Вы находитесь на странице: 1из 37

Associate Editor: P.

Molenaar
Molecular distinction between physiological and pathological cardiac hypertrophy:
Experimental ndings and therapeutic strategies
Bianca C. Bernardo
a
, Kate L. Weeks
a,b
, Lynette Pretorius
a,c
, Julie R. McMullen
a,

a
Cardiac Hypertrophy Laboratory, Baker IDI Heart & Diabetes Institute, Melbourne, Australia
b
Faculty of Medicine, Dentistry and Health Sciences, Department of Biochemistry and Molecular Biology, University of Melbourne, Melbourne, Australia
c
Faculty of Medicine, Nursing and Health Sciences, Department of Medicine (Alfred Hospital), Monash University, Melbourne, Australia
a b s t r a c t a r t i c l e i n f o
Keywords:
Heart failure
Phosphoinositide 3-kinase
Insulin-like growth factor 1
Physiological cardiac hypertrophy
Pathological cardiac hypertrophy
Gender differences
Therapeutic applications
Cardiac hypertrophy can be dened as an increase in heart mass. Pathological cardiac hypertrophy (heart
growth that occurs in settings of disease, e.g. hypertension) is a key risk factor for heart failure. Pathological
hypertrophy is associated with increased interstitial brosis, cell death and cardiac dysfunction. In contrast,
physiological cardiac hypertrophy (heart growth that occurs in response to chronic exercise training, i.e. the
athlete's heart) is reversible and is characterized by normal cardiac morphology (i.e. no brosis or
apoptosis) and normal or enhanced cardiac function. Given that there are clear functional, structural,
metabolic and molecular differences between pathological and physiological hypertrophy, a key question in
cardiovascular medicine is whether mechanisms responsible for enhancing function of the athlete's heart
can be exploited to benet patients with pathological hypertrophy and heart failure. This review summarizes
key experimental ndings that have contributed to our understanding of pathological and physiological
heart growth. In particular, we focus on signaling pathways that play a causal role in the development of
pathological and physiological hypertrophy. We discuss molecular mechanisms associated with features of
cardiac hypertrophy, including protein synthesis, sarcomeric organization, brosis, cell death and energy
metabolism and provide a summary of proling studies that have examined genes, microRNAs and proteins
that are differentially expressed in models of pathological and physiological hypertrophy. How gender and
sex hormones affect cardiac hypertrophy is also discussed. Finally, we explore how knowledge of molecular
mechanisms underlying pathological and physiological hypertrophy may inuence therapeutic strategies for
the treatment of cardiovascular disease and heart failure.
2010 Elsevier Inc. All rights reserved.
Pharmacology & Therapeutics 128 (2010) 191227
Abbreviations: 4E-BP1, 4E binding protein 1; ACE, Angiotensin converting enzyme; ACEI, Angiotensin converting enzyme inhibitors; Adr, adrenaline; Ang II, Angiotensin II; ANP,
Atrial natriuretic peptide; ARB, Angiotensin receptor blockers; ARK, Adrenergic receptor kinase; ARs, Adrenergic receptors; AT
1
, Angiotensin type 1 receptor; BNP, B-type natriuretic
peptide; ca, constitutively active; CaM, calmodulin; CaMK, calcium/calmodulin-dependent protein kinases; c-fos, c-fos oncogene; c-jun, c-jun oncogene; c-myc, c-myc oncogene; Cn/
CN, calcineurin; CoA, coenzyme A; CREB, cAMP response element-binding protein; CT-1, cardiotrophin 1; cTNT, cardiac troponin T; DAG, diacylglycerol; DCM, dilated
cardiomyopathy; dn, dominant negative; ECM, extracellular matrix; EGF, epidermal growth factor; EGFR, epidermal growth factor receptor; eIF2, eukaryotic initiation factor 2; eIF4E,
eukaryotic initiation factor 4E; ER, estrogen receptor; ERK, extracellular signal-regulated kinases; ET
A
, endothelin type A receptor; ET-I, endothelin 1; FAK, focal adhesion kinase;
GATA, GATA protein binding; Gp, guanine nucleotide binding proteins; GPCR, G protein-coupled receptor; Grb2, growth factor receptor bound protein 2; GSK3, glycogen synthase
kinase 3; HDAC, histone deacetylase; HIF, hypoxia-inducible factor; HSF1, heat shock transcription factor 1; Hsp, heat shock protein; HW/BW, heart weight/body weight ratio; IGF1,
insulin-like growth factor 1; IGF1R, insulin-like growth factor 1 receptor; IP
3
, inositol 1,4,5-trisphosphate; JAK, Janus kinase; JNK, c-Jun amino-terminal kinase; JVS, juvenile visceral
steatosis; KO, knockout; LIF, leukemia inhibitory factor; LVPW, left ventricular posterior wall; MAPK, mitogen activated protein kinase; MCAD, medium chain acyl coenzyme A
dehydrogenase; MCIP, mitogen-enriched calcineurin-interacting protein; mCPT-1, muscle-type carnitine palmitoyltransferase 1; MEF2, myocyte enhancer factor 2; MEK, mitogen
activated protein kinase kinase; MEKK, mitogen activated protein kinase kinase kinase; MHC, myosin heavy chain; miRNAs, microRNAs; MLC, myosin light chain; MLCK, myosin light
chain kinase; mTOR, mammalian target of rapamycin; Nab1, NGF1A-binding protein; NADPH, nicotinamide adenine dinucleotide phosphate; NE, noradrenaline, norepinephrine;
NFAT, nuclear factor of activated T cells; Ntg, non-transgenic; PDE, phosphodiesterase; PE, phenylephrine; PI3K, phosphoinositide 3-kinase; PKA, protein kinase A; PKC, protein
kinase C; PKD, protein kinase D; PLB, phospholamban; PLC, phospholipase C; PPAR, peroxisome proliferation-activated receptors; pS6, phosphorylation of 40 S ribosomal S6 protein;
Rab, member of RAS oncogene family; Raf1, member of RAS oncogene family; Ran, member of RAS oncogene family; Ras, Ras oncogene; Rho, rhodopsin; ROS, reactive oxygen
species; RTK, receptor tyrosine kinase; S6K, ribosomal S6 kinase; SERCA, sarcoplasmic reticulum Ca
2+
-ATPase; Src, Rous sarcoma oncogene; STAT, signal transducer and activator of
transcription; TAK, TGF- activated kinase; Tg, transgenic; TGF, transforming growth factor; TNF, tumor necrosis factor; TR, thyroid hormone receptor; Ub, ubiquitin; WT, wildtype.
Corresponding author. P.O. Box 6492 St Kilda Road Central, Melbourne Victoria 8008, Australia. Tel.: +61 385321194; fax: +61 385321100.
E-mail address: Julie.mcmullen@bakeridi.edu.au (J.R. McMullen).
0163-7258/$ see front matter 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.pharmthera.2010.04.005
Contents lists available at ScienceDirect
Pharmacology & Therapeutics
j our nal homepage: www. el sevi er. com/ l ocat e/ phar mt her a
Contents
1. Cardiac hypertrophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
2. Experimental and genetic mouse models utilized in the identication of
signaling pathways that mediate cardiac hypertrophy . . . . . . . . . . . . . . . . . . . . . . . . . . 192
3. Overview of signaling cascades/proteins implicated
in mediating physiological and pathological cardiac growth . . . . . . . . . . . . . . . . . . . . . . . 192
4. Molecular mechanisms associated with structural features of pathological and physiological hypertrophy . . . 193
5. Molecular mechanisms associated with differences in
energy metabolism in pathological and physiological hypertrophy . . . . . . . . . . . . . . . . . . . . 193
6. Characteristic gene expression changes associated with pathological and physiological hypertrophy . . . . 194
7. Gender differences in cardiac hypertrophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
8. Therapeutic strategies for the treatment of heart failure . . . . . . . . . . . . . . . . . . . . . . . . . 194
9. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
1. Cardiac hypertrophy
1.1. Introduction and overview
Cardiac hypertrophy can broadly be dened as an increase in heart
mass. Growth of the postnatal heart is closely matched to its
functional load (Zak, 1984). In response to an increase in load (e.g.
pressure overload in a setting of hypertension), the heart must work
harder than under normal conditions. To counterbalance the chronic
increase in wall stress the muscle cells within the heart enlarge
leading to an increase in size and mass (Cooper, 1987; Sugden & Clerk,
1998; Hunter & Chien, 1999). The increase in heart mass is largely due
to an increase in ventricular weight. In the subsequent sections we
have described cardiac hypertrophy at the cellular level, different
types of cardiac hypertrophy (pathological and physiological), the
molecular mechanisms responsible for different forms of cardiac
hypertrophy, gender differences, and possible treatment strategies
based on the distinct molecular mechanisms associated with
physiological and pathological cardiac hypertrophy.
1.2. Cardiac hypertrophy at the cellular level
The heart is composed of cardiac myocytes (muscle cells), non-
myocytes (e.g. broblasts, endothelial cells, mast cells, vascular smooth
muscle cells), and the surrounding extracellular matrix (Nag, 1980; Zak,
1984). Ventricular cardiac myocytes make up only one-third of the total
Fig. 1. Cellular processes involved in the development of cardiac hypertrophy. ECM: extracellular matrix, FAO: fatty acid oxidation, GPCR: G protein-coupled receptor, MAPK:
mitogen-activated protein kinase, PI3K: phosphoinositide 3-kinase, ROS: reactive oxygen species, SERCA: sarcoplasmic reticulum Ca
2+
ATPase.
196
197
207
209
210
212
214
217
217
217
192 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
cell number, but account for 7080%of the heart's mass (Nag, 1980; Zak,
1984; Popescu et al., 2006). In mammals, at birth or soon after, the
majority of cardiac myocytes lose the ability to proliferate, thus heart
growth occurs primarily via an increase in myocyte size (Soonpaa et al.,
1996). The inability of adult cardiac myocytes to divide has come under
some debate (Anversa & Nadal-Ginard, 2002; Anversa et al., 2002;
Pasumarthi & Field, 2002). However, estimates of DNA labeling indicate
that DNA synthesis is taking place in a very small fraction of the total
adult cardiac myocyte population(Nakagawa et al., 1988; Soonpaa et al.,
1996; MacLellan & Schneider, 2000; Anversa et al., 2002; Pasumarthi &
Field, 2002), indicating that the postnatal heart enlarges primarily by an
increase in myocyte size.
Myocytes are composed of bundles of myobrils. Myobrils contain
myolaments which consist of sarcomeres, the basic contractile unit of
the heart. Myocytes are arranged in a circumferential and spiral
orientation around the left ventricle, and need to contract simulta-
neously to ensure the heart pumps with a normal rhythm. Intercalated
discs, located at the bipolar ends of cardiac myocytes, are responsible for
maintaining cellcell adhesion while allowing contractile force to be
transmitted between adjacent cardiac myocytes (Estigoy et al., 2009).
Growth of cardiac myocytes is dependent on the initiation of several
events in response to anincrease in functional load, including activation
of signaling pathways, changes in gene expression, increases in the rate
of protein synthesis, and the organization of contractile proteins into
sarcomeric units (Fig. 1). Cardiac myocytes appear to have an intrinsic
mechano-sensing mechanism. Stretch sensitive ion channels present in
theplasmamembraneof cardiac myocytes andstructural proteins (such
as integrins) play a role in linking the extracellular matrix, cytoskeleton,
sarcomere, calcium handling proteins and nucleus (Knoll et al., 2003;
Hoshijima, 2006). Thus, thereis aninteractivecontinuumfromintegrins
at the cell surface to the contractile apparatus and nucleus (Fig. 1).
1.3. Association between cardiac hypertrophy and heart failure
Understanding the molecular mechanisms responsible for the
induction of cardiac hypertrophy has been of great interest due to the
known association between cardiac hypertrophy and nearly all forms of
heart failure (Levy et al., 1990). Cardiac hypertrophy is also an
independent risk factor for myocardial infarction, arrhythmia and
sudden death (Levy et al., 1990). In response to a chronic increase in
load, there is an initial increase in heart mass to normalize wall stress
and permit normal cardiovascular function at rest i.e. compensated
growth. However, if the chronic increase in wall stress is not relieved,
the hypertrophied heart can dilate, contractile function falls and the
heart can fail.
Heart failure affects approximately 13% of people in Western
society. The incidence of heart failure increases with age, affecting 34%
of those over 45 years old, 5% of those aged between 60 and 69 years of
age, and 10% of people over the age of 70 (Davies et al., 2001; AIHW,
2004; Thom et al., 2006; Lloyd-Jones et al., 2009). Symptoms of heart
failure patients include fatigue, insomnia, anxiety, depression, shortness
of breath, edema, dizziness, and nausea, all of which contribute to a
reduced quality of life for these patients (Blinderman et al., 2008). With
an aging population, rising rates of obesity and diabetes, as well as the
availability of interventions that prolong survival following cardiac
insults, the incidence of heart failure is likely to rise over the coming
decades. Thecosts associatedwithanexpandingnumber of patients and
specialized treatment strategies are expected to contribute signicantly
to the economic burden caused by heart failure (Blinderman, et al.,
2008). Currently there is nocure for heart failure, andlong termsurvival
following heart failure remains poor, with one third of patients dying
within a year of diagnosis (Zannad et al., 1999; Cowie et al., 2000;
Bleumink et al., 2004; McMurray & Pfeffer, 2005). Thus, a number of
studies have focused on identifying the molecular mechanisms
associated with cardiac hypertrophy and the transition to heart failure,
to identify new therapeutic targets to prevent or reverse cardiac
hypertrophy and heart failure.
1.4. Cardiac hypertrophy and the athlete's heart
The athlete's heart has generally beendened as a benignincrease in
heart mass, associated with morphological alterations, that represents a
Fig. 2. Cardiac hypertrophy can be classied as physiological, which occurs during pregnancy or in response to chronic exercise training, is reversible and characterized by normal
cardiac morphology and function. In contrast, hypertrophy that occurs in settings of disease is detrimental for cardiac structure and function and can lead to heart failure.
Developmental hypertrophy is associated with the normal growth of the heart after birth until adulthood. RV: right ventricle, LV: left ventricle. Normal/ physiological heart growth is
shown in green, pathological heart growth is shown in red.
193 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
physiological adaptation to chronic training. Though, with attention
from media reports of sudden death in young athletes, it has been
questioned whether highly trained athletes develop pathological
conditions. Notably, there is currently no evidence in the healthy
population (excluding persons with underlying cardiovascular disease
or genetic disorders) showing that remodeling due to exercise training
leads to long-termcardiac disease progression, cardiovascular disability,
or sudden cardiac death (Maron & Pelliccia, 2006). The overall risk of
sudden death in athletes is not well dened but considered low. A 12-
year survey inhigh-school athletes participating inorganized sports in a
US state (Maron et al., 1998) reported a frequency of sudden death of
1:200,000 per year (based on only 3 deaths among 1.4 million students;
including27sports). Suddendeathinyoungtrainedathletes inresponse
to physical exertion has largely been causally linked to congenital but
clinically unsuspected cardiovascular disease (see Maron, 2003). In
large autopsy-based surveys of athletes in the US, hypertrophic
cardiomyopathy is the most common cause of sudden death (account-
ing for about one-third of events), followed by congenital coronary-
artery anomalies. A wide range of other, largely congenital malforma-
tions account for the remaining sudden deaths from cardiovascular
disease among athletes (Maron, 2003). Thus, it is generally accepted
that cardiac hypertrophy in response to exercise is protective, in some
instances improves cardiac function, and does not progress to heart
failure. A comprehensive understanding of why cardiac hypertrophy
progresses to heart failure in a setting of disease, but does not in
response to exercise, is considered important for identifying and
targeting the critical molecular mechanisms responsible for the
transition from hypertrophy to heart failure.
1.5. Distinct forms of cardiac growth and hypertrophy
1.5.1. Pathological and physiological cardiac growth and hypertrophy
Cardiac growth or hypertrophy can broadly be classied as either
physiological (normal) or pathological (detrimental). Physiolog-
ical heart growth includes normal postnatal growth, pregnancy-
induced growth, and exercise-induced cardiac hypertrophy. In
contrast, pathological growth occurs in response to chronic pressure
or volume overload in a disease setting (e.g. hypertension, valvular
heart disease), myocardial infarction or ischemia associated with
coronary artery disease, or abnormalities/conditions that lead to
cardiomyopathy (e.g. inherited genetic mutations, diabetes) (Fig. 2).
