mechanism due to LV dilation may trigger then LV dysfunction. (Rouf et al., 2017). This may be why we see signs of structural remodeling and increasing BNP in correlation with increased arterial load first. ...
Article
Full-text available
Background: Marfan Syndrome (MFS) has been associated with increased aortic stiffness and left ventricular dysfunction. The latter may be due to the underlying genotype and/or secondary to aortic stiffening (vascular-ventricular interaction). The aim of this study was to characterize arterial and cardiac function in MFS using a multimodal approach. Methods: Prospective observational study of MFS patients and healthy controls. Methods included echocardiography, ascending aortic distensibility, common carotid intima media thickness [cIMT], parameters of wave reflection, carotid-femoral pulse wave velocity [cfPWV]), reactive hyperemia index [RHI], and biomarker analysis (Olink, CVII panel). Results: We included 20 patients with MFS and 67 controls. Ascending aortic distensibility, cIMT and RHI were decreased, while all parameters of arterial wave reflection, stiffness and BNP levels were increased in the MFS group. Both systolic and diastolic function were impaired relative to controls. Within the MFS group, no significant correlation between arterial and cardiac function was identified. However, cfPWV correlated significantly with indexed left ventricular mass and volume in MFS. Bran natriuretic peptide (BNP) was the only biomarker significantly elevated in MFS following correction for age and sex. Conclusions: MFS patients have generally increased aortic stiffness, endothelial dysfunction and BNP levels while cIMT is decreased, supporting that the mechanism of general stiffening is different from acquired vascular disease. CfPWV is associated with cardiac size, blood pressure and BNP in MFS patients. These may be early markers of disease progression that are suitable for monitoring pharmacological treatment effects in MFS patients.
... [130][131][132][133] There are several published examples illuminating the significance of ISH in establishing disease pathogenesis and we highlight a few here. [134][135][136][137][138][139][140][141][142][143] For example, FISH was used to localize Ifnar1 and Ifnar2 in mouse dorsal root ganglia (DRG) sensory neurons supporting the idea that activation of these receptors could modulate nociceptive signaling during viral infection. 144 In addition, downregulated expression of neuronal genes, such as TRPV1, SLC17A7, and PRDM12 was demonstrated in L4 DRGs of patients further providing new insights into intraganglionic pathology and mechanisms causing neuropathic pain in patients with type 2 diabetes mellitus. ...
Article
In situ hybridization (ISH) is used for the localization of specific nucleic acid sequences in cells or tissues by complementary binding of a nucleotide probe to a specific target nucleic acid sequence. In the last years, the specificity and sensitivity of ISH assays were improved by innovative techniques like synthetic nucleic acids and tandem oligonucleotide probes combined with signal amplification methods like branched DNA, hybridization chain reaction and tyramide signal amplification. These improvements increased the application spectrum for ISH on formalin-fixed paraffin-embedded tissues. ISH is a powerful tool to investigate DNA, mRNA transcripts, regulatory noncoding RNA, and therapeutic oligonucleotides. ISH can be used to obtain spatial information of a cell type, subcellular localization, or expression levels of targets. Since immunohistochemistry and ISH share similar workflows, their combination can address simultaneous transcriptomics and proteomics questions. The goal of this review paper is to revisit the current state of the scientific approaches in ISH and its application in drug research and development.
... Extracellular-regulated protein kinases (ERK) can also get activated independently of Smad through small GTPase Ras under growth factor stimulation. In a study on Marfan mice ERK, it was found to induce cardiac fibrosis (Rouf et al., 2017). MAPK and ERK can also activate the Smad-signaling pathway. ...