Both physiological and pathological heart growth are associated with
an increase in heart size, however pathological hypertrophy is also
typically associated with loss of myocytes and brotic replacement,
cardiac dysfunction, and increased risk of heart failure and sudden
death (Levy et al., 1990; Weber et al., 1993; Cohn et al., 1997). In
contrast, physiological growth is associated with normal cardiac
structure, normal or improved cardiac function, and is reversible in
the instance of exercise- or pregnancy-induced hypertrophy (Ferrans,
1984; Schaible & Scheuer, 1984; Fagard, 1997) (Fig. 2).
1.5.2. Concentric and eccentric hypertrophy
Pathological and physiological hypertrophy has classically been
subdivided as concentric or eccentric. These classications are based on
changes in shape, which is dependent on the initiating stimulus
(Grossman et al., 1975; Pluim et al., 2000) (Fig. 3). Concentric
hypertrophy refers to an increase in relative wall thickness and
cardiac mass, with a small reduction or no change in chamber volume.
Concentric hypertrophy is characterized by a parallel pattern of
sarcomere addition leading to an increase in myocyte cell width
(Fig. 3). Eccentric hypertrophy refers to an increase in cardiac mass
with increased chamber volume, i.e. dilated chambers. Relative wall
thickness may be normal, decreased, or increased. In eccentric
hypertrophy, addition of sarcomeres in series leads to an increase in
myocyte cell length (Fig. 3) (Grossman et al., 1975).
A pathological stimulus causing pressure overload (e.g. hyperten-
sion, aortic stenosis) produces an increase in systolic wall stress which
results in concentric hypertrophy (Grossman et al., 1975). In contrast,
a stimulus causing volume overload (e.g. aortic regurgitation,
arteriovenous stulas) produces an increase in diastolic wall stress
and results in eccentric hypertrophy (Grossman et al., 1975; Pluim et
al., 2000). Clinical studies suggest that eccentric cardiac hypertrophy
induced by pathological stimuli poses a greater risk than concentric
cardiac hypertrophy (Berenji et al., 2005).
Physiological stimuli can also produce concentric or eccentric
hypertrophy. Aerobic exercise (also referred to as endurance training,
Fig. 3. Different stimuli induce different forms of cardiac hypertrophy. Pressure overload causes thickening of the left ventricle wall due to the addition of sarcomeres in parallel and
results in concentric hypertrophy. Volume overload induces an increase in muscle mass via the addition of sarcomeres in series and results in eccentric hypertrophy.
194 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
isotonic or dynamic exercise e.g. long-distance running, swimming)
and pregnancy increase venous return to the heart resulting in
volume overload and eccentric hypertrophy (Zak, 1984; Pluim et al.,
2000; Eghbali et al., 2005). This type of eccentric hypertrophy is
usually characterized by chamber enlargement and a proportional
change in wall thickness, whereas eccentric hypertrophy in settings of
disease is generally associated with thinning of the ventricular walls.
Strength training (also referred to as isometric or static exercise, e.g.
weight lifting, wrestling, throwing heavy objects) results in a pressure
load on the heart rather than volume load and concentric hypertrophy
(Zak 1984; Pluim et al., 2000) (Fig. 3).
1.6. Distinct features of pathological and physiological hypertrophy
Despite comparable increases in heart size, pathological and
physiological hypertrophy are associated with distinct 1) structural
and functional, 2) metabolic, and 3) biochemical and molecular
features (Fig. 4, Table 1).
Fig. 4. Four distinct features of cardiac hypertrophy include heart size, cardiac function, cardiac brosis and gene expression. The upper left quadrant shows representative images of
mouse hearts that have been subjected to a pathological (aortic banding; band) stimulus for one week, compared with sham (sham) operated controls, or mice subjected to a
physiological stimulus for four weeks (chronic swim training; exercise) compared with sedentary controls. An increase in heart size is observed in mice that have undergone aortic
banding (band) or chronic swim training (exercise) compared to sham and sedentary controls. The upper right quadrant depicts cardiac function as shown by M-mode
echocardiography. Cardiac function is depressed in a mouse model of pathological (decompensated) hypertrophy but is conserved in a mouse model of physiological hypertrophy
(exercise). The lower left quadrant depicts histological analysis of heart sections stained with Masson's trichrome; increased brosis is shown in blue and is only present in
pathological hypertrophy. Representative sections from the left ventricular wall of untrained mice (control), aortic banded mice (pressure overload, band) and swim trained mice
(exercise). Bar = 10 m. Sections from sham operated mice were similar to those of untrained (control). The lower right quadrant shows gene expression changes associated with
cardiac hypertrophy. Representative Northern blot showing total RNA from ventricles of sham, aortic banded (band), untrained (sedentary) and trained (exercise) mice. Expression
of GAPDH was determined to verify equal loading of RNA. There is increased expression of atrial natriuretic peptide (ANP), B-type natriuretic peptide (BNP), -myosin heavy chain
(-MHC) and -skeletal actin, and decreased expression of sarcoplasmic reticulum Ca
2+
ATPase (SERCA) and -MHC in mice subjected to aortic banding (band) compared to
controls, while gene expression remains relatively unchanged in exercised trained mice compared to sedentary controls.
Table 1
Features of pathological and physiological hypertrophy.
Feature Pathological cardiac hypertrophy Physiological cardiac hypertrophy
Stimuli Disease Aerobic exercise training
Pressure or volume overload Postnatal growth
Cardiomyopathy (familial, viral, diabetes, metabolic, alcoholic/toxic) Pregnancy
Cardiac morphology Increased myocyte volume Increased myocyte volume
Formation of new sarcomeres Formation of new sarcomeres
Increase in heart size Increase in heart size
Cardiac brosis Yes No
Apoptosis Yes No
Fetal gene expression Upregulation of ANP, BNP, -MHC, and -skeletal actin Relatively unchanged
Expression of genes associated with contractile function Downregulation of SERCA2a, -MHC Normal or increased
Cardiac function Depressed Normal or enhanced
Metabolism Decreased fatty acid oxidation Enhanced fatty acid oxidation
Increased glucose utilization Enhanced glucose utilization
Reversible No Yes
Association with heart failure and mortality Yes No
195 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
1.6.1. Structural and functional features
Cardiac hypertrophy is associated with structural remodeling of
components of the ventricular walls to accommodate increases in
myocyte size, including changes in the brillar collagen network
and angiogenesis. Under basal conditions or a setting of physiolog-
ical hypertrophy, the brillar collagen network provides structural
integrity of adjoining myocytes, facilitating myocyte shortening
which translates into efcient cardiac pump function (Gunasinghe
& Spinale, 2004). Pathological hypertrophy is associated with cell
death (apoptosis, necrosis) and the loss of myocytes is replaced
with excessive collagen (known as brosis). The main brillar
collagen present in cardiac brosis is type 1 collagen. Excessive
accumulation of collagen stiffens the ventricles, which impairs
contraction and relaxation, impairs the electrical coupling of cardiac
myocytes with extracellular matrix proteins, and reduces capillary
density. Fibrosis and reduced capillary density increases oxygen
diffusion distances, leading to myocardial ischemia, and is likely to
contribute to the transition from hypertrophy to failure (Guna-
singhe & Spinale, 2004).
1.6.2. Cardiac metabolism
In the normal healthy heart, fatty acid oxidation is the main
metabolic pathway responsible for generating energy, accounting for
6070% of ATP production (van der Vusse et al., 1992); glucose and
lactate metabolism account for approximately 30% of ATP synthesis.
The heart is capable of switching energy substrates depending on
workload and the relative concentrations of fuel molecules in the
bloodstream (see van der Vusse et al., 1992). This is considered an
adaptive mechanism which allows the heart to produce a continuous
supply of ATP under various physiological conditions (e.g. fasting,
during exercise, etc.).
Pathological cardiac hypertrophy is associated with decreases in
fatty acid oxidation and increases in glucose metabolism (Allard et al.,
1994; Christe & Rodgers 1994; Davila-Roman et al., 2002). This switch
in substrate utilization may be a protective mechanism, allowing the
heart to produce more ATP per molecule of oxygen consumed (see
van Bilsen et al., 2009). This is reminiscent of what occurs during fetal
cardiac development, when oxygen supply is limited and fatty acid
transport and metabolism are impaired (due to carnitine deciency
and delayed maturation of enzymes involved in fatty acid oxidation).
Thus, glucose is the primary substrate used by the fetal heart to
generate ATP (Ostadal, et al., 1999). In contrast, physiological cardiac
hypertrophy induced by exercise training is characterized by
enhanced fatty acid and glucose oxidation (Gertz et al., 1988). Of
note, in advanced pathological hypertrophy and failure, glucose
metabolism decreases as the heart becomes resistant to insulin,
reducing the overall ability of the heart to generate sufcient ATP (see
Neubauer, 2007).
1.6.3. Biochemical and molecular features
In the 1970s and early 1980s it was recognized that physiological
hypertrophy (induced by exercise training/ thyroid hormone) was
associated with elevations in myosin ATPase activity and enhance-
ment of contractility, whereas pathological hypertrophy (induced by
renal hypertension, aortic banding) was associated with decreased
myosin ATPase activity and depressed contractile function (Wikman-
Coffelt et al., 1979; Rupp, 1981). Since then, there have been a number
of studies demonstrating that physiological and pathological cardiac
hypertrophy are associated with some distinct biochemical and
molecular signatures. Iemitsu et al. (2001) compared mRNA expres-
sion in a rat model of pathological cardiac hypertrophy (spontane-
ously hypertensive rat) and physiological hypertrophy (chronic swim
training). The investigators reported a distinct pattern of gene
expression in the two models (Iemitsu et al., 2001). It is now well
recognized that pathological cardiac hypertrophy is associated with
distinct alterations in cardiac contractile proteins (- and -myosin
heavy chain (MHC)), increased expression of fetal genes (e.g. atrial
natriuretic peptide (ANP), B-type natriuretic peptide (BNP), -
skeletal actin) but down-regulation of calcium-handling proteins
(e.g. sarcoplasmic reticulum Ca
2+
-ATPase 2a (SERCA2a)). Since this
time, generation of numerous transgenic and knockout mouse models
in combination with models of physiological and pathological
hypertrophy have allowed investigators to delineate signaling
proteins that appear to play distinct roles in regulating physiological
and pathological cardiac hypertrophy (discussed in detail in
Section 3).
2. Experimental and genetic mouse
models utilized in the identication of
signaling pathways that mediate cardiac hypertrophy
The denition of cardiac hypertrophy as either physiological or
pathological has not been without contention (Dorn et al., 2003).
However, there is now substantial evidence from animal studies that
the different phenotypes associated with pathological and physiolog-
ical hypertrophy can be due to distinct stimuli and activation of some
distinct signaling pathways, at least under certain conditions.
Studies utilizing genetic mouse models (transgenic and knockout)
alone or in combination with morphologically distinct models of
hypertrophy (e.g. pathological, physiological, concentric and eccen-
tric) have become powerful tools for understanding molecular
pathways responsible for different forms of heart growth in vivo.
Genetic mouse models have typically utilized the -MHC promoter to
achieve cardiac myocyte specic expression (Subramaniam et al.,
1991), or the Cre-loxP systemto generate cardiac-specic or inducible
knockout mice (Chien, 2001). Inducible transgenic mouse models
have also been valuable as they allow investigators to switch on the
activity of the protein of interest in myocytes by injection of a drug
(e.g. tamoxifen) at a specic time point, i.e. after completion of
developmental growth (Fan et al., 2005b; Hoesl et al., 2008; Lu et al.,
2009; Ruan et al., 2009). Commonly used experimental mouse models
of pathological hypertrophy include pressure overload (constriction/
banding of the renal, abdominal, ascending or transverse aorta),
volume overload (aortocaval shunt), and minipump infusions of
vasoactive substances (e.g. isoproterenol, angiotensin II (Ang II)).
Physiological models include treadmill running, freewheel running
and chronic swim training. Below, we have described hypertrophic
triggers/stimuli, signaling proteins and cascades which appear to play
an important role for the development of pathological or physiological
hypertrophy.
2.1. Hypertrophic triggers/stimuli
In response to hemodynamic overload, cardiac myocytes are
subjected to mechanical stretch, and autocrine and paracrine humoral
factors including Ang II, endothelin 1 (ET-1), insulin-like growth
factor 1 (IGF1), transforming growth factor- (TGF-) and cardio-
trophin 1 (CT-1) are released. These factors bind to receptors on
cardiac cells, in turn activating intracellular signaling pathways that
leads to cell growth. Signaling cascades and proteins responsible for
cardiac growth and hypertrophy are complex and extensive crosstalk
has been identied (Fig. 5). The subsequent sections of this review
focus on signaling cascades and proteins that have been reported to
play distinct roles in regulating pathological and physiological
hypertrophy.
2.1.1. Physiological and pathological triggers/stimuli
Human and animal studies have demonstrated that certain factors
are preferentially released in response to pathological and physiolog-
ical stimuli. It is well recognized that IGF1 is released during postnatal
development and in response to exercise training (Yeh et al., 1994;
Conlon and Raff 1999; Koziris et al., 1999; Neri Serneri et al., 2001b;
196 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Perrino et al., 2006), and IGF1 levels in the heart were increased in
swim-trained rats (Scheinowitz et al., 2003). Furthermore, production
of cardiac IGF1 (but not Ang II or ET-1) was increased in professional
athletes compared with control subjects (Neri Serneri et al., 2001b). In
contrast, pressure overload is associated with elevated levels of Ang II,
catecholamines and ET-1 (Schunkert et al., 1990; Arai et al., 1995;
Yamazaki et al., 1999; Rapacciuolo et al., 2001; Yayama et al., 2004),
and cardiac formation of Ang II was increased in heart failure patients
with hypertrophied hearts (Neri Serneri et al., 2001a).
3. Overview of signaling cascades/proteins implicated
in mediating physiological and pathological cardiac growth
The best characterized signaling cascades responsible for mediat-
ing physiological and pathological cardiac hypertrophy are the IGF1-
phosphoinositide 3-kinase [PI3K, (p110)]-Akt pathway and Gq
signaling (downstream of G protein-coupled receptors (GPCR)
activated by Ang II, ET-1 and catecholamines), respectively (Fig. 6).
Other signaling pathways associated with physiological cardiac
hypertrophy and/or protection include the gp130/JAK/STAT pathway,
thyroid hormone signaling, and heat shock transcription factor 1
(HSF1). In contrast, pathological hypertrophy has also been associated
with abnormalities leading to enhanced PI3K(p110), mitogen
activated protein kinases (MAPKs), protein kinase C (PKC) and D
(PKD), and calcineurin.
3.1. Signaling proteins/pathways
implicated in mediating physiological hypertrophy
3.1.1. IGF1-PI3K(p110)-Akt pathway
Substantial evidence from genetic mouse models has demonstrat-
ed the critical role of the IGF1-PI3K(p110)-Akt pathway in
regulating physiological cardiac growth.
3.1.1.1. Insulin-like growth factor (IGF1) receptor signaling. IGF1 is best
known for being produced by the liver in response to growth
hormone stimulation and is essential for normal fetal and postnatal
growth and development (Adams et al., 2000). IGF1 is also produced
by the heart (see reviews: Ren et al., 1999; McMullen, 2008) and binds
to a cell surface receptor, insulin-like growth factor 1 receptor
(IGF1R), a receptor tyrosine kinase that activates downstream
signaling proteins. A number of studies have examined the role of
IGF1 in the heart using gene targeted mice.
a) Mice with increased cardiac myocyte specic expression of IGF1
(human IGF1B transgene expression driven by the -MHC
promoter) had enlarged hearts with normal cardiac function
(Reiss et al., 1996). A confounding factor of this study was that
transgene expression increased IGF1 secretion from cardiac
myocytes which resulted in a signicant rise in systemic plasma
levels of IGF1 (approximately 80%) and an increase in other organ
weights. The increase in heart size was attributed to an increase in
Fig. 5. A schematic of the major signaling pathways involved in cardiac hypertrophy, showing cross-talk and integration of various pathways. N.B. due to the complex nature of
signaling cascades and on-going discoveries it was not possible to illustrate all interactions.