Article
Full-text available
Extracellular matrix (ECM) plays a critical role in maintaining elasticity in cardiac tissues. Elasticity is required in the heart for properly pumping blood to the whole body. Dysregulated ECM remodeling causes fibrosis in the cardiac tissues. Cardiac fibrosis leads to stiffness in the heart tissues, resulting in heart failure. During cardiac fibrosis, ECM proteins get excessively deposited in the cardiac tissues. In the ECM, cardiac fibroblast proliferates into myofibroblast upon various kinds of stimulations. Fibroblast activation (myofibroblast) contributes majorly toward cardiac fibrosis. Other than cardiac fibroblasts, cardiomyocytes, epithelial/endothelial cells, and immune system cells can also contribute to cardiac fibrosis. Alteration in the expression of the ECM core and ECM-modifier proteins causes different types of cardiac fibrosis. These different components of ECM culminated into different pathways inducing transdifferentiation of cardiac fibroblast into myofibroblast. In this review, we summarize the role of different ECM components during cardiac fibrosis progression leading to heart failure. Furthermore, we highlight the importance of applying mass-spectrometry-based proteomics to understand the key changes occurring in the ECM during fibrotic progression. Next-gen proteomics studies will broaden the potential to identify key targets to combat cardiac fibrosis in order to achieve precise medicine-development in the future.
... Images were acquired with an Olympus DP72 camera on an Olympus BX43 microscope using cellSens Standard version 1.4.1 (Olympus). Quantification of myocyte cross-sectional area was performed as previously described (66). In short, following deparaffinizing and citrate-based heatmediated antigen retrieval, slides were incubated with Alexa Fluor 647-conjugated wheat germ agglutinin (W32466, Invitrogen) at 4°C overnight and mounted using Prolong Gold mounting medium (P36934, Thermofisher). ...
Preprint
Full-text available
Hypereosinophilic syndrome is a progressive disease with extensive eosinophilia that results in organ damage. Cardiac pathologies are the main reason for its high mortality rate. A better understanding of the mechanisms of eosinophil-mediated tissue damage would benefit therapeutic development. Here, we describe the cardiac pathologies that developed in a mouse model of hypereosinophilic syndrome. These IL-5 transgenic mice exhibited decreased left ventricular function at a young age which worsened with age. Mechanistically, we demonstrated infiltration of activated eosinophils into the heart tissue that led to an inflammatory environment. Gene expression signatures showed tissue damage as well as repair and remodeling processes. Cardiomyocytes from IL-5Tg mice exhibited significantly reduced contractility relative to WT controls. This impairment may result from the inflammatory stress experienced by the cardiomyocytes and suggest that dysregulation of contractility and Ca ²⁺ reuptake in cardiomyocytes contributes to cardiac dysfunction at the whole organ level in hypereosinophilic mice. Teaser Too many eosinophils cause inflammation in the heart and change cardiomyocyte contraction leading to poor heart function.
... This aortic remodeling occurred in absence of cardiac remodeling, as expected (Table S1 in the Supplemental Material). 21 Abdominal aortas (Abd Ao) of mMFS mice had milder dilatation in the absence of a decrease in aortic velocity ( Figure S2A through S2D in the Supplemental Material). ...
Article
Background Marfan syndrome (MFS) is associated with TGF (transforming growth factor) β–stimulated ERK (extracellular signal-regulated kinase) activity in vascular smooth muscle cells (VSMCs), which adopt a mixed synthetic/contractile phenotype. In VSMCs, TGFβ induces IL (interleukin) 11) that stimulates ERK-dependent secretion of collagens and MMPs (matrix metalloproteinases). Here, we examined the role of IL11 in the MFS aorta. Methods We used echocardiography, histology, immunostaining, and biochemical methods to study aortic anatomy, physiology, and molecular endophenotypes in Fbn1 C1041G/+ mice, an established murine model of MFS (mMFS). mMFS mice were crossed to an IL11-tagged EGFP (enhanced green fluorescent protein; Il11 EGFP/+ ) reporter strain or to a strain deleted for the IL11 receptor ( Il11ra1 −/− ). In therapeutic studies, mMFS were administered an X209 (neutralizing antibody against IL11RA [IL11 receptor subunit alpha]) or IgG for 20 weeks and imaged longitudinally. Results IL11 mRNA and protein were elevated in the aortas of mMFS mice, as compared to controls. mMFS mice crossed to Il11 EGFP/+ mice had increased IL11 expression in VSMCs, notably in the aortic root and ascending aorta. As compared to the mMFS parental strain, double mutant mMFS: Il11ra1 −/ ⁻ mice had reduced aortic dilatation and exhibited lesser fibrosis, inflammation, elastin breaks, and VSMC loss, which was associated with reduced aortic COL1A1 (collagen type I alpha 1 chain), IL11, MMP2/9, and phospho-ERK expression. To explore therapeutic targeting of IL11 signaling in MFS, we administered either a neutralizing antibody against IL11RA (X209) or an IgG control. After 20 weeks of antibody administration, as compared to IgG, mMFS mice receiving X209 had reduced thoracic and abdominal aortic dilation as well as lesser fibrosis, inflammation, elastin breaks, and VSMC loss. By immunoblotting, X209 was shown to reduce aortic COL1A1, IL11, MMP2/9, and phospho-ERK expression. Conclusions In MFS, IL11 is upregulated in aortic VSMCs to cause ERK-related thoracic aortic dilatation, inflammation, and fibrosis. Therapeutic inhibition of IL11, imminent in clinical trials, might be considered as a new approach in MFS.
... This phenomenon may be associated with intrinsic cardiac impairment in MFS. In addition to case reports on heart transplantation, some researchers have verified the existence of intrinsic cardiac impairment in MFS through animal experiments.27,28 Jason et al.29 proved that partial FBN1 gene inactivation in cardiomyocytes was sufficient to lead to the cardiac dysfunction in mice. ...
Article
Full-text available
Background Intrinsic cardiac impairment in Marfan syndrome (MFS) has been explored in many clinical studies; however, their results have been inconsistent. This meta-analysis aimed to assess the difference in cardiac structure and function between Marfan patients and healthy individuals, and to verify the hypothesis of intrinsic cardiac impairment in MFS. Methods Electronic searches for studies were performed in the PubMed, Embase, and Cochrane Library databases. Nine studies with 490 patients with MFS and 478 controls were included in the analysis. Age and sex were strictly matched between Marfan patients and healthy controls in every study. Results There was no difference in the left ventricular end systolic diameter index (mean difference [MD]: 0.33; 95% confidence interval [CI]: (−0.24, 0.89); p = 0.26) and left ventricular end diastolic diameter index (MD: 0.18; 95% CI: [−0.47, 0.83]; p = 0.58) between Marfan patients and controls. Marfan patients showed larger left ventricular end systolic volume index (MD: 2.62; 95% CI: [0.27, 4.97]; p = 0.03) and left ventricular end diastolic volume index (MD: 4.16; 95% CI: [2.70, 5.63]; p < 0.01) than the control group. Furthermore, Marfan patients showed a lower left ventricular ejection fraction than healthy people (MD: −2.59%; 95% CI: [−4.64%, −0.54%]; p = 0.01). Conclusions Intrinsic cardiac impairment was observed in MFS. MFS patients showed the larger left ventricular volume and poorer left ventricular function than matched controls. Considering the potentially adverse impact on cardiac function, intrinsic cardiac impairment in MFS should be considered during the cardiac surgery.
... Additionally, focal adhesion signalling in SMCs is predicted to be disrupted by genetic alterations affecting components of the elastin-contractile unit, and focal adhesion signalling is the most altered pathway in SMCs in Fbn1 C1039G/+ mice when compared with wild-type mice, on the basis of single-cell RNA sequencing of the aorta 87 . Thus, mechanosignalling is emerging as having a role in aortic disease, and studies in MFS mouse models have also suggested that the aetiology of the cardiomyopathy is altered mechanosignalling owing to the underlying fibrillin-1 defect [88][89][90] . ...