197 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
cardiac myocyte number rather than myocyte size. This result was
unexpected given that the majority of mammalian cardiac
myocytes are thought to lose their ability to proliferate at birth
or within the rst few weeks of postnatal life (Soonpaa et al.,
1996).
b) In a second study, IGF1 transgenic mice were generated using the
-skeletal actin promoter (transgene expression in heart and
skeletal muscle) (Delaughter et al., 1999). Serum IGF1 levels were
not elevated but persistent transgenic expression was associated
with increases in gut, liver, and spleen weight (Fiorotto et al.,
2003). Up to 10 weeks of age, IGF1 transgenic mice displayed
cardiac hypertrophy, which was associated with enhanced cardiac
systolic function i.e. physiological hypertrophy (Delaughter et al.,
1999). However, the investigators concluded that the physiolog-
ical cardiac phenotype ultimately progressed to pathological
hypertrophy, because mice had depressed cardiac function by
12 months of age (Delaughter et al., 1999).
c) We examined the role of IGF1 specically in cardiac myocytes by
over-expressing the IGF1R rather than IGF1 (transgenic expression
of human IGF1R using the -MHC promoter) (McMullen et al.,
2004b). Expression of the IGF1R was considered an advantage
because it allowed examination of IGF1 signaling in the absence of
effects of secreted IGF1 on other tissues or non-myocytes. At
3 months of age, IGF1R transgenic mice had enlarged hearts
(approximately 40% increase in weight, with a proportional
increase of all chambers and ventricular wall thickness), increased
myocyte size, no evidence of histopathology (e.g. necrosis, brosis,
myocyte disarray) and enhanced systolic function (McMullen et
al., 2004b). These characteristic features of physiological hyper-
trophy were maintained at 1216 months of age (IGF1R transgenic
mice had enhanced systolic function in comparison to non-
transgenic mice) (McMullen et al., 2004b). Thus, IGF1R transgenic
mice did not progress to pathological hypertrophy with aging,
which was observed in the earlier study (model b) (Delaughter et
al., 1999). Consistent with the hypothesis that IGF1 activates PI3K
(p110) and Akt (a known downstream target of PI3K) to induce
physiological cardiac hypertrophy (Fig. 6), activation of PI3K and
phosphorylation of Akt were elevated in hearts of IGF1R transgenic
mice (McMullen et al., 2004b). In contrast, signaling proteins
downstream of Gq (implicated in pathological hypertrophy)
including MAPKs and calcineurin were not activated in hearts of
IGF1R transgenic mice (McMullen et al., 2004b).
d) In corroboration with the idea that IGF1 signaling is critical for
physiological heart growth, cardiac myocyte-specic ablation of
the IGF1R gene in mice attenuated the hypertrophic response to
swim exercise training compared to non-transgenic mice (Kim et
al., 2008b). A basal cardiac phenotype was not observed.
3.1.1.2. Phosphoinositide 3-kinase (PI3K, p110) signaling. PI3Ks are a
family of enzymes and have been linked to a diverse group of cellular
functions, particularly cell growth, survival, differentiation, and
proliferation (Cantley, 2002). PI3K is a lipid kinase that releases
inositol lipid products from the plasma membrane which in turn
mediate intracellular signaling (Toker & Cantley, 1997; Vanhaeseb-
roeck et al., 1997). Activation of PI3Ks is coupled to both receptor
tyrosine kinases (e.g. insulin receptor and IGF1R) and GPCRs. There
are three major classes of PI3Ks (classes I, II and III), which are
classied based on sequence homology in the catalytic domain,
structure and substrate specicity (Vanhaesebroeck et al., 2001; Kok
et al., 2009). Class I PI3Ks are heterodimers and further divided into
Fig. 6. A schematic overviewof pathological and physiological hypertrophy outlining key differences in initiating stimuli, signaling pathways, cellular responses and cardiac function.
For simplicity we have focussed on the best characterized signaling pathways implicated in mediating pathological (shaded red) and physiological (shaded green) cardiac
hypertrophy. Other important mediators are described in detail in Section 3. Ang II: angiotensin II, ET-1: endothelin-1, GPCR: G protein-coupled receptor, IGF-1: insulin-like growth
factor 1, MAPK: mitogen-activated protein kinase, NE: norepinephrine, PI3K(p110): phosphoinositide 3-kinase p110, RTK: receptor tyrosine kinase.
198 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
the subclasses I
A
and I
B
. Class I
A
PI3Ks consist of a p110 catalytic
subunit (, or ) and a p85 or p55 regulatory subunit. The only Class
I

PI3K is p110, which is regulated by p101 (Vanhaesebroeck et al.,


1997). Of the Class I PI3Ks, p110 and p110 are abundantly
expressed in the heart. p110 is also expressed in the heart but at a
lower level (Crackower et al., 2002). The p110 isoform of PI3K is
exclusively expressed in leukocytes (Vanhaesebroeck & Watereld,
1999).
PI3Ks were rst shown to regulate organ size in Drosophila. Over-
expression of the Drosophila PI3K homolog, Dp110, resulted in
formation of larger wings and eyes (Leevers et al., 1996). In contrast,
expression of a catalytically inactive Dp110 caused the opposite
phenotype, i.e. smaller wings and eyes (Leevers et al., 1996). Null
homozygous mice for p110 were embryonically lethal due to
proliferation defects in the embryo (Bi et al., 1999). Evidence that
PI3K activity regulates heart size was obtained from studies that
expressed a cardiac-specic constitutively active (ca) form of PI3K
(p110) or dominant negative (dn) form of PI3K(p110) in
transgenic mice utilizing the -MHC promoter (Shioi et al., 2000;
McMullen et al., 2003). PI3K activity was increased by 6.5-fold in
hearts of caPI3K transgenic mice, and hearts were 20% larger than
non-transgenic mice. The increase in heart size of the caPI3K mice was
proportional, resembling physiological hypertrophy. In contrast, PI3K
activity was 77% lower in hearts of dnPI3K transgenic mice, which
resulted in a 17% decrease in heart size compared to non-transgenic
mice (Shioi et al., 2000). Changes in heart size were due to changes in
myocyte size rather than number (Shioi et al., 2000). Cardiac function,
structure and life span were normal in caPI3K and dnPI3K transgenic
mice under basal conditions (Shioi et al., 2000; McMullen et al., 2003).
These studies demonstrated that PI3K(p110) is critical for physio-
logical postnatal growth of the heart.
It was later demonstrated that PI3K(p110) is also critical for
physiological exercise-induced growth of the heart but not patholog-
ical hypertrophy. Adult dnPI3K transgenic mice were subjected to a
physiological stimulus (exercise; chronic swim training) and a
pathological stimulus (pressure overload; ascending aortic banding).
dnPI3K mice showed signicant hypertrophy in response to pressure
overload, but an attenuated hypertrophic response to swim training,
compared with non-transgenic mice (McMullen et al., 2003; McMul-
len et al., 2007). These studies were also conrmed using a muscle-
specic knockout approach of the p85/p55/p50 and p85
(global) regulatory subunits in mice (Luo et al., 2005), as well as
cardiac-specic ablation of p110 (Lu et al., 2009). These mice showed
a decrease in heart weight to body weight ratio of approximately 20%
and 16%, respectively (Luo et al., 2005; Lu et al., 2009), similar to that
reported in dnPI3K mice (Shioi et al., 2000; McMullen et al., 2003).
The small heart phenotype of p85/p85 knockout mice was
accompanied by a reduction in mean myocyte cell area, and mice
exhibited an attenuated hypertrophic response to exercise training
(Luo et al., 2005). Together, these studies support the critical role of
class I
A
PI3Ks in the regulation of physiological cardiac hypertrophy.
Finally, to assess whether PI3K(p110) was the critical mediator
responsible for physiological growth in the IGF1R transgenic mice, we
crossed transgenic mice over-expressing IGF1R with dnPI3K trans-
genic mice and examined heart size. Hearts of double transgenic mice
(i.e. expressing both the IGF1R and dnPI3K transgenes) were not
signicantly different in size to that in dnPI3K mice alone (McMullen
et al., 2004b). This result demonstrated that the physiological heart
growth in the IGF1R transgenic mice was dependent on PI3K(p110)
signaling (McMullen et al., 2004b).
3.1.1.3. Akt. Akt, a serine/threonine kinase (also known as protein
kinase B), is a well characterized target of PI3K. The Akt family is
involved in a number of cellular processes, including cell survival, cell
cycle, metabolism, and protein synthesis. There are three isoforms of
Akt (Akt1, Akt2 and Akt3), each is encoded by distinct genetic loci, in
which the genes code for enzymes that are members of the serine/
threonine-specic protein kinase family (Matsui & Rosenzweig,
2005). The Akt homologue Dakt regulates cell and organ growth in
Drosophila, in the same manner as PI3K; activation of Dakt increased
cell size whereas loss-of-function caused a reduction in cell size, but
had no impact on cell proliferation (Verdu et al., 1999; Rintelen et al.,
2001). In mammals, Akt1 null mice had a 20% reduction in body
weight (Cho et al., 2001). Although all three isoforms are broadly
expressed, only Akt1 and Akt2 are highly expressed in the heart
(Matsui & Rosenzweig 2005).
Initial characterization of the cardiac phenotypes of Akt transgenic
mice led to confounding results. Phenotypes ranged from absence of
hypertrophy associated with protection from ischemiareperfusion
injury to substantial hypertrophy associated with a pathological
phenotype and premature death (Condorelli et al., 2002; Matsui et al.,
2002; Shioi et al., 2002; Shiraishi et al., 2004; Shiojima et al., 2005). The
varying phenotypes have been attributed to different degrees of Akt
activation, angiogenesis, and subcellular localization. Of note, Akt canbe
activated by both receptor tyrosine kinases (e.g. IGF1R) and GPCR
(Fig. 5), and appears to be differently regulated depending on the
initiating stimulus. Myostatin, an inhibitor of cardiac growth, reduced
GPCR-induced Akt phosphorylation but not receptor tyrosine kinase
(IGF1R)-induced phosphorylation in neonatal cardiac myocytes (Mor-
issette et al., 2006). The biological signicance of this differential
activation is currently unclear.
More recent studies in Akt knockout mice suggest Akt1 is required
for physiological rather than pathological heart growth. Akt1
knockout mice (normal cardiac phenotype under basal conditions)
showed a blunted hypertrophic response to swim training but not to
pressure overload (DeBosch et al., 2006b). These ndings are
reminiscent of those in mice with reduced PI3K activity (McMullen
et al., 2003; Luo et al., 2005). It is now generally accepted that Akt1
mediates cardiac cell growth whereas Akt2 is important for cardiac
metabolism (DeBosch et al., 2006a,b).
Glycogen synthase kinase 3 (GSK3), a cellular substrate of Akt, is
an important regulatory kinase with a number of cellular targets,
including cytoskeletal proteins and transcription factors. Both GSK3
isoforms (GSK3 and GSK3) are expressed in the heart (Ferkey &
Kimelman, 2000; Harwood, 2001; Hardt & Sadoshima, 2002). Initial
reports demonstrated that GSK3 negatively regulated heart growth
and that inhibition of GSK3 by hypertrophic stimuli was an
important mechanism for stimulating growth (Haq et al., 2000;
Morisco et al., 2000; Morisco et al., 2001; Antos et al., 2002; Badorff et
al., 2002). More recent studies have shown that GSK3 inhibits
postnatal cardiac growth and reduces pressure overload-induced
hypertrophy (Zhai et al., 2007). However, transgene expression was
associated with increased brosis and apoptosis both under basal
conditions and during pressure overload. Furthermore, the reduced
hypertrophic phenotype in response to pressure overload was
associated with severe cardiac dysfunction and heart failure (Zhai et
al., 2007). Interestingly, the GSK3 and GSK3 isoforms appear to
have distinct roles in a setting of pressure overload. Phosphorylation
of GSK3 was essential for the development of pathological
hypertrophy whereas phosphorylation of GSK3 played a compen-
satory role (Matsuda et al., 2008). Thus, selective modulation of the
phosphorylation status of the two isoforms may be required to
maximize the therapeutic potential of modulating this kinase.
3.1.2. Gp130/JAK/STAT pathway
Leukemia inhibitory factor (LIF), CT-1, and other members of the
interleukin-6 cytokine family activate the gp130 receptor associated
with the LIF receptor (Fig 5). Once activated, this cytokine receptor
interacts with janus kinase 1 (JAK1), leading to phosphorylation of the
signal transducer and activator of transcription (STAT) class of
transcription factors (Kodama et al., 1997; Pellegrini & Dusanter-
Fourt 1997; Aoki & Izumo, 2001; Molkentin & Dorn 2001). Cardiac-
199 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
specic transgenic mice over-expressing STAT3 displayed cardiac
hypertrophy that was protective against doxorubicin-induced car-
diomyopathy (Kunisada et al., 2000). In contrast, mice with
ventricular deletion of gp130 had normal cardiac structure and
function under basal conditions, but displayed a rapid onset of dilated
cardiomyopathy in response to pressure overload (Hirota et al., 1999).
However, expression of a dominant negative mutant of gp130 (to
decrease activation of this pathway) appeared to protect transgenic
mice against pressure overload-induced hypertrophy (Uozumi et al.,
2001). Despite this discrepancy, the majority of data in the literature
suggests that the gp130/JAK/STAT pathway has a protective role in the
heart (see Fischer & Hilker-Kleiner, 2007; Boengler et al., 2008;
Fischer & Hilker-Kleiner, 2008).
3.1.3. Thyroid hormone receptor signaling
Thyroid hormone is a classic hormonal mediator of normal
postnatal heart growth. The thyroid gland secretes two biologically
active hormones: thyroxine (T
4
, prohormone) and triiodothyronine
(T
3
). T
4
and T
3
diffuse across the plasma membrane and T
4
is
converted to T
3
(Danzi & Klein 2002; Dillmann, 2002). Postnatal heart
growth was reduced in a setting of depressed thyroid gland activity,
whereas administration of excess thyroid hormone to animals led to
an increase in heart weight (Bedotto et al., 1989; Hudlicka & Brown
1996). Furthermore, administration of T
4
in humans was associated
with increased heart mass but no fall in systolic function (Ching et al.,
1996), and patients with chronic hyperthyroidism have increased
cardiac contractility which is often associated with cardiac hypertro-
phy (Forfar et al., 1982; Feldman et al., 1986). Thus, it has been
suggested that thyroid hormone induces physiological heart growth.
The biological effects of thyroid hormone have largely been
attributed to nuclear transcriptional mechanisms. T
3
passes through
the nuclear membrane to bind to nuclear thyroid hormone receptors
(TRs), which act as transcription factors to directly repress or activate
cardiac genes (Lazar & Chin 1990; Lazar 1993; Mangelsdorf et al.,
1995; Danzi & Klein, 2002; Dillmann, 2002; Harvey & Williams, 2002).
Thyroid hormone has been shown to regulate - and -MHC, cardiac
troponin, SERCA2a, and voltage gated potassium channels (Izumo et
al., 1986; Rohrer & Dillmann, 1988; Nishiyama et al., 1998; Danzi &
Klein, 2002). In mammals TRs are encoded by two genes: TRand TR
(Lazar, 1993; Harvey & Williams, 2002). Thyroid hormone treatment
increased heart mass by approximately 64% in wildtype mice, 44% in
TRknockout mice but only 6% in TR knockout mice, suggesting that
TR plays a predominant role in regulating heart growth (Weiss et al.,
2002).
More recently, cytosolic and membrane-initiated effects of thyroid
hormones have been reported (Bassett et al., 2003; Farach-Carson &
Davis, 2003). Kenessey and Ojamaa demonstrateda direct interaction of
cytosol-localized TR1 with the p85 regulatory subunit of PI3K in
neonatal rat ventricular myocytes (Kenessey & Ojamaa, 2006). This
interaction was shown to be critical for T
3
-induced protein synthesis.
The authors concluded that rapid T
3
mediated activation of PI3K by
TR1 may underlie the mechanisms via which thyroid hormone
induces physiological heart growth (Kenessey & Ojamaa, 2006) (Fig 5).
3.1.4. Heat shock transcription factor 1
Sakamoto and colleagues identied HSF1 as a promising critical
mediator of physiological cardiac hypertrophy froma genetic proling
study that compared gene expression in hearts from rats subjected to
pressure overload with exercise trained rats (voluntary running
wheel). Gene expression of HSF1, which regulates heat shock proteins
(Hsp) including Hsp70 and Hsp27, was upregulated in hearts from
exercise trained rats but not in hearts subjected to pressure overload
(Sakamoto et al., 2006). To examine whether HSF1 had a protective
role in exercise-induced cardiac hypertrophy, HSF1-decient hetero-
zygote mice (HSF1
+/
) were subjected to voluntary wheel running
for 4 weeks. Interestingly, exercise-induced hypertrophy was not
blunted in HSF1
+/
but cardiac function was signicantly reduced
(Sakamoto et al., 2006).
3.2. Signaling proteins/pathways
implicated in mediating pathological hypertrophy
3.2.1. G proteins
Gproteins can be divided into two main subgroups: heterotrimeric
G proteins and small-molecular-weight monomeric G proteins (small
G proteins).
3.2.1.1. Heterotrimeric G proteins. Heterotrimeric G proteins consist of
three subunits (, and ) and couple to GPCR. Binding of an agonist
to the GPCR leads to dissociation of the G and G subunits,
followed by activation of downstream signaling pathways (Gutkind,
1998a,b; Rockman et al., 2002). Isoforms of the heterotrimeric G
proteins are largely determined by the isoform of the subunits,
which fall into four subfamilies: G
s
, G
i
, G
12
, and G
q
(e.g. G
q
, G
11
)
(Simon et al., 1991; Neer, 1995).
In response to a pathological stimulus (e.g. pressure overload),
hormones/vasoactive factors such as Ang II, ET-1 and noradrenaline
(norepinephrine, NE) are released and induce cardiac growth
(Schunkert et al., 1990; Arai et al., 1995; Yamazaki et al., 1999;
Rapacciuolo et al., 2001; Yayama et al., 2004). These ligands bind to
GPCR: Ang II receptor type 1 (AT
1
receptor), endothelin receptors (ET
A
and ET
B
) and
1
-adrenergic receptors (ARs), respectively. This causes
activation of G
q/11
and downstream signaling proteins, including
phospholipase C (PLC), MAPKs, PKC and protein kinase A (PKA).
Transgenic mouse studies have highlighted the critical role of G
q/11
in mediating pathological cardiac hypertrophy (Fig. 6). Cardiac-
specic transgenic mice over-expressing G
q
developed cardiac
hypertrophy that was associated with cardiac dysfunction and
premature death (D'Angelo et al., 1997; Mende et al., 1998). In
contrast, mice lacking G proteins (G
q/11
) in cardiac myocytes and
cardiac-specic transgenic mice expressing a peptide specic for
inhibiting Gq-coupled receptor signaling displayed no hypertrophy or
a signicantly blunted response to pressure overload (Akhter et al.,
1998; Wettschureck et al., 2001). Taken together, these studies
suggest that the G
q/11
pathway is important for the induction of
pathological hypertrophy.
3.2.1.1.1. Angiotensin II receptors. Ang II is the principal vasoactive
substance of the reninangiotensin system with a variety of
pathophysiological actions in the cardiovascular system via systemic
and local effects including vasoconstriction, aldosterone release, and
cell growth (Zimmerman & Dunham, 1997; de Gasparo et al., 2000).
Two pharmacologically distinct Ang II receptors have been cloned
(AT
1
and AT
2
). Rodents have two AT
1
receptor isoforms (AT
1A
and
AT
1B
) (de Gasparo et al., 1995; Lorell 1999). The heart contains a local
reninangiotensin system which is activated in response to hemody-
namic stress (e.g. pressure overload) (Yamazaki & Yazaki, 1997;
Lijnen & Petrov, 1999). It is well established that blocking Ang II
formation with angiotensin converting enzyme inhibitors (ACEI)
attenuates pressure overload-induced hypertrophy in animal models
and humans (Sadoshima et al., 1996; Zhu et al., 1997; Lijnen and
Petrov 1999; Yamazaki et al., 1999; Devereux 2000; Modesti et al.,
2000).
Hypertrophy due to activation of the AT
1
receptor in rodent cardiac
neonatal myocytes or mouse models has been associated with
activation of MAPKs, increased intracellular calcium, PKC, and
transactivation of the epidermal growth factor receptor (EGFR)
(Sadoshima & Izumo, 1993; Miyata & Haneda, 1994; Sadoshima et
al., 1995; Kagiyama et al., 2002; Thomas et al., 2002; Chan et al., 2006).
Under basal conditions, single global Ang II receptor knockout mice
(AT
1A
, AT
1B
, AT
2
) have been reported to have no cardiac phenotype or
a small decrease in heart mass (Hein et al., 1995; Ichiki et al., 1995;
Hamawaki et al., 1998; Harada et al., 1998; Oliverio et al., 1998;
200 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Gembardt et al., 2008). Interestingly, double knockouts lacking AT
1A
and AT
1B
had smaller hearts than AT
1A
alone, whereas the reduction
was less pronounced in triple knockout mice lacking AT
1A
, AT
1B
and
AT
2
(Gembardt et al., 2008). Cardiac-specic AT
1
transgenic mice
developed pathological hypertrophy and died prematurely of heart
failure (Paradis et al., 2000). Unexpectedly, pressure overload-
induced hypertrophy was not blunted in AT
1A
knockout mice
(Hamawaki et al., 1998; Harada et al., 1998), possibly suggesting
that AT
1B
is able to compensate for loss of AT
1A
. Mice with a gain of
function mutated AT
1A
receptor did not display cardiac hypertrophy
but developed progressive cardiac brosis and diastolic function
abnormalities (Billet et al., 2007). Though, this study is confounded by
a modest increase in blood pressure in the mutant mice. Cardiac-
specic AT
2
transgenic mice had no cardiac phenotype (Masaki et al.,
1998), but pressure overload-induced hypertrophy was inhibited in
AT
2
knockout mice (Senbonmatsu et al., 2000). While it is reasonably
clear that Ang II is critical for mediating cardiac hypertrophy, the
precise role of Ang II receptor subtypes requires further examination
(see Billet et al., 2008)). In addition, recent studies have demonstrated
the cardiovascular effects of a number of breakdown products of Ang
II including Ang 17, Ang III and Ang IV; as well as new mechanisms
concerning the functional regulation of Ang II receptors such as
receptor dimerization, ligand-independent activation, receptor-inter-
acting proteins, and the existence of an agonistic antibody against the
AT
1
receptor (Jones et al., 2008; Mogi et al., 2009). The role of these
peptides and new regulation of Ang II receptors in relation to cardiac
hypertrophy also requires further investigation.
3.2.1.1.2. Endothelin-1 receptors. ET-1 is the predominant endothe-
lin in the heart and is a potent hypertrophic stimulus in neonatal
cardiac myocytes (Shubeita et al., 1990). ET-1 binds to two GPCRs: ET
A
and ET
B
. ET
A
receptors account for 90% of endothelin receptors on
cardiac myocytes (Kedzierski & Yanagisawa, 2001). Ang II increased
ET-1 levels in rat cultured cardiac myocytes, and an ET
A
antagonist
(BQ123) reduced Ang II-induced myocyte hypertrophy (Ito et al.,
1993). ET
A
receptor antagonists (alone or in combination with ET
B
receptor antagonists) have been shown to attenuate pathological
hypertrophic responses in animal models (see Brunner et al., 2006).
Though, cardiac myocyte specic ET
A
knockout mice displayed a
normal hypertrophic response to Ang II (Kedzierski et al., 2003).
A number of clinical trials have assessed the potential of ET-1
antagonists to treat heart failure patients; results have largely been
disappointing, and/or remain unpublished (reviewed by (Seed et al.,
2001; Rich & McLaughlin, 2003; Kelland & Webb, 2006; Kirkby et al.,
2008)). In the majority of clinical trials using ET-1 antagonists [either
non-selective for ET
A
or ET
B
(e.g. bosentan, tezosentan) or selective for
ET
A
(e.g. darusentan)] heart failure patients developed adverse side
effects without improvement in cardiac remodeling or clinical
symptoms (Louis et al., 2001; Kalra et al., 2002; Luscher et al., 2002;
Rich & McLaughlin, 2003; Anand et al., 2004; Packer et al., 2005;
Kaluski et al., 2008).
3.2.1.1.3. Adrenergic receptors (ARs). Catecholamines activate ARs
(members of the GPCR superfamily) (Scheuer 1999; Lomasney &
Allen, 2001; Rockman et al., 2002). There are three major AR
subfamilies:
1
-AR,
2
-AR, and -AR. At least 6 types of ARs are
present in the mammalian heart (three
1
-ARs:
1A
,
1B
,
1D
and
three -ARs:
1
,
2
,
3
), with
1
-ARs predominating, accounting for
approximately 80% of the -ARs in the healthy heart (Xiang & Kobilka,
2003; Barki-Harrington et al., 2004; Salazar et al., 2007; Woodcock et
al., 2008). ARs are coupled to G
q
, G
s
, and G
i
, leading to
modulation of adenylate cyclase, PLC and ion channels (Rockman et
al., 2002). Specically,
1A
,
1B
, and
1D
-ARs activate G
q
signaling,

1
-ARs couple to G
s
, and
2
-ARs couple to G
s
and G
i
(Exton, 1985;
Garcia-Sainz et al., 1999; Rockman et al., 2002). Species-dependent
differences exist between the functions of
1
- and
2
-AR subtypes in
the heart which may be attributed to differential coupling to G
s
and
G
i
(see Kaumann et al., 1999). In the human heart,
2
-ARs appear to
largely couple to the G
s
/cAMP pathway (Kaumann et al., 1999). The
role of
3
-ARs in the heart remains unclear (see Barki-Harrington et
al., 2004; Kaumann & Molenaar, 2008).
Heart failure patients have elevated circulating catecholamines
and increased adrenergic drive, which initially increases contractility
and may be benecial. However, prolonged adrenergic drive is
detrimental and associated with desensitization and downregulation
of -ARs (Bristow, 2000). This is consistent with ndings from
cardiac-specic transgenic mice over-expressing
1
-ARs. Before
15 weeks of age, transgenic mice displayed increased cardiac
contractility compared to controls. However, cardiac function pro-
gressively fell in
1
-AR transgenic mice after 16 weeks and the mice
rapidly developed cardiac dysfunction and heart failure (Engelhardt
et al., 1999). Progressive deterioration in cardiac function with
chronic transgenic expression of
1
-AR was later conrmed by
another report (Bisognano et al., 2000).
Cardiac-specic transgenic mice expressing a dominant negative
-AR receptor kinase 1 (-ARK1/GRK2; dominant negative mutant
restores -ARsignaling; -ARK1phosphorylates -ARs leading to their
desensitization) were protected against pathological hypertrophy and
heart failure (Koch et al., 1995), and targeted deletion of GRK2 in
cardiac myocytes of mice prevented and rescued heart failure induced
by myocardial infarction(Raake et al., 2008). Inthis respect, it has been
difcult to explain why -AR agonists are poorly tolerated in heart
failure patients but -blockers have a protective role (Bristow, 2003;
Molenaar &Parsonage, 2005, discussedin further detail inSection8.1).
Though it has been suggested that at the molecular level, inhibition of
-ARK1/GRK2 shares a number of properties with -blockade as
opposed to -AR agonism (Rockman et al., 2002). For instance, -AR
agonists promote desensitization and receptor downregulation due to
constant activation of the -AR system. In contrast, inhibition of -
ARK1 may allow -ARs to return to a more normal state of signaling
because desensitization will be inhibited (Rockman et al., 2002).
While it is generally accepted that chronic stimulation of the -AR
system has an adverse effect on the heart that contributes to the
pathogenesis of heart failure, preservation of normal -AR-G protein
coupling is critical during times of need, such as periods of stress and
during exercise (Christensen & Galbo, 1983; Lefkowitz et al., 2000;
Rockman et al., 2002). In this instance, acute versus chronic activation
of -ARs may explain differences in phenotype observed with
pathological and physiological hypertrophy. Plasma resting levels of
catecholamines were signicantly higher in a mouse model of
pathological hypertrophy induced by chronic pressure overload in
comparison to a mouse model of physiological hypertrophy induced
by swim training (Perrino et al., 2006).
The role of AR subtypes in mediating cardiac hypertrophy based on
genetically modied mouse models has been described in detail (see
Du, 2008). Global knockout mice decient of
1
and
2
-ARs displayed
an attenuated hypertrophic response to pressure overload, with
reduced brosis (Kiriazis et al., 2008). Based on various transgenic and
knockout models,
1B
-AR and
2
-AR appear to contribute to
maladaptation and the onset of heart failure in a setting of pressure
overload, whereas activation of
1A
-AR may be benecial (Du, 2008).
Of note, in rat neonatal cardiac myocytes, it was shown that
2
-ARs
that couple with G
i
proteins mediate cardiac protection due to
activation of the PI3K-Akt pathway (Chesley et al., 2000). Further-
more,
1
-ARs were critical for normal postnatal heart growth in male
but not female mice (O'Connell et al., 2003), and protected the heart
against pressure overload-induced maladaptive hypertrophy (O'Con-
nell et al., 2006).
3.2.1.2. Small G proteins (also called GTPases). The family of small G
proteins can be divided into 5 subfamilies (Ras, Rho, ADP ribosylation
factors, Rab, and Ran). Small G proteins act as molecular switches,
which link receptors to downstream signaling cascades. Ras and Rho
can be activated in myocytes in response to Ang II, ET-1,
201 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
phenylephrine (PE) and mechanical stress and have been implicated
in the development of cardiac hypertrophy (Ramirez et al., 1997a;
Aoki et al., 1998; Aikawa et al., 1999; Chiloeches et al., 1999; Clerk and
Sugden, 2000; Clerk et al., 2001). Cardiac-specic transgenic mice
expressing a constitutively active form of Ras or over-expressing
Rab1a developed pathological cardiac hypertrophy (Hunter et al.,
1995; Wu et al., 2001). Furthermore, using a Rho-kinase inhibitor
(Fasudil), it was shown that Rho-kinase was critical for pressure
overload-induced pathological hypertrophy in rats but not swim-
ming-induced physiological cardiac hypertrophy (Balakumar & Singh,
2006).
3.2.2. PI3K(p110) signaling
In contrast to the p110 isoform of PI3K (coupled to RTKs e.g.
IGF1R), PI3K(p110) is coupled to GPCRs (specic GPCR not all
denitively dened but thought to include G
s
/G
i
/G
q
e.g. adren-
ergic receptors, Ang II receptors and endothelin receptors) and
appears to have a detrimental effect in the heart (Oudit et al., 2004).
PI3K(p110) does not affect heart size under basal conditions but is a
negative regulator of cardiac contractility, as PI3K(p110) knockout
mice displayed enhanced contractile function (Crackower et al.,
2002). PI3K(p110) may have an impact on heart growth in settings
of pathological stress (Naga Prasad et al., 2000; Oudit et al., 2003),
although the role of PI3K(p110) in the diseased heart is complex and
appears to differ depending on the nature of the pathological
stimulus. PI3K(p110) knockout mice were protected from heart
failure induced by chronic activation of -ARs, displaying less
hypertrophy and brosis, and better cardiac function than controls
(Oudit et al., 2003). However, PI3K(p110) knockout mice displayed
an accelerated progression to dilated cardiomyopathy in response to
pressure overload (Patrucco et al., 2004; Oudit & Kassiri, 2007). PI3K
(p110) might contribute to cardiac dysfunction via its effects on -
AR internalization and regulation of phosphodiesterases (Fig. 5)
(Oudit & Kassiri, 2007; Pretorius et al., 2009b). Downregulation and
desensitization of -ARs is detrimental for heart function (Bristow et
al., 1982; Perrino et al., 2007), and is dependent on the binding of
p110 to -ARK1 (Naga Prasad et al., 2001). Expression of a
catalytically inactive p110 mutant or disruption of the interaction
between -ARK1 and p110 restored -AR signaling and contractile
function in transgenic mice subjected to chronic -AR stimulation
(Nienaber et al., 2003; Perrino et al., 2005; Perrino et al., 2007). PI3K
(p110) may also reduce cardiomyocyte contractility by regulating
the activity of phosphodiesterases (PDEs) (Patrucco et al., 2004;
Kerfant et al., 2007). PDEs hydrolyse cAMP, a second messenger which
plays a critical role in mediating Ca
2+
release from the sarcoplasmic
reticulum to induce contraction. PDE inhibitors improve contractile
function by increasing intracellular cAMP levels, however the safety of
PDE inhibitors as therapeutic agents in patients with heart failure is
still being investigated (see Osadchii, 2007 for review).