Article
Marfan syndrome (MFS) is an autosomal dominant, age-related but highly penetrant condition with substantial intrafamilial and interfamilial variability. MFS is caused by pathogenetic variants in FBN1, which encodes fibrillin-1, a major structural component of the extracellular matrix that provides support to connective tissues, particularly in arteries, the pericondrium and structures in the eye. Up to 25% of individuals with MFS have de novo variants. The most prominent manifestations of MFS are asymptomatic aortic root aneurysms, aortic dissections, dislocation of the ocular lens (ectopia lentis) and skeletal abnormalities that are characterized by overgrowth of the long bones. MFS is diagnosed based on the Ghent II nosology; genetic testing confirming the presence of a FBN1 pathogenetic variant is not always required for diagnosis but can help distinguish MFS from other heritable thoracic aortic disease syndromes that can present with skeletal features similar to those in MFS. Untreated aortic root aneurysms can progress to life-threatening acute aortic dissections. Management of MFS requires medical therapy to slow the rate of growth of aneurysms and decrease the risk of dissection. Routine surveillance with imaging techniques such as transthoracic echocardiography, CT or MRI is necessary to monitor aneurysm growth and determine when to perform prophylactic repair surgery to prevent an acute aortic dissection.
Article
Full-text available
Hypereosinophilic syndrome is a progressive disease with extensive eosinophilia that results in organ damage. Cardiac pathologies are the main reason for its high mortality rate. A better understanding of the mechanisms of eosinophil-mediated tissue damage would benefit therapeutic development. Here, we describe the cardiac pathologies that developed in a mouse model of hypereosinophilic syndrome. These IL-5 transgenic mice exhibited decreased left ventricular function at a young age which worsened with age. Mechanistically, we demonstrated infiltration of activated eosinophils into the heart tissue that led to an inflammatory environment. Gene expression signatures showed tissue damage as well as repair and remodeling processes. Cardiomyocytes from IL-5Tg mice exhibited significantly reduced contractility relative to wild type (WT) controls. This impairment may result from the inflammatory stress experienced by the cardiomyocytes and suggest that dysregulation of contractility and Ca²⁺ reuptake in cardiomyocytes contributes to cardiac dysfunction at the whole organ level in hypereosinophilic mice.
Article
BACKGROUND Organ fibrosis due to excessive production of extracellular matrix by resident fibroblasts is estimated to contribute to >45% of deaths in the Western world, including those due to cardiovascular diseases such as heart failure. Here, we screened for small molecule inhibitors with a common ability to suppress activation of fibroblasts across organ systems. METHODS High-content imaging of cultured cardiac, pulmonary, and renal fibroblasts was used to identify nontoxic compounds that blocked induction of markers of activation in response to the profibrotic stimulus, transforming growth factor-β1. SW033291, which inhibits the eicosanoid-degrading enzyme, 15-hydroxyprostaglandin dehydrogenase, was chosen for follow-up studies with cultured adult rat ventricular fibroblasts and human cardiac fibroblasts (CFs), and for evaluation in mouse models of cardiac fibrosis and diastolic dysfunction. Additional mechanistic studies were performed with CFs treated with exogenous eicosanoids. RESULTS Nine compounds, including SW033291, shared a common ability to suppress transforming growth factor-β1–mediated activation of cardiac, pulmonary, and renal fibroblasts. SW033291 dose-dependently inhibited transforming growth factor-β1–induced expression of activation markers (eg, α-smooth muscle actin and periostin) in adult rat ventricular fibroblasts and normal human CFs, and reduced contractile capacity of the cells. Remarkably, the 15-hydroxyprostaglandin dehydrogenase inhibitor also reversed constitutive activation of fibroblasts obtained from explanted hearts from patients with heart failure. SW033291 blocked cardiac fibrosis induced by angiotensin II infusion and ameliorated diastolic dysfunction in an alternative model of systemic hypertension driven by combined uninephrectomy and deoxycorticosterone acetate administration. Mechanistically, SW033291-mediated stimulation of extracellular signal-regulated kinase 1/2 mitogen-activated protein kinase signaling was required for SW033291 to block CF activation. Of the 12 exogenous eicosanoids that were tested, only 12-hydroxyeicosatetraenoic acid, which signals through the G protein-coupled receptor, GPR31, recapitulated the suppressive effects of SW033291 on CF activation. CONCLUSIONS Inhibition of degradation of eicosanoids, arachidonic acid-derived fatty acids that signal through G protein-coupled receptors, is a potential therapeutic strategy for suppression of pathological organ fibrosis. In the heart, we propose that 15-hydroxyprostaglandin dehydrogenase inhibition triggers CF-derived autocrine/paracrine signaling by eicosanoids, including 12-hydroxyeicosatetraenoic acid, to stimulate extracellular signal-regulated kinase 1/2 and block conversion of fibroblasts into activated cells that secrete excessive amounts of extracellular matrix and contribute of heart failure pathogenesis.