3.2.3. Mitogen activated protein kinase (MAPK) pathways
MAPKs are divided into 3 subfamilies based on the terminal kinase
in the pathway: the extracellular signal-regulated kinases (ERKs), the
c-Jun amino-terminal kinase (JNKs), and the p38-MAPKs (Clerk &
Sugden, 1999; Widmann et al., 1999; Pearson et al., 2001). All three
types of MAPKs are activated in cultured cardiac myocytes in response
to GPCR agonists (couple to G
q
: AT
1
receptors, endothelin receptors
and
1
-ARs) and mechanical stress, as well as in pressure overloaded
hearts and failing human hearts, but the exact role of MAPKs has
remained unclear (Yamazaki et al., 1993; Sadoshima et al., 1995;
Komuro et al., 1996; Sugden & Clerk, 1998; Cook et al., 1999; Esposito
et al., 2001; Pearson et al., 2001; Takeishi et al., 2001; Purcell et al.,
2007).
3.2.3.1. ERK1/2. ERKs are protein kinases that phosphorylate a range of
cytosolic and nuclear substrates (see Chen et al., 2001b for review).
ERK1/2 are ubiquitously expressed (Boulton et al., 1991) and
activation has been reported in numerous settings of cardiac
hypertrophy and failure (see Muslin, 2008) however it is still unclear
whether ERK1/2 is a critical mediator of hypertrophic responses.
ERK1/2 was activated in response to agonists that induce pathological
heart growth, such as Ang II, ET-1 and NE, but not in response to the
physiological hypertrophic agonist IGF1 (Clerk et al., 2006). In isolated
cardiac myocytes, activation of ERK1/2 was essential for protein
synthesis (a requirement for cell growth) following stimulation with
hypertrophic agonists that signal via G
q
protein coupled receptors
(Wang & Proud, 2002). Consistent with this nding, expression of a
dominant negative mutant of Raf-1 (a MAPK kinase kinase down-
stream of G
q
; Fig. 5) blunted cardiac hypertrophy in mice subjected
to pressure overload, implicating ERK1/2 in the development of
pathological cardiac hypertrophy (Harris et al., 2004). However,
transgenic mice expressing cardiac-specic constitutively active
MAPK kinase 1 (MEK1; a MAPK kinase immediately upstream of
ERK1/2; does not activate JNK or p38-MAPK) developed a physiolog-
ical rather than a pathological phenotype, which was characterized by
concentric cardiac hypertrophy, enhanced systolic cardiac function
and no interstitial brosis (Bueno et al., 2000). Furthermore, loss of
ERK1 (global knockout mice) had no effect on heart size in mice
subjected to pressure overload or swim training, indicating that ERK1
is not a critical mediator of pathological or physiological hypertrophy,
or that the remaining ERK2 activity was sufcient to drive the
hypertrophic response (Purcell et al., 2007). Reduced expression of
ERK2 (ERK2
+/
mice; ERK2
/
mice are embryonically lethal) alone
or in mice decient for ERK1 (ERK1
/
ERK2
+/
mice) also failed to
block hypertrophy induced by pressure overload or swim training
(Purcell et al., 2007). Together, these studies suggest that activation of
ERK1/2 is sufcient, but not critical, for inducing cardiac hypertrophy,
although the results of the latter study were confounded by the fact
that the ERK1
/
and ERK2
+/
mice were not cardiac-specic.
Fig. 7. ERK1/2 appears to contribute to hypertrophic responses via two distinct
mechanisms. Stimulation of GPCRs leads to dissociation of Gq proteins. A) The G
subunit of Gq activates traditional MAP kinase signaling cascades, resulting in
phosphorylation and activation of ERK1/2 by MEK1/2. This in turn leads to protein
synthesis and cell growth. B) Association of G subunits with the Raf1/MEK/ERK1/2
complex is necessary for autophosphorylation of ERK1/2 at residue Thr188 and
localization of ERK1/2 in the nucleus. This leads to phosphorylation of nuclear targets
(such as Elk1, MSK1 and c-Myc) and transcription of hypertrophic genes.
202 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Induction of ERK1/2 kinase activity requires phosphorylation of
the threonine and tyrosine residues within the TEY motif of the
activation loop by MEK1/2 (Anderson et al., 1990; Payne et al., 1991;
Robbins et al., 1993). A recent study reported that autophosphoryla-
tion of ERK2 at residue Thr188 promoted nuclear localization and
subsequent phosphorylation of hypertrophic factors, including Elk1,
MSK1 and c-Myc (Lorenz et al., 2009). This occurred following
association of the Raf/MEK/ERK complex with the subunits of G
q
proteins (Lorenz et al., 2009). Thr188 phosphorylation appears to be a
key event in the development of ERK1/2-mediated cardiac hypertro-
phy, as transgenic mice with suppressed ERK2 Thr188 phosphoryla-
tion were resistant to cardiac hypertrophy induced by pressure
overload, while mice with enhanced ERK2 Thr188 phosphorylation
displayed more pronounced hypertrophy compared to wildtype mice
(Lorenz et al., 2009). Phosphorylation of Thr188 was also evident in
biopsies from human failing hearts, suggesting that ERK2 autopho-
sphorylation of Thr188 is clinically relevant (Lorenz et al., 2009). Thus,
ERK1/2 appears to contribute to cardiac hypertrophic responses via
two distinct mechanisms (Fig. 7). Activation of the traditional MAPK
signaling cascade (Raf/MEK/ERK) following binding of G
q
proteins to
GPCRs results in phosphorylation of ERK1/2 within the TEY motif and
induction of ERK1/2 kinase activity. Subsequent phosphorylation of
substrates (such as p90 ribosomal S6 kinase (Lorenz et al., 2009)) may
contribute to cell hypertrophy by increasing protein synthesis. This
mechanism may be responsible for driving the physiological hyper-
trophy observed in the MEK1 transgenic mice. Secondly, interaction of
the Raf/MEK/ERK complex with G proteins causes autophosphor-
ylation of ERK2 at Thr188. This results in nuclear localization, allowing
ERK1/2 to phosphorylate nuclear targets, which in turn promotes
transcription of hypertrophic genes. This event may be critical for
inducing the maladaptive phenotype associated with pathological
hypertrophic responses (Fig 7.)
3.2.3.2. ERK5. ERK5 (also known as big MAP kinase 1, BMK1) may play
a role in mediating pathological eccentric cardiac hypertrophy, as
transgenic mice expressing constitutively active MEK5 developed
eccentric cardiac hypertrophy which progressed to dilated cardiomy-
opathy and death (Nicol et al., 2001).
3.2.3.3. JNKs. The JNK family consists of at least ten isoforms, derived
from three genes: JNK1, JNK2 and JNK3 (Waetzig & Herdegen, 2005).
JNK1 and JNK2 are ubiquitously expressed, whereas JNK3 has a more
restricted expression prole in the heart, brain and testis (Waetzig &
Herdegen, 2005). JNK was activated in hearts from heart failure
patients (Cook et al., 1999) and in the remote myocardium of
infarcted rat hearts (Li et al., 1998). A number of in vitro studies have
suggested that JNKs may be important regulators of pathological
hypertrophy (Bogoyevitch et al., 1996; Ramirez et al., 1997a;
Choukroun et al., 1998; Wang et al., 1998b; Choukroun et al., 1999),
although in vivo studies have been more difcult to interpret
(described below and summarized in Table 2).
JNK is phosphorylated and activated by MAPK kinase 4 (MEK4) and
MEK7 (Fig. 5), and preferentially upregulated by MAPK kinase kinase 1
(MEKK1) (Fig 5). Transgenic mice withcardiac-specic activationof JNK
(via constitutive activation of MEK7) did not develop cardiac hypertro-
phy but died prematurely from congestive heart failure (Petrich et al.,
2003). Several loss of function approaches have also been utilized to
determinethe role of JNKs inpathological cardiac hypertrophy. Pressure
overload-induced hypertrophy was attenuated in dnMEK4 transgenic
mice(Choukrounet al., 1999), andMEKK1was essential for pathological
cardiac hypertrophy and dysfunction induced by cardiac-specic
transgenic expression of Gq (Minamino et al., 2002). These data
suggest JNKs may be necessary regulators of pathological cardiac
hypertrophy. However, dnJNK1/2 transgenic mice and JNK1/2 gene-
targetedmicedisplayedanenhancedhypertrophic response topressure
overload (Liang et al., 2003), suggesting JNKs antagonize cardiac
growth. Furthermore, cardiac-specic MEK4 knockout mice with
reduced JNK activity displayed normal cardiac growth and function
under basal conditions (Liu et al., 2009). However, following aortic-
banding and chronic -adrenergic stimulation, unlike the dnMEK4
transgenic mice, cardiac specic MEK4 knockout mice had enhanced
cardiac growth, increased hypertrophic gene transcription and ventric-
ular brosis compared to wildtype aortic-banded controls. Following
swim training, the hypertrophic response was unchanged compared to
wildtype controls inthis model (Liuet al., 2009). Null MEKK1 transgenic
mice had comparable heart weights to wildtype mice under basal
conditions. In response to aortic-banding, MEKK1
/
mice displayed an
enhanced hypertrophic response rather than a blunted response after
Table 2
JNK mouse models and their phenotype under basal conditions and in response to pressure overload as assessed by heart weight/body weight or left ventricular/body weight ratios.
Mouse model Basal phenotype Pressure overload Reference
Versus Ntg/WT Hypertrophic response versus Ntg/WT
caMEK7
(cardiac-specic)
n/a Petrich et al., 2004
dnMEK4
(cardiac-specic by adenovirus-mediated gene transfer)
n/a Choukroun et al., 1999
dnJNK1/2
(cardiac-specic)
Liang et al., 2003
Jnk1
+/
/Jnk2
/
(targeted deletion, global)
Liang et al., 2003
MEK4
/
(cardiac-specic)
Liu et al., 2009
MEKK1
/
(targeted deletion, global)
Sadoshima et al., 2002
Jnk1
/
(targeted deletion, global)
Tachibana et al., 2006
Jnk2
/
(targeted deletion, global)
Tachibana et al., 2006
Jnk3
/
(targeted deletion, global)
Tachibana et al., 2006
Key:
= JNK has no effect on heart growth, possibly due to redundancy of isoforms.
= Indicates JNK antagonizes pathological growth.
= JNK is essential for pathological growth.
n/a = not examined/assessed.
203 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
14 days compared to wildtype controls (Sadoshima et al., 2002).
Furthermore, aortic-banded MEKK1
/
mice had a higher mortality
rate and congestive heart failure compared to wildtype banded mice.
This study implies that cardiac hypertrophy induced by pressure
overload occurs in the absence of JNK activation. In further support of
the latter, mice with selective deletion of the three JNK genes (Jnk1
/
,
Jnk2
/
and Jnk3
/
) subjected to aortic-banding developed cardiac
hypertrophy that was comparable to wildtype mice (Tachibana et al.,
2006). Thus, it appears individual members of the JNK family are not
required to induce cardiac growth, or that the JNK isoforms are
functionally redundant.
3.2.3.4. p38-MAPK. p38-MAPK is an important mediator of numerous
biological functions including cell growth, cell proliferation, cell cycle
and cell death, and is considered a critical component of stress
response pathways (Wilson et al., 1996; Bassi et al., 2008). In the
heart, p38-MAPK is known to be activated during ischemia, and p38-
MAPK activity was increased in the myocardium from patients with
ischemic heart disease (Cook et al., 1999). p38-MAPK has been
implicated in the regulation of cardiac gene expression, cardiac
myocyte apoptosis, myocyte hypertrophy, contractility, remodeling
and metabolism (Liao et al., 2001; Petrich & Wang, 2004; Baines &
Molkentin, 2005; Wang, 2007).
The p38-MAPK family consists of four isoforms (, , and ),
though it appears only and are expressed in the heart (Jiang et al.,
1997; Clark et al., 2007). The p38 isoform is predominately
expressed in the human and rodent myocardium (Lemke et al.,
2001; Braz et al., 2003). The role of p38 and p38 in the heart has
been examined extensively using both transgenic and knockout
mouse models. As with JNK, many of the studies appear to be
contradictory (summarized in Table 3). Cardiac-specic transgenic
mice expressing a dominant negative form of p38 were generated
and analyzed by two independent groups (Braz et al., 2003; Zhang et
al., 2003a). Zhang et al. (2003a) reported no basal phenotype in dn-
p38 transgenics. In contrast, Braz et al. (2003) found that dn-p38
transgenics developed concentric hypertrophy associated with a fall
in cardiac function and elevated fetal gene expression (ANP and BNP),
with the majority of mice dying prematurely fromcardiomyopathy by
8 months of age. Both groups subjected the dn-p38 transgenics to
pressure overload and reported a signicant increase in heart size.
Zhang et al. reported a hypertrophic response (to pressure overload)
in the dn-p38 transgenics similar to control animals but, interest-
ingly, the transgenics had signicantly less brosis (Zhang et al.,
2003a). In contrast, Braz et al. (2003) reported an exaggerated
hypertrophic response in dn-p38 transgenics compared with
controls in response to pressure overload or 14 day minipump
infusions of phenylephrine, Ang II and isoproterenol. Finally,
cardiac-specic p38 knockout mice displayed normal cardiac
structure and function under basal conditions (Nishida et al., 2004).
In response to pressure overload, knockout mice developed a similar
degree of cardiac hypertrophy to controls, but displayed greater
cardiac dysfunction, more brosis and apoptosis (Nishida et al., 2004).
The authors concluded that p38 plays a critical role in protecting the
heart in a setting of pressure overload.
Cardiac-specic dn-p38 transgenic mice had no cardiac hypertro-
phy under basal conditions but seemed to have reduced systolic
function (Zhang et al., 2003a). The hypertrophic response to pressure
overload was not different from that observed in non-transgenic mice,
though the dn-p38 transgenics were reported to display less brosis
(Zhang et al., 2003a).
MEK3 and MEK6 are regulators of p38-MAPK (Fig 5). Under basal
conditions dnMEK3 and dnMEK6 transgenic mice developed patho-
logical hypertrophy by 2 and 8 months of age, respectively. Both
models displayed cardiac dysfunction, and brosis was also reported
in dnMEK3 transgenics at 4 months of age. The majority of dnMEK3
mice died by 8 months of age due to cardiomyopathy (Braz et al.,
2003). In response to pressure overload, dnMEK3 and dnMEK6
transgenic mice displayed an exacerbated cardiac hypertrophic
response, increased brosis and depressed cardiac function, similar
to the observations in dn-p38 transgenic mice (Braz et al., 2003).
Activation of p38-MAPK in either MEK3 or MEK6 transgenic hearts
under baseline conditions led to the increased expression of the fetal
gene program, substantial induction of interstitial brosis, and loss of
contractility (Liaoet al., 2001). Bothtransgenic mousemodels developed
heart failure, although this was not associated with hypertrophy of
cardiac myocytes (Liao et al., 2001). Inthis set of studies, p38-MAPKdoes
not appear to promote hypertrophy, but seems to contribute to brosis,
loss of contractility and the development of dilated cardiomyopathy.
Explanations for the possible discrepancies between many of these
models (see Table 3) may be attributed to the generation of transgenic
mice on different genetic backgrounds (Braz et al., 2003; Zhang et al.,
2003a), gender differences (Liu et al., 2006) (discussed in detail in
Table 3
p38 mouse models and their phenotype under basal conditions and in response to pressure overload as assessed by heart weight/body weight or left ventricular/body weight ratios.
Mouse model Basal phenotype Pressure overload Reference
Versus Ntg/WT Hypertrophic response versus Ntg/WT
dn-p38
(cardiac-specic, Black Swiss background)
Zhang et al., 2003a
dn-p38
(cardiac-specic, FVB/N background)
Braz et al., 2003
dn-p38
(cardiac-specic females compared to males)
Liu et al., 2006
p38 CKO
(cardiac-specic KO)
Nishida et al., 2004
dn-p38
(cardiac-specic)
Zhang et al., 2003a
dnMEK3
(cardiac-specic)
Braz et al., 2003
dnMEK6
(cardiac-specic)
Braz et al., 2003
MEK3
(cardiac-specic)
n/a Liao et al., 2001
MEK6
(cardiac-specic)
n/a Liao et al., 2001
Key:
= p38 has no effect on heart growth.
= p38 is essential for pathological hypertrophy.
n/a = not examined/assessed.
204 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Section 7), p38-MAPK having an anti-apoptotic function that is
independent of kinase activity or a biphasic dose response curve to
p38-MAPK (Muslin, 2008), as well as distinct roles of p38and that
may bedifferentially regulatedby MEK3andMEK6(Wanget al., 1998a).
3.2.3.5. Protein kinases. Extracellular stimuli such as pressure overload
activate PKC and PKD via GPCRs to trigger hypertrophic responses
(Dorn & Force, 2005; Harrison et al., 2006) (Fig 5).
3.2.3.5.1. PKC. PKC is considered a critical signal transducer
downstream of G
q
. There are at least 12 isoforms of PKC and at least
4 (, , , and ) have been implicated in the induction of cardiac
hypertrophy (see Dorn & Force, 2005). PKC and PKC are
conventional Ca
2+
-dependent isoforms whereas PKC and PKC are
novel Ca
2+
-independent isoforms (Mackay & Mochly-Rosen, 2001;
Sabri & Steinberg, 2003). Mice null for PKC, , , or had no obvious
cardiac phenotype under basal conditions (see Dorn & Force, 2005),
however it has been suggested that the role of specic PKC isoforms
may be masked by compensatory signaling by other PKC isoforms
(Dorn & Force, 2005). Cardiac-specic PKC transgenic mice devel-
oped cardiac hypertrophy associated with cardiac dysfunction,
brosis and premature death (Bowman et al., 1997; Wakasaki et al.,
1997; Chen et al., 2001a), but knockout mice displayed a typical
hypertrophic response to a GPCR agonist (PE) or aortic-banding
(Roman et al., 2001). Thus, it appears that PKC is not required for the
pathological hypertrophic response. Cardiac-specic transgenic mice
over-expressing PKC or PKC displayed mild concentric hypertrophy
with a physiological phenotype (no evidence of brosis, and normal
cardiac function) (Takeishi et al., 2000; Chen et al., 2001a). However,
in response to a cardiac insult (ischemia-induced damage), PKC had
a protective role, whereas activation of PKC exacerbated the damage
(Chen et al., 2001a). PKC appears to be critical for regulating cardiac
contractility but not cardiac hypertrophy (Braz et al., 2002; Hahn et
al., 2003; Braz et al., 2004). Transgenic mice with over-expression of
PKC had diminished cardiac contractility while PKC
/
mice had
improved cardiac contractility (Braz et al., 2004). Furthermore,
inhibition of PKC activity in a model of pathological hypertrophy
(G
q
transgenic mice) improved cardiac contractility, whereas
activation of PKC resulted in a lethal cardiomyopathy (Hahn et al.,
2003). There were no apparent effects on heart size under basal
conditions or in response to cardiac stress (e.g. pressure overload)
with the PKC mouse models.
3.2.3.5.2. PKD. Cardiac-specic transgenic mice expressing a
constitutively active form of PKD1 developed pathological hypertro-
phy and died prematurely (Harrison et al., 2006). In contrast, mice
with conditional cardiac-specic deletion of PKD1 had no phenotype
under basal conditions but displayed a blunted hypertrophic response
to various pathological models (pressure overload, Ang II-dependent
hypertrophy and isoproterenol-dependent hypertrophy) which was
associated with better cardiac function, less brosis, and less fetal
gene activation compared to control mice (Fielitz et al., 2008).
3.2.4. Calcium signaling
Each heartbeat is associated with entry of calcium into cardiac
myocytes. Calcium is central to the control of contractile function and
cardiac growth. Calcium/calmodulin is an important second messen-
ger for GPCR agonists and biomechanical stress (Frey et al., 2000; Aoki
& Izumo, 2001; Sugden, 2001). The best described calcium-dependent
signaling proteins include calcineurin and calcium/calmodulin-de-
pendent protein kinases (CaMKs). Calcineurin is a serinethreonine
phosphatase that consists of a catalytic A subunit and a regulatory B
subunit. Two regulatory subunit genes (B1, B2) have been identied,
and three genes encode the A catalytic subunit, calcineurinA (CnA),
calcineurinA (CnA) and calcineurinA (CnA). Only the and
genes have been shown to be expressed in human, mouse and rat
hearts (Klee et al., 1998; Molkentin & Dorn, 2001). Calcineurin
dephosphorylates nuclear factor of activated T cells (NFAT) transcrip-
tion factors which promotes nuclear translocation and activation of
gene transcription (Fig. 5).
Following stimulation with GPCR hypertrophic agonists (Ang II
and PE) in cultured rat neonatal cardiac myocytes, calcineurin
enzymatic activity, CnA (but not CnA or CnA) mRNA and protein
levels were increased (Taigen et al., 2000). Calcineurin activity was
also increased in hypertrophied and failing hearts from human
patients (Haq et al., 2001), and human failing heart ventricular
muscle exposed to ET-1, Ang II and urotensin II (Li et al., 2005).
Furthermore, calcineurin activity was upregulated in hypertrophied
hearts following aortic-banding in rodents (Shimoyama et al., 1999;
Lim et al., 2000; De Windt et al., 2001; Zou et al., 2001; Saito et al.,
2003). Finally, transgenic mice expressing an activated form of
calcineurin in the heart developed profound cardiac hypertrophy
which rapidly progressed to dilated cardiomyopathy, with extensive
interstitial brosis, congestive heart failure and sudden death, often
by 3 months of age (Molkentin et al., 1998). Taken together, these
studies imply that elevated calcineurin induces pathological cardiac
hypertrophy.
Consistent with the idea that calcineurin/NFAT coupling induces
pathological cardiac growth, when NFAT-luciferase reporter mice
were subjected to both physiological stimuli (exercise training,
growth hormone-IGF1 infusion) and pathological stimuli (pressure
overload, myocardial infarction), NFAT luciferase reporter activity was
upregulated in both pathological models but not in the physiological
models (Wilkins et al., 2004). In addition, transgenic mice with
targeted inactivation of calcineurin A displayed an impaired
hypertrophic response to pressure overload and infusion of GPCR
agonists (Bueno et al., 2002), and cardiac hypertrophy in transgenic
mice expressing a dominant negative form of calcineurin A displayed
a blunted hypertrophic response to pressure overload compared to
wildtype mice (Zou et al., 2001). Finally, pharmacological inhibition of
calcineurin activity prevented cardiac hypertrophy in constitutively
active calcineurin A transgenic mice (Molkentin et al., 1998).
As noted previously, calcineurin is considered to regulate the
pathological hypertrophic response via dephosphorylation of NFAT
(Fig. 5)(Olson & Williams, 2000). NFAT translocates to the nucleus,
where it associates with other transcription factors such as GATA4 and
myocyte enhancer factor 2 (MEF2), to regulate the expression of
cardiac genes (Wilkins et al., 2002; Frey & Olson, 2003). In support of
this, Molkentin et al. (1998) showed that cardiac-specic transgenic
mice expressing a constitutively activate mutant form (nuclear
localized) of NFAT3 developed cardiac hypertrophy and heart failure,
whereas expression of the wildtype protein did not lead to
hypertrophy (Molkentin et al., 1998).
The contribution of calcineurin in mediating cardiac hypertrophy
has also been examined utilizing myocyte-enriched calcineurin-
interacting protein 1 (MCIP1) transgenic mice. MCIP is able to inhibit
calcineurin signaling by binding directly to the catalytic subunit
(CnA), which inactivates its ability to dephosphorylate NFAT and
MEF2 (Fig. 5). Forced over-expression of MCIP1 selectively in the
heart of transgenic mice was able to attenuate hypertrophy and
prevent progression to dilated cardiomyopathy in response to various
pathological stimuli, including aortic-banding, transgenic expression
of CnA, and the -adrenergic receptor agonist isoproterenol (Rother-
mel et al., 2001; Hill et al., 2002). This is consistent with the idea that
enhanced calcineurin signaling mediates pathological growth.
Though, interestingly, transgenic expression of MCIP also inhibited
exercise-induced hypertrophy (voluntary running wheel) (Rothermel
et al., 2001). Studies utilizing MCIP1 null mice have also been
complicated to interpret. MCIP
/
mice had a normal cardiac
phenotype under basal conditions, displayed an exaggerated hyper-
trophic response to transgenic expression of CnA, but a blunted
hypertrophic response to pressure overload and adrenergic stimula-
tion (isoproterenol) (Vega et al., 2003). The authors concluded that
MCIP1 may have a dual role in hypertrophic signaling, acting as a
205 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
suppressor or activator, depending on the initiating stimulus (Vega et
al., 2003).
Calcium/calmodulin-dependent protein kinase II (CaMKII) is a
serine/threonine protein kinase that has been implicated in cardiac
hypertrophy and heart failure. Upregulation of CaMKII has been
reported in hearts of patients and animal models with heart failure
(Hoch et al., 1999; Kirchhefer et al., 1999; Bossuyt et al., 2008). Four
isoforms (, , and ) of CaMKII exist, which are encoded by
separate genes. CaMKII is the predominant isoform in the heart and
has distinct splice variants (
A
,
B
,
C
); is also expressed in the heart
(Tobimatsu & Fujisawa, 1989; Edman & Schulman, 1994; Srinivasan et
al., 1994; Baltas et al., 1995; Mayer et al., 1995; Ramirez et al., 1997b).
Transgenic studies have provided evidence for the involvement of
CaMKII in pathological cardiac hypertrophy. Over-expression of
CaMKII
B
(nuclear isoform) in the mouse heart induced cardiac
hypertrophy and dilated cardiomyopathy (Zhang et al., 2002b).
Similarly, cardiac-specic transgenic mice over-expressing CaMKII
C
(cytoplasmic isoform) developed cardiac hypertrophy (with mild
brosis), dilated cardiomyopathy and heart failure (lung congestion,
atrial dilation and severe edema) (Zhang et al., 2003b). Both CaMKII
B
and CaMKII
C
transgenic expression induced a genetic program
associated with hypertrophy and heart failure (i.e., ANP, -MHC and
-skeletal actin mRNA were increased, -MHC, SERCA2a and
phospholamban (PLB) were decreased). Further supporting this
role, a genetic mouse model of CaMKII inhibition prevented cardiac
dilation and dysfunction resulting from myocardial infarction and -
AR stimulation (Zhang et al., 2005). Under basal conditions, these
transgenic mice had a stable cardiac phenotype (no hypertrophy),
normal cardiac development and function (Zhang et al., 2005). More
recently, CaMKII-null mice (generated using a Cre-loxP approach)
were shown to be protected against hypertrophy and brosis in
response to pressure overload caused by aortic-constriction (Backs et
al., 2009). Fetal gene expression (ANP, BNP, -MHC) was also
attenuated in this model. In contrast, germline ablation of CaMKII
did not affect the development of hypertrophy following pressure
overload induced by two weeks of aortic constriction, which may be
explained by the upregulation of CaMKII compensating for the loss of
CaMKII (Ling et al., 2009). However, after long term aortic banding
(6 weeks), cardiac function was maintained, enlargement of the heart
was attenuated, pulmonary congestion and lung edema were
reduced, and survival rates improved in the knockout model
compared to wildtype mice, suggesting that CaMKII deletion inhibits
the development of heart failure induced by long term pressure
overload (Ling et al., 2009).
CaMKII signaling is thought to exert its effect oncardiac hypertrophy
by causing phosphorylation of class II histone deacetylase 4 (HDAC4),
which in turn dissociates from the MEF2 transcription factor, and is
translocatedfromthe nucleus to the cytoplasm. This causes activationof
MEF2 which is sufcient to promote pathological hypertrophy (Backs &
Olson, 2006; Backs et al., 2006; Kim et al., 2008a). In further support of
this role, CaMKII-null mice demonstrateda clear reductionof HDAC4in
ventricular lysates (Backs et al., 2009).
3.3. Complexities and future work in relation to signaling cascades
The most frequently used animal models of pathological hyper-
trophy (e.g. aortic-banding, hypertension) represent a chronic
pressure load that results in concentric hypertrophy. In contrast,
models of physiological hypertrophy (e.g. treadmill, voluntary free-
wheel, swimming) represent an intermittent volume load that results
in eccentric hypertrophy. Thus, it has been argued that differences in
phenotypes and signaling observed in models of pathological and
physiological hypertrophy may be a consequence of the duration of
the insult (constant versus intermittent) or type of load (volume
versus pressure). These issues have been addressed to some extent by
the following studies: i) mouse hearts subjected to intermittent
pressure overload displayed pathological features including diastolic
dysfunction, reduced capillary density, histological and cellular
abnormalities, and a fall in SERCA2a (Perrino et al., 2006); ii)
cardiac-specic chronic transgenic expression of IGF1R or caPI3K
was associated with a physiological phenotype that did not progress
to a pathological phenotype (Shioi et al., 2000; McMullen et al.,
2004b); iii) transgenic mice with enhanced ERK5 activation devel-
oped eccentric hypertrophy that progressed to dilated cardiomyop-
athy (Nicol et al., 2001), iv) a pathological model of eccentric
hypertrophy (myocardial infarction) and a physiological model of
eccentric hypertrophy (voluntary exercise wheel) in rats were
associated with differential regulation of signaling proteins (Gosselin
et al., 2006). Together these data suggest that it is not whether a
stimulus is chronic or intermittent, or whether the initiating
hypertrophic stimuli represent a pressure or volume overload that
determines whether the resultant cardiac hypertrophy is pathological
or physiological.
Further studies are required to determine whether postnatal
cardiac growth, pregnancy-induced growth, and exercise-induced
hypertrophy are mediated by similar molecular mechanisms. PI3K
(p110) is critical for postnatal heart growth and exercise-induced
growth (Shioi et al., 2000; McMullen et al., 2003); whereas the Gq
signaling pathway was critical for pathological growth but not
postnatal heart growth (Wettschureck et al., 2001). Signaling
cascades implicated in mediating pregnancy-induced heart growth
have not been extensively studied but may be coupled to Kv4.3, c-Src,
cGMP and estrogen receptors. Consistent with the classication of
pregnancy-induced growth as physiological, pregnancy does not
trigger changes in classic markers of pathological hypertrophy
including -MHC or ANP (Eghbali et al., 2005; Eghbali et al., 2006).
Examination of signaling pathways that specically induce and
differentiate between eccentric or concentric hypertrophy have not
been clearly dened. Patients and animal models with eccentric
hypertrophy have a particularly poor prognosis (Berenji et al., 2005).
An understanding of eccentric hypertrophy at the molecular level is
likely to provide important insight into why compensated hypertro-
phy can progress to heart failure. It will also be important to
differentiate between physiological eccentric hypertrophy and path-
ological eccentric hypertrophy. For instance, hypertrophy in response
to isotonic exercise is classied as eccentric but is quite different to
eccentric hypertrophy associated with decompensation of the heart.
Furthermore, pregnancy-induced hypertrophy due to volume over-
load is not associated with recapitulation of the fetal gene program
whereas a similar volume overload in response to aortocaval shunt
caused an increase in ANP and -MHC gene expression (Sopontam-
marak et al., 2005; Eghbali et al., 2006). Gp130-mediated signals
contribute to the development of eccentric hypertrophy. Both CT-1
and LIF cause elongation of myocytes due to assembly of sarcomeric
units in series rather than in parallel (Wollert et al., 1996). But as
previously noted, activation of this pathway is considered protective.