Article
Full-text available
Aortic root aneurysm is a common cause of morbidity and mortality in Loeys-Dietz and Marfan syndromes, where perturbations in transforming growth factor beta (TGFβ) signaling play a causal or contributory role, respectively. Despite the advantages of cross-species disease modeling, animal models of aortic root aneurysm are largely restricted to genetically engineered mice. Here, we report that zebrafish devoid of the genes encoding latent-transforming growth factor beta-binding protein 1 and 3 (ltbp1 and ltbp3, respectively) develop rapid and severe aneurysm of the outflow tract (OFT), the aortic root equivalent. Similar to syndromic aneurysm tissue, the distended OFTs display evidence for paradoxical hyperactivated TGFβ signaling. RNA-sequencing revealed significant overlap between the molecular signatures of disease tissue from mutant zebrafish and a mouse model of Marfan syndrome. Moreover, chemical inhibition of TGFβ signaling in wild-type animals phenocopied mutants but chemical activation did not, demonstrating that TGFβ signaling is protective against aneurysm. Human relevance is supported by recent studies implicating genetic lesions in LTBP3 and, potentially, LTBP1 as heritable causes of aortic root aneurysm. Ultimately, our data demonstrate that zebrafish can now be leveraged to interrogate thoracic aneurysmal disease and identify novel lead compounds through small-molecule suppressor screens. This article has an associated First Person interview with the first author of the paper.
Article
Full-text available
Purpose: Infantile Marfan syndrome (MFS) is a rare congenital inheritable connective tissue disorder with poor prognosis. This study aimed to evaluate the cardiovascular manifestations and overall prognosis of infantile MFS diagnosed in a tertiary referral center in Korea. Methods: Eight patients diagnosed with infantile MFS between 2004 and 2014 were retrospectively evaluated. Results: Their median age at the time of diagnosis was 2.5 months (range, 0-20 months). The median follow-up period was 25.5 months (range, 0-94 months). The median length at birth was 50.0 cm (range, 48-53 cm); however, height became more prominent over time, and the patients were taller than the 97th percentile at the time of the study. None of the patients had any relevant family history. Four of the 5 patients who underwent DNA sequencing had a fibrillin 1 gene mutation. All the patients with echocardiographic data of the aortic root had a z score of >2. All had mitral and tricuspid valve prolapse, and various degrees of mitral and tricuspid regurgitation. Five patients underwent open-heart surgery, including mitral valve replacement, of whom two required multiple operations. The median age at mitral valve replacement was 28.5 months (range, 5-69 months). Seven patients showed congestive heart failure before surgery or during follow-up, and required multiple anti-heart failure medications. Four patients died of heart failure at a median age of 12 months. Conclusion: The prognosis of infantile MFS is poor; thus, early diagnosis and timely cautious treatment are essential to prevent further morbidity and mortality.
Article
Full-text available
Background: Transforming growth factor (TGF)-β activation is known to play a central role in progressive ventricular remodeling in advanced heart failure in animal models, but there has been no direct evidence of increased TGF-β activity in the myocardium of patients with advanced human heart failure. METHODS AND RESULTS: Using a recently developed bioassay that measures TGF-β bioactivity rather than TGF-β abundance, we measured bioactive TGF-β in human myocardium from control non-failing donors (NF), and patients with ischemic cardiomyopathy (ICM) and dilated cardiomyopathy (DCM). Both free and total soluble TGF-β were significantly increased in ICM and DCM compared with NF. Free TGF-β had an excellent correlation with phosphorylated Smad2 (R(2)=0.55, P<0.0001), a downstream marker of TGF-β signaling. Collagen type I and type III were significantly upregulated in DCM compared with NF, consistent with histological evidence of myocardial fibrosis. Expression of fibulin-2, a positive modulator of TGF-β, was significantly increased in DCM compared with NF, and the free TGF-β level was correlated with fibulin-2 mRNA (R(2)=0.24, P<0.006). Conclusions: Although both free and total soluble TGF-β are significantly increased in ICM and DCM compared with NF, the superior correlation of free TGF-β with downstream signaling suggests that this is the most functionally relevant form. The present findings suggest that sustained TGF-β activation in both ICM and DCM contributes to excess myocardial fibrosis.
Article
Full-text available
Activation and accumulation of cardiac fibroblasts, which result in excessive extracellular matrix deposition and consequent mechanical stiffness, myocyte uncoupling, and ischemia, are key contributors to heart failure progression. Recently, endothelial-to-mesenchymal transition (EndoMT) and the recruitment of circulating hematopoietic progenitors to the heart have been reported to generate substantial numbers of cardiac fibroblasts in response to pressure overload-induced injury; therefore, these processes are widely considered to be promising therapeutic targets. Here, using multiple independent murine Cre lines and a collagen1a1-GFP fusion reporter, which specifically labels fibroblasts, we found that following pressure overload, fibroblasts were not derived from hematopoietic cells, EndoMT, or epicardial epithelial-to-mesenchymal transition. Instead, pressure overload promoted comparable proliferation and activation of two resident fibroblast lineages, including a previously described epicardial population and a population of endothelial origin. Together, these data present a paradigm for the origins of cardiac fibroblasts during development and in fibrosis. Furthermore, these data indicate that therapeutic strategies for reducing pathogenic cardiac fibroblasts should shift from targeting presumptive EndoMT or infiltrating hematopoietically derived fibroblasts, toward common pathways upregulated in two endogenous fibroblast populations.
Article
Full-text available
Patients with Marfan syndrome (MFS), a multisystem disorder caused by mutations in the gene encoding the extracellular matrix (ECM) protein fibrillin 1, are unusually vulnerable to stress-induced cardiac dysfunction. The prevailing view is that MFS-associated cardiac dysfunction is the result of aortic and/or valvular disease. Here, we determined that dilated cardiomyopathy (DCM) in fibrillin 1-deficient mice is a primary manifestation resulting from ECM-induced abnormal mechanosignaling by cardiomyocytes. MFS mice displayed spontaneous emergence of an enlarged and dysfunctional heart, altered physical properties of myocardial tissue, and biochemical evidence of chronic mechanical stress, including increased angiotensin II type I receptor (AT1R) signaling and abated focal adhesion kinase (FAK) activity. Partial fibrillin 1 gene inactivation in cardiomyocytes was sufficient to precipitate DCM in otherwise phenotypically normal mice. Consistent with abnormal mechanosignaling, normal cardiac size and function were restored in MFS mice treated with an AT1R antagonist and in MFS mice lacking AT1R or β-arrestin 2, but not in MFS mice treated with an angiotensin-converting enzyme inhibitor or lacking angiotensinogen. Conversely, DCM associated with abnormal AT1R and FAK signaling was the sole abnormality in mice that were haploinsufficient for both fibrillin 1 and β1 integrin. Collectively, these findings implicate fibrillin 1 in the physiological adaptation of cardiac muscle to elevated workload.