In contrast, MEK5 induces eccentric hypertrophy and heart failure
with no report of intervening concentric hypertrophy (Nicol et al.,
2001). An intact sensing apparatus that can detect changes in
functional load appears essential in the compensatory response to
pathological insults. Mutations or deletions of sensor proteins have
been associated with dilated cardiomyopathy/eccentric hypertrophy
characterized by ventricular chamber dilation, brosis and heart
failure. Integrins (heterodimeric transmembrane receptors), link the
extracellular matrix to the intracellular cytoskeleton. Interacting
proteins/downstream effectors of integrins (e.g. melusin, focal
adhesion kinase (FAK), small GTPases) as well as proteins at the
level of the Z-disc within sarcomeres (e.g. muscle LIM protein) are
considered biomechanical stretch sensors. Control mice typically
develop concentric hypertrophy in response to pressure overload
(aortic banding) that may later progress to eccentric hypertrophy and
heart failure. In contrast, mice lacking stretch-sensor proteins (e.g. -
206 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
integrin, melusin, FAK, and muscle LIM protein) developed eccentric
hypertrophy and cardiac dysfunction under basal conditions or in
response to pressure overload, i.e. no intermediate concentric
hypertrophy (Arber et al., 1997; Shai et al., 2002; Brancaccio et al.,
2003; Peng et al., 2006). On the other hand, melusin expression is
increased during compensated hypertrophy induced by pressure
overload, and over-expression of melusin in transgenic mice
prolonged concentric hypertrophy and protected against the transi-
tion to eccentric hypertrophy and failure (De Acetis et al., 2005).
4. Molecular mechanisms associated with structural
features of pathological and physiological hypertrophy
4.1. Molecular mechanisms associated with protein synthesis
An essential feature of both physiological and pathological cardiac
hypertrophy is increasedproteinsynthesis. Ribosomal S6 kinases (S6Ks:
S6K1 and S6K2) are consideredcritical regulators of protein synthesis in
response to hypertrophic stimuli. Activation of S6K1 was increased in
hearts of transgenic mouse models of physiological hypertrophy (mice
with increased activation of the IGF1-PI3K(p110) pathway) and
pathological hypertrophy (pressure overload) (Shioi et al., 2000; Shioi
et al., 2003; McMullen et al., 2004a,b). Rapamycin, an inhibitor of the
mammalian target of rapamycin (mTOR; a proximal effector of S6Ks)
inhibited S6K1 and attenuated and regressed pathological hypertrophy
induced by pressure overload (Shioi et al., 2003; McMullen et al.,
2004a). Unexpectedly, deletion of S6Ks (utilizing global S6K knockout
mice) did not attenuate cardiac hypertrophy induced by exercise
training, transgenic expression of IGF1R/caPI3K or pressure overload
(McMullen et al., 2004c). Together, these data suggest that S6Ks are not
essential for the induction of physiological or pathological cardiac
hypertrophy, or that other proteins were upregulated in the knockout
models.
4.2. Molecular mechanisms associated with sarcomeric organization
The formation of sarcomeres is a complex process involving the
synthesis of numerous proteins (Zak, 1984; Vigoreaux, 1994). These
proteins aggregate into laments, which are organized into specic
three-dimensional arrays and aligned with other contractile elements
already present in the cardiac myocyte (Zak, 1984; Sanger et al., 2000).
Cardiac hypertrophy facilitates the increased workload (e.g. strain of
pressure or volume overload) by increasing contractile capacity.
Organization of sarcomeres, and thereby an increase in the contractile
units, is an essential component to maximize force generation. The
transition from pathological hypertrophy to heart failure has been
associated with loss of contractile laments in the presence of
microtubule densication and desmin disorganization, and loss of
proteins of the sarcomeric skeleton (titin, -actin, myomesin) (Tsutsui
et al., 1993; Heinet al., 2000; Hamdani et al., 2008). Defects or mutations
of sarcomeric proteins, including cardiac troponin I or T, -MHC, -
MHC, myosin light chain (MLC), -tropomyosin, titin, and actin, have
been associated with familial hypertrophic cardiomyopathy in humans
(Margulies & Houser, 2004; Lind et al., 2006; Morimoto, 2008;
Tsoutsman et al., 2008). Signaling proteins and regulators of gene
expressionthat have beenimplicated insarcomere organizationinclude
Rac1, RhoA(Aoki et al., 1998; Hoshijima et al., 1998; Pracyk et al., 1998),
FAK and p130Cas (substrates for the non-receptor tyrosine kinase Src)
(Kovacic-Milivojevic et al., 2001), multiple kinases (see Seguchi et al.,
2007; Solaro 2008), Nkx2.5/Csx (Kasahara et al., 2003), and HDAC4
(Gupta et al., 2008).
4.3. Molecular mechanisms associated with brosis
Cardiac brosis is a common feature in animal models of
pathological hypertrophy and patients with advanced heart failure.
In response to a pathological insult, cardiac broblasts and extracel-
lular matrix proteins accumulate disproportionately and excessively.
This leads to mechanical stiffness which contributes to diastolic
dysfunction and can progress to systolic dysfunction (Weber & Brilla,
1991; Villarreal & Dillmann, 1992; Weber et al., 1993; Brower et al.,
2006). It is recognized that anti-brotic therapies may be useful in
improving cardiac function of the diseased heart. However, the
development of such therapies has been limited by an incomplete
understanding of the source of broblasts and mechanisms respon-
sible for the induction of excessive collagen accumulation. Tradition-
ally, adult broblasts are considered to be derived directly from
embryonic mesenchymal stem cells and to increase in number in a
setting of pathological stress due to proliferation of resident
broblasts (Weber & Brilla, 1991; Maric et al., 1997; Weber, 1997;
Lang and Fekete 2001). More recently, it has been demonstrated that
bone marrow derived broblasts, as well as endothelial cell derived
broblasts (via endothelialmesenchymal transition) contribute to
the population of cardiac broblasts (Zeisberg et al., 2007).
Genetic mouse models have provided some insight in relation to
the molecular mechanisms responsible of brosis associated with
pathological hypertrophy. Triggers, receptors, signaling proteins, and
transcription factors implicated in the development of cardiac brosis
include: Ang II, AT
1
receptor, TGF-, tumor necrosis factor (TNF)-, G
proteins (G
q
, G
s
), PKC2, RhoA, calcineurin, calsequestrin, NFATs,
Csx/Nkx2.5, and serum response factor (see Manabe et al., 2002;
Gunasinghe & Spinale, 2004; Rosenkranz, 2004).
TGF-1 is probably the best described mediator of cardiac brosis. In
the heart, TGF-1 is secretedby cardiac broblasts inresponse tostimuli
suchas Ang II (Khan&Sheppard, 2006). The effects of TGF-1 oncardiac
brosis appear to be mediated in part via Smads and TGF--activated
kinase-1 (TAK1) leading to increased transcription of extracellular
matrix proteins (Rosenkranz, 2004; Khan & Sheppard, 2006). Trans-
genic mice over-expressing TGF-1 developed cardiac hypertrophy that
was associated with interstitial brosis and increased myocyte size
(Rosenkranz et al., 2002). Incontrast, heterozygous TGF-1
+/
decient
mice displayed less cardiac brosis with aging (Brooks &Conrad, 2000).
TGF-1 was also responsible for endothelial cells undergoing endothe-
lialmesenchymal transition and the induction of increased brosis in a
setting of pathological hypertrophy induced by pressure overload
(Zeisberg et al., 2007). Furthermore, inhibition of TGF-1 with
recombinant bone morphogenic protein-7 (member of the TGF-
superfamily of growth factors) was able to inhibit endothelial
mesenchymal transition and brosis in mice with pressure overload-
induced hypertrophy (Zeisberg et al., 2007).
Myocyte death by apoptosis or necrosis is typically observed in
parallel with the onset of brosis. Debate remains as to whether
brosis is simply a non-specic response to myocyte loss or whether
the release of secretory factors from myocytes promote or inhibit
collagen synthesis also plays a signicant role (Benjamin et al., 1989).
It was shown that rat cardiac myocytes induce secretion of active TGF-
in the presence of Ang II and that a paracrine action of TGF-
induced cytokines in broblasts to promote collagen synthesis (Sarkar
et al., 2004).
4.4. Molecular mechanisms associated with cell death
Pathological hypertrophy is typically associated with increased cell
death (apoptosis and necrosis) whereas physiological hypertrophy is
not. Thus, differential activation of pro-survival and pro-death signals
is likely to be an important factor that contributes to the distinct
phenotypes of these two forms of heart growth. Low levels of
apoptosis were sufcient to induce heart failure in mice, and
inhibition of cell death with a polycaspase inhibitor largely prevented
the heart failure phenotype (Wencker et al., 2003). In another study,
inhibition of apoptosis with a caspase inhibitor improved cardiac
207 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
function and prevented mortality in pregnant cardiac-specic Gq
transgenic mice (Hayakawa et al., 2003).
Growth factors (e.g. IGF1), and cytokines (e.g. CT-1) have anti-
apoptotic effects, in part, via PI3K and/or ERK signaling (Parrizas et al.,
1997; Sheng et al., 1997; Haunstetter and Izumo 1998; Kang et al.,
2002). In contrast, signaling via G
q
or G
s
in transgenic mice was
shown to promote cardiac myocyte apoptosis (Adams et al., 1998;
Geng et al., 1999). Calcineurin has been reported to have both pro-
and anti-apoptotic effects (De Windt et al., 2000; Saito et al., 2000; Pu
et al., 2003). JNK, p38-MAPK, and loss of gp130 have been associated
with increased apoptosis in cardiac myocytes (Wang et al., 1998a;
Hirota et al., 1999; Kang et al., 2002).
Rats with pathological hypertrophy (Dahl salt sensitive rats on a
high salt diet) were more sensitive to apoptotic stimulation than
cardiac myocytes from rats with exercise-induced physiological
hypertrophy (treadmill). The physiological hypertrophic model was
associated with changes in Bcl-2 family members and caspases
favoring survival, whereas the pathological model was associated
with changes in mitochondrion- and death receptor-mediated path-
ways (e.g. a decreased Bcl-x
L
/Bax ratio, increase in Fas) that have
previously been associated with pathological hypertrophy and heart
failure (Kang et al., 2004).
4.4.1. Autophagy
Autophagy is the process by which cells degrade and recycle aged
proteins and damaged organelles. Inhibition of autophagy triggers
apoptosis, indicating that autophagy plays a key role in cell survival
(Boya et al., 2005). Autophagy is upregulated in numerous models of
pathological hypertrophy and in failing hearts (see Nishida et al.,
2008; Nishida et al., 2009). This may be a protective mechanism,
preventing the accumulation of cytoplasmic components that disrupt
cardiac function (Nishida et al., 2009). Little is known about the role of
autophagy in settings of physiological cardiac hypertrophy, however
autophagy is upregulated by exercise and appears to be important for
amino acid turnover and protein synthesis in skeletal muscle post
exercise (see Gottlieb et al., 2009).
Fig. 8. Alterations in substrate utilization in pathological and physiological cardiac
hypertrophy. Pathological hypertrophy is associated with a switch from fatty acid to
glucose utilization, although glucose metabolismalso decreases with the progression to
heart failure. In contrast, physiological hypertrophy is associated with enhanced rates
of fatty acid and glucose oxidation.
Fig. 9. Overview of energy metabolism in cardiac myocytes. Circulating fatty acids, glucose and lactate are the main fuel sources utilized by cardiac myocytes to generate ATP.
Following uptake into cardiac myocytes, fatty acids are transported to the mitochondria by carnitine palmitoyltransferase I (CPTI) and II (CPTII) for -oxidation, entry into the
tricarboxylic acid (TCA) cycle and oxidative phosphorylation. Glucose and lactate are converted to pyruvate before entering the mitochondria and converging with fatty acid
oxidation pathways. Asterisks denote proteins and processes that can induce cardiac hypertrophy when altered or defective. ATP: adenosine triphosphate, CoA: coenzyme A, HIF1:
hypoxia inducible factor 1, NADPH: nicotinamide adenine dinucleotide phosphate, PPAR: peroxisome proliferator-activated receptor, ROS: reactive oxygen species.
208 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
5. Molecular mechanisms associated with differences in
energy metabolism in pathological and physiological hypertrophy
As described earlier, pathological and physiological hypertrophy
are associated with distinct metabolic proles. Substrate utilization in
pathological hypertrophy resembles that of the fetal heart (decreased
fatty acid oxidation, increased glycolysis), while physiological
hypertrophy induced by exercise training is associated with enhanced
fatty acid and glucose oxidation (see Lehman & Kelly, 2002; Fig. 8).
Whether alterations in energy metabolismare a cause or consequence
of cardiac hypertrophy and failure is controversial, however there is a
growing body of evidence to suggest that switches in substrate
utilization and other alterations in energy metabolism do contribute
to the development of pathological hypertrophy and failure. The
accumulation of lipids in cardiac myocytes is deleterious and can lead
to pathological hypertrophy and heart failure, although the mechan-
isms responsible for lipotoxic cardiomyopathy are still unknown.
Reactive oxygen species (ROS), a by-product of mitochondrial energy
metabolism, may also induce hypertrophy in pathological settings
(Fig. 9).
5.1. Peroxisome proliferator-activated receptors (PPARs)
The PPAR family of transcription factors are key metabolic
regulators that have been implicated in the development of cardiac
hypertrophy (see Robinson & Grieve, 2009 for review). PPAR is a
nuclear receptor that regulates lipid metabolism by increasing
transcription of genes involved in fatty acid oxidation, such as
muscle-type carnitine palmitoyltransferase 1 (mCPT-1) and medium
chain acyl coenzyme A dehydrogenase (MCAD) (Fig. 9) (see Barger &
Kelly, 2000). PPAR expression was elevated in hearts of trained rats
(Rimbaud et al., 2009) and downregulated in several models of
pathological hypertrophy (see Barger et al., 2000; Lehman & Kelly,
2002; Akki et al., 2008) as well as in the fetal heart (Sack et al., 1997).
Transgenic mice decient for PPAR developed greater hypertrophy
and had worse cardiac function compared to wildtype mice when
subjected to pressure overload (Smeets et al., 2008). In addition,
administration of a PPAR agonist attenuated hypertrophy induced
by aldosterone treatment in mice (Lebrasseur et al., 2007) and in two
rat models of pathological hypertrophy (Ichihara et al., 2006; Linz et
al., 2009). These studies suggest that downregulation of PPAR (and
therefore fatty acid oxidation) contributes to the development of
pathological hypertrophy, although a study in cardiac-specic
transgenic mice has demonstrated that enhanced PPAR protein
expression can also induce hypertrophy and cardiac dysfunction
(Finck et al., 2003).
PPAR is a target of ERK1/2 (Barger et al., 2000), a signaling
protein which is activated in settings of pathological cardiac
hypertrophy (see Section 3.2.3.1). Deactivation of PPAR by ERK1/2
was responsible for the reduction in fatty acid oxidation observed in
cardiac myocytes stimulated with the hypertrophic agonist PE (Barger
et al., 2000), indicating that signaling pathways involved in the
development of cardiac hypertrophy can also exert changes in cellular
metabolism.
The role of PPAR in the development of cardiac hypertrophy is
less clear. Several studies suggest that PPAR is protective, as mice
lacking PPAR in the heart developed cardiac hypertrophy and
dysfunction (Ding et al., 2007), and treatment with a PPAR agonist
reduced cardiac remodeling and brosis in a rat model of hyperten-
sion (Henderson et al., 2007). However, cardiac-specic over-
expression of PPAR in mice also resulted in cardiac dysfunction
(Son et al., 2007), indicating that PPAR activation is not always
protective (Fig. 9). A recent study demonstrated that transgenic mice
with enhanced PPAR activity (due to elevated levels of ventricular
hypoxia-inducible factor 1, HIF1) developed concentric hypertro-
phy which progressed to dilated cardiomyopathy at around 5
6 months of age (Fig. 9) (Krishnan et al., 2009). This was attributed
to PPAR-dependent lipid accumulation and cardiac myocyte apo-
ptosis. HIF1 is a key regulator of genes involved in glycolytic
metabolism and was found to be highly expressed in ventricular
biopsies frompatients with hypertrophic cardiomyopathy. HIF1was
also highly expressed in a mouse model of pathological hypertrophy
induced by pressure overload, but not in physiological hypertrophy
induced by exercise training. This study provides strong evidence that
a switch fromfatty acid to glycolytic metabolism leads to pathological
hypertrophy and is ultimately detrimental for cardiac function.
5.2. Lipid accumulation and pathological hypertrophy
Lipid accumulation appears to act as a trigger for cardiac myocyte
hypertrophy (Chiu et al., 2001) and is evident in the failing human
heart (Sharma et al., 2004), however mechanisms responsible for
lipotoxic cardiomyopathy are unclear (see McGavock et al., 2006 and
Park et al., 2007 for reviews). Lipid accumulation can result from
alterations in fatty acid uptake, transport or oxidation. For example,
carnitine is responsible for transporting long-chain fatty acids from
the cytosol into the mitochondria for subsequent -oxidation (Fig. 9).
Carnitine deciency impairs fatty acid oxidation and has been
associated with the development of left ventricular hypertrophy and
cardiomyopathy in humans (Koizumi et al., 1999). Mice with systemic
carnitine deciency (juvenile visceral steatosis (JVS) mice) displayed
signicant cardiac hypertrophy at eight weeks of age (Horiuchi et al.,
1993; Kuwajima et al., 1998) and were more susceptible to pressure
overload (Takahashi et al., 2007). Treatment with carnitine attenu-
ated the basal hypertrophic phenotype in JVS mice (Horiuchi et al.,
1993), and improved cardiac function in JVS mice subjected to
pressure overload (Takahashi et al., 2007). Furthermore, limiting
dietary lipid intake attenuated hypertrophy in JVS mice and this was
associated with reduced triglyceride accumulation in the ventricles
(Jalil et al., 2006).
Enhanced fatty acid uptake (e.g. due to excess dietary lipids, as
occurs in obesity) has also been linked with the development of
pathological hypertrophy (Kankaanpaa et al., 2006; Koonen et al.,
2007) and may play a role in the development of diabetic
cardiomyopathy (see Carley & Severson, 2005), although high-fat
feeding attenuated hypertrophy in rats with hypertension (Okere et
al., 2005).
5.3. Reactive oxygen species
(ROS)/oxidative stress and pathological hypertrophy
ROS are a natural by-product of mitochondrial energy production
(Fig. 9). Oxidative stress occurs when ROS production outweighs the
antioxidant capabilities of the cell, and has been implicated in the
pathogenesis of cardiac hypertrophy and heart failure (McMurray
et al., 1993). Hypertrophic stimuli, such as Ang II, ET-1 and
catecholamines, are capable of stimulating ROS production in cardiac
myocytes (Liu et al., 2004; Laskowski et al., 2006). Defective NADPH
oxidase activity is another source of ROS that has been implicated in
the development of pathological hypertrophy (see Murdoch et al.,
2006). A study in transgenic mice demonstrated that NADPH oxidase
plays an important role in the development of cardiac hypertrophy, as
mice decient for the gp91
phox
subunit of NADPH oxidase were
protected from developing pathological hypertrophy induced by
chronic infusion of Ang II (Bendall et al., 2002).
ROS production also increases during exercise although this
appears to have a preconditioning effect, reducing susceptibility to
oxidative stress-related disorders (see Radak et al., 2008). Exercise-
induced ROS production is much lower in the heart compared with
liver and skeletal muscle (Traverse et al., 2006). Thus it seems unlikely
that ROS play a role in the development of physiological hypertrophy.
209 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
5.4. Changes in energy metabolism
associated with physiological cardiac hypertrophy
There are clear differences in energy metabolism between
pathological and physiological hypertrophy. While pathological
hypertrophy is typically associated with a switch from fatty acid to
glucose utilization, physiological hypertrophy is accompanied by
enhanced rates of fatty acid and glucose oxidation. This has been
demonstrated in numerous animal models, as well as in humans. For
example, the expression of genes encoding enzymes involved in fatty
acid oxidation (i.e. PPAR, mCPT-1, and MCAD) was signicantly
increased in the hearts of trained rats but not hypertensive rats,
despite a similar degree of hypertrophy (Rimbaud et al., 2009). In
another study, phosphofructokinase and lactate dehydrogenase, key
regulators of glycolytic metabolism, were upregulated in pathological
hypertrophy induced by hypertension but not in physiological
hypertrophy induced by exercise training (Iemitsu et al., 2003). As
mentioned previously, HIF1 (a key regulator of glycolytic metabo-
lism) was elevated in mice and human patients with pathological
hypertrophy, but not in a mouse model of physiological hypertrophy
(Fig. 9) (Krishnan et al., 2009). Rats with exercise-induced hypertro-
phy had lower rates of glycolysis compared to sedentary controls,
while glucose and palmitate (i.e. fatty acid) oxidation increased by
45% and 5065%, respectively (Burelle et al., 2004). The enhanced
oxidative capacity of hearts that have undergone physiological
hypertrophy suggests a protected phenotype.
6. Characteristic gene expression changes
associated with pathological and physiological hypertrophy
Intracellular signaling pathways are coupled with transcription
factors in the nucleus to regulate the long-term alterations in gene
expression that are associated with cardiac hypertrophy. Classically,
pathological hypertrophy has been associated with upregulation of
fetal genes including ANP, BNP, -skeletal actin, atrial MLC-1, and -
MHC; and downregulation of genes normally expressed at higher
levels in the adult than in the fetal ventricle, such as -MHC and
SERCA2a (Izumo et al., 1988; Chien et al., 1991; MacLellan &
Schneider, 2000). MHC is the major component of myosin, the
protein complex responsible for driving contraction in muscle cells. In
rodents, -MHC is the predominant isoform present in the prenatal
heart, but is down-regulated soon after birth when -MHC is
expressed (Morkin, 2000). During pathological hypertrophy the
increased expression of -MHC and fall in -MHC may represent an
adaptive response, as -MHC is slower than -MHC at catalyzing the
hydrolysis of ATP (the chemical reaction driving myocyte contraction)
leading to slower, more economical, contractile function (Swynghe-
dauw, 1986; Izumo et al., 1987; Dorn et al., 1994; Swynghedauw,
1999). Recapitulation of the fetal gene program and switch in
contractile protein composition does not commonly occur in models
of physiological hypertrophy e.g. exercise-induced hypertrophy
(McMullen et al., 2003) (Fig. 4). The direct impact of changes in
fetal gene expression on cardiac growth, function and brosis remain
unclear. While fetal genes are often upregulated in models of
pathological hypertrophy, this may be a compensatory response to
protect the heart. ANP and BNP signaling have antihypertrophic
actions within cardiac myocytes (Woods, 2004). Furthermore, some
transgenic models of physiological hypertrophy have been associated
with a modest upregulation of fetal genes (e.g. IGF1R (McMullen et al.,
2004b) and MEK1 (Bueno et al., 2000)), whereas a model of
intermittent pressure overload was associated with a pathological
phenotype but no upregulation of the fetal gene program (Perrino et
al., 2006). Finally, some studies have demonstrated that a small/
normal heart size phenotype can be associated with activation of the
fetal gene program (Shioi et al., 2000; Antos et al., 2002). Thus, until
the biological signicance of changes in the fetal gene program is
better understood models of cardiac hypertrophy are probably best
dened based on functional and structural parameters.
Transcription factors including GATA4, GATA6, Csx/Nkx2.5, MEF2,
c-jun, c-fos, c-myc, nuclear factor-B, and NFAT have been implicated
for the activation of cardiac genes in response to hypertrophic stimuli
(Sadoshima & Izumo, 1997; Aoki & Izumo, 2001; Akazawa & Komuro,
2003). Of note, Gata4 has been implicated in the regulation of a
number of genes associated with pathological hypertrophy (e.g. ANP,
BNP, -skeletal actin, and -MHC) but may also be important for
physiological growth. A recent study demonstrated that mice with
reduced GATA4 had mild cardiac dysfunction and reduced heart
weight under basal conditions. In response to pressure overload,
GATA4 decient mice developed eccentric hypertrophy and heart
failure associated with marked apoptosis and brosis (Bisping et al.,
2006).
NGF1A-binding protein (Nab1) is a transcriptional repressor for
early growth response transcription factors. Nab1 appears to be a
specic regulator of pathological cardiac hypertrophy. Cardiac-
specic Nab1 transgenic mice displayed a blunted hypertrophic
response to adrenergically-induced and pressure overload-induced
hypertrophy whereas physiological growth (development or exer-
cise-induced hypertrophy) was not affected (Buitrago et al., 2005).
Proling studies are beginning to provide a more global view of
differences in gene expression in various hypertrophic models (see
Section 6.2).
6.1. Chromatin-modifying enzymes
Histone-dependent packaging of genomic DNA into chromatin is a
central mechanism for gene regulation. Nucleosomes (the basic unit
of chromatin) interact to create a highly compact structure that limits
access of genomic DNA to transcription factors, thus repressing gene
expression (McKinsey et al., 2002). Chromatin-remodeling enzymes
have been implicated in the re-expression of the fetal gene program
(McKinsey et al., 2002). The chromatin-modifying enzymes, HDACs,
promote chromatin condensation and thus repress transcription
(McKinsey et al., 2002).
HDACs have been grouped into three classes (IIII). Class IIa
HDACs (4, 5, 7 and 9) which are regulated by signaling proteins
including PKC and CaMK have been implicated as suppressors of
pathological hypertrophy (Bush & McKinsey, 2009). HDAC5 and
HDAC9 knockout mice developed hypertrophy with increased age and
showed an exaggerated hypertrophic response to pressure overload
(aortic constriction), however postnatal heart growth was unaffected
in the knockout models (Zhang et al., 2002a; Chang et al., 2004). By
Fig. 10. Venn diagram showing the shared and distinct gene clusters (as determined
from microarray analysis) that have been shown to be associated with pathological and
physiological hypertrophy. HSF1 is a heat shock protein transcription factor that has
recently been shown to be involved in physiological hypertrophy. Data collected from
(Kong et al., 2005; Mirotsou et al., 2006; Dorn 2007).
210 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
contrast, class I HDACs (1 and 2) are reported to promote
hypertrophy. Cardiac-specic HDAC2 transgenic mice developed
cardiac hypertrophy associated with increased fetal gene expression
(Trivedi et al., 2007). Class III HDACs (also known as sirtuins) are also
reported to inhibit pathological hypertrophy, including apoptosis (see
Bush & McKinsey, 2009).
6.2. Proling studies of cardiac hypertrophy
In another approach to identify transcriptome changes associated
with pathological and physiological cardiac hypertrophy, a number of
investigators have conducted comprehensive microarray gene ex-
pression proling studies in hearts from rodent models of patholog-
ical and physiological cardiac hypertrophy, as well as tissue from
heart failure patients (Friddle et al., 2000; Hwang et al., 2000; Yang et
al., 2000; Barrans et al., 2001; Hwang et al., 2002b; Diffee et al., 2003;
McMullen et al., 2004b; Kong et al., 2005; Strom et al., 2005). More
recent studies have used proteomic methods, such as 2-dimensional
polyacrylamide gel electrophoresis (2-D PAGE) in conjunction with
mass spectrometry to catalogue exercise-induced changes in the
cardiac proteome (Boluyt et al., 2006; Burniston, 2009).
6.2.1. Gene expression proling (microarray)
Microarray technology is a powerful technique and quite often the
method of choice to study mRNA expression differences between a
cohort of samples, different transgenic mouse models, or tissues
subjected to different conditions. This technology has been used to
identify molecular differences between physiological and pathological
hypertrophy in a number of transgenic rodent models (Friddle et al.,
2000; McMullen et al., 2004b; Kong et al., 2005; Stromet al., 2005), as
well as in human tissue samples of non-failing hearts versus failing
hearts (Hwang et al., 2000; Yang et al., 2000). The outcomes of these
studies have been reviewed in depth (Barrans et al., 2001; Hwang et
al., 2002a; Dorn 2007). In brief, these studies have identied shared
and distinct gene cluster expression proles of physiological and
pathological hypertrophy (Fig. 10). Genes associated with patholog-
ical hypertrophy were largely from inammation, apoptotic, cardiac
fetal gene, and oxidative stress clusters. On the other hand, genes that
were associated with physiological hypertrophy were predominately
involved in cell survival, fatty acid oxidation, insulin signaling,
epidermal growth factor signaling and HSF1 expression (Fig. 10)
(Aronowet al., 2001; McMullen et al., 2004b; Kong et al., 2005; Strom
et al., 2005; Mirotsou et al., 2006). Cardiac apoptosis has been shown
to be involved in various human and animal models of heart failure
(Kang & Izumo, 2000), including the G
q
transgenic mouse which
develops hypertrophy and heart failure (Aronow et al., 2001).
Expression of genes that are involved with fatty acid oxidation in
physiological models are consistent with the known increase in fatty
acid oxidation in response to exercise (see Section 5.4). These studies,
in addition to the data obtained from genetic mouse models, lend
further support to the notion that distinct molecular mechanisms can
regulate pathological and physiological cardiac hypertrophy. Mem-
bers of the IGF1/EGF signaling pathway showed statistically signi-
cant changes in a rat model of physiological hypertrophy, in
comparison to compensated pathological hypertrophy (Dahl salt
sensitive rats on a high salt diet) (Kong et al., 2005). Another
application of microarray technology is the discovery of new drug
targets. Identication of novel genes that play important roles in
mediating physiological hypertrophy and protection may open up
new avenues for treating patients with heart failure (discussed
further in Section 8). In general, research and therapy has concen-
trated on identifying and inhibiting pathological processes. Another
therapeutic strategy may be to activate novel regulators of physio-
logical hypertrophy that have been identied from gene expression
proling studies (McMullen et al., 2004b; Kong et al., 2005; Lin et al.,
2010).
Gene expression proling has also been utilized to examine
differences between pathological eccentric and concentric hypertro-
phy. Rat models of concentric hypertrophy (aortic constriction for
pressure overload) and eccentric hypertrophy (aortocaval shunt for
volume overload) with similar degrees of cardiac hypertrophy were
associated with different gene expression patterns (Miyazaki et al.,
2006). Sixty-four genes behaved similarly between the 2 models but
93 genes were altered only in the pressure overload model and 134
genes were differentially expressed in the volume overload model. In
another study, comparisons were made between a model of
pathological hypertrophy (pressure overload by ascending aortic
banding for 1 week) and physiological hypertrophy (chronic swim
training for 4 weeks) in mice. Only a small percentage (approximately
3%) of genes were regulated in a similar manner between the 2
models (Cardiogenomics, 20012003).
6.2.2. Proteomics
Investigators have also used proteomic techniques to identify
changes in expression, splice variation and post translational
modications in models of cardiac hypertrophy (Boluyt et al., 2006;
Burniston, 2009). Treadmill exercise-trained rats showed cardiac
hypertrophy, with a 1418% increase in heart weight to body weight
ratio compared to sedentary controls (Boluyt et al., 2006). Analysis of
2-dimensional electrophoresis gels revealed protein spots that were
decreased, increased, or detected exclusively in exercise trained
hearts compared to controls. Heat shock protein 20 (Hsp20)
represented a protein exclusively expressed in exercise trained hearts
(identied by mass spectrometry and immunoblotting). A later study
corroborated this nding in another rat model, but in a setting of
moderate exercise that is similar to physical activity guidelines for
humans (Burniston, 2009).
Hsp20 belongs to the subfamily of small heat shock proteins in
which there are 10 known members in mammalian species (Fan et al.,
2005a). Hsp20 contains three phosphorylation sites, serine 16, serine
59, and serine 157, but conclusions on the role of Hsp20 have mostly
been based on serine 16 phosphorylation (Fan et al., 2005a). Heat
shock proteins have been involved in the preservation of myocardial
function after ischemia/reperfusion (Kingma, 1999). Studies have
shown that Hsp20 enhances myocardial contraction in vitro (Chu et
al., 2004), and cardiac-specic transgenic mice over-expressing
Hsp20 had enhanced cardiac function in the absence of pathological
abnormalities (Fan et al., 2005c). Furthermore, following coronary
artery occlusion and reperfusion (24 h), Hsp20 transgenic hearts
exhibited better cardiac functional recovery, and reduced infarct area
compared