Article
An infant with neonatal Marfan syndrome (nMFS), a condition that is nearly always lethal during infancy, was referred to our hospital with symptoms of congestive heart failure resulting from severe mitral valve insufficiency. During mitral valve repair, the use of an annuloplasty ring was waived until annular dilatation was achieved after 2 palliative mitral valvuloplasty procedures. After the definitive operation, the patient's mitral valve function remained within normal limits until the last follow-up when the patient was 11 years old. To the best of our knowledge, this patient has the longest recorded survival after mitral valve repair.
Article
Mitral valve prolapse (MVP) is a common human phenotype, yet little is known about the pathogenesis of this condition. MVP can occur in the context of genetic syndromes, including Marfan syndrome (MFS), an autosomal-dominant connective tissue disorder caused by mutations in fibrillin-1. Fibrillin-1 contributes to the regulated activation of the cytokine TGF-β, and enhanced signaling is a consequence of fibrillin-1 deficiency. We thus hypothesized that increased TGF-β signaling may contribute to the multisystem pathogenesis of MFS, including the development of myxomatous changes of the atrioventricular valves. Mitral valves from fibrillin-1–deficient mice exhibited postnatally acquired alterations in architecture that correlated both temporally and spatially with increased cell proliferation, decreased apoptosis, and excess TGF-β activation and signaling. In addition, TGF-β antagonism in vivo rescued the valve phenotype, suggesting a cause and effect relationship. Expression analyses identified increased expression of numerous TGF-β–related genes that regulate cell proliferation and survival and plausibly contribute to myxomatous valve disease. These studies validate a novel, genetically engineered murine model of myxomatous changes of the mitral valve and provide critical insight into the pathogenetic mechanism of such changes in MFS and perhaps more common nonsyndromic variants of mitral valve disease. Histologic and morphometric assessment of mitral valve architecture in P6.5 mice. Fbn1 genotypes are indicated as follows: +/+ (Fbn1+/+), +/− (Fbn1C1039G/+), and −/− (Fbn1C1039G/C1039G). (A) Representative mitral valve sections from each genotype at P6.5 showing increased length and thickness in mutant valves as compared to wild-type littermates. Magnification, ×20. Scale bars: 100 μm. (B) Morphometric analysis of mitral valve length during the first week of postnatal life. Fbn1C1039G/C1039G valves were significantly longer by P6.5 when compared with those of Fbn1+/+ animals (–;P –; 0.05). (C) Morphometric analysis of mitral valve thickness during the first week of postnatal life. Increased valve thickness in Fbn1C1039G/C1039G versus Fbn1+/+ mice was significant by P4.5 (*P –; 0.005 vs. Fbn1+/+), and by P6.5, differences between all genotypes were statistically significant (*P –; 0.005 vs. Fbn1+/+; –;P –; 0.05 vs. Fbn1C1039G/+). Error bars indicate 95% confidence intervals. (D) Echocardiographic parasternal long axis views of 9-month Fbn1C1039G/+ and Fbn1+/+ mouse hearts. The anterior mitral valve leaflet (arrowhead) shows increased length and thickening, as well as systolic prolapse into the LA in fibrillin-1–;deficient mice. AoR, aortic root; AscAo, ascending aorta; LA, left atrium.
Article
Marfan syndrome (MFS) is a systemic disorder of connective tissue caused by mutations in fibrillin-1. Cardiac dysfunction in MFS has not been characterized halting the development of therapies of cardiac complications in Marfan syndrome. We aimed to study the age-dependent cardiac remodeling in the mouse model of MFS Fbn1039G+/- mouse (Marfan HT mouse) and its association with valvular regurgitation. Marfan HT mice of 2-4 months demonstrated a mild hypertrophic cardiac remodeling with predominant decline of diastolic function and increased TGF-β canonical (p SMAD2/3) and non-canonical (p ERK 1/2 and p p38 MAPK) signaling and up regulation of hypertrophic markers natriuretic peptides atrium natriuretic peptide (ANP) and brain natriuretic peptide (BNP). Among older HT mice (6-14 months) cardiac remodeling was associated with two distinct phenotypes manifesting either dilated or constricted LV chamber. Dilatation of LV chamber was accompanied by biochemical evidence of greater mechanical stress, including elevated ERK1/2 and pMAPK38 phosphorylation and higher BNP expression. The aortic valve regurgitation was registered in 20% of constricted group and 60% of dilated, while mitral insufficiency was observed in 40% of constricted group and 100% of dilated group. Cardiac dysfunction was not associated with the increase of interstitial fibrosis and non-myocyte proliferation. In the mouse model fibrillin-1 haploinsufficiency results in the early onset of non-fibrotic hypertrophic cardiac remodeling and dysfunction independently from valvular abnormalities. MFS heart is vulnerable to stress-induced cardiac dilatation in the face of valvular regurgitation and stress-activated MAPK signals represent a potential target for cardiac management in MFS .
Article
Extracellular microfibrils, alone or in association with elastin, confer critical biomechanical properties on a variety of connective tissues. Little is known about the composition of the microfibrils or the factors responsible for their spatial organization into tissue-specific macroaggregates. Recent work has revealed the existence of two structurally related microfibrillar components, termed fibrillin-1 and fibrillin-2. The functional relationships between these glycoproteins and between them and other components of the microfibrils and elastic fibers are obscure. As a first step toward elucidating these important points, we compared the expression pattern of the fibrillin genes during mammalian embryogenesis. The results revealed that the two genes are differentially expressed, in terms of both developmental stages and tissue distribution. In the majority of cases, fibrillin-2 transcripts appear earlier and accumulate for a shorter period of time than fibrillin-1 transcripts. Synthesis of fibrillin-1 correlates with late morphogenesis and the appearance of well-defined organ structures; fibrillin-2 synthesis, on the other hand, coincides with early morphogenesis and, in particular, with the beginning of elastogenesis. The findings lend indirect support to our original hypothesis stating that fibrillins contribute to the compositional and functional heterogeneity of the microfibrils. The available evidence is also consistent with the notion that the fibrillins might have distinct, but related roles in microfibril physiology. Accordingly, we propose that fibrillin-1 provides mostly force-bearing structural support, whereas fibrillin-2 predominantly regulates the early process of elastic fiber assembly.
Article
OBJECTIVE: Studies of mice with mild Marfan syndrome (MFS) have correlated the development of thoracic aortic aneurysm (TAA) with improper stimulation of noncanonical (Erk-mediated) TGFβ signaling by the angiotensin type I receptor (AT1r). This correlation was largely based on comparable TAA modifications by either systemic TGFβ neutralization or AT1r antagonism. However, subsequent investigations have called into question some key aspects of this mechanism of arterial disease in MFS. To resolve these controversial points, here we made a head-to-head comparison of the therapeutic benefits of TGFβ neutralization and AT1r antagonism in mice with progressively severe MFS (Fbn1mgR/mgR mice). APPROACH AND RESULTS: Aneurysm growth, media degeneration, aortic levels of phosphorylated Erk and Smad proteins and the average survival of Fbn1mgR/mgR mice were compared after a ≈3-month-long treatment with placebo and either the AT1r antagonist losartan or the TGFβ-neutralizing antibody 1D11. In contrast to the beneficial effect of losartan, TGFβ neutralization either exacerbated or mitigated TAA formation depending on whether treatment was initiated before (postnatal day 16; P16) or after (P45) aneurysm formation, respectively. Biochemical evidence-related aneurysm growth with Erk-mediated AT1r signaling, and medial degeneration with TGFβ hyperactivity that was in part AT1r dependent. Importantly, P16-initiated treatment with losartan combined with P45-initiated administration of 1D11 prevented death of Fbn1mgR/mgR mice from ruptured TAA. CONCLUSIONS: By demonstrating that promiscuous AT1r and TGFβ drive partially overlapping processes of arterial disease in MFS mice, our study argues for a therapeutic strategy against TAA that targets both signaling pathways although sparing the early protective role of TGFβ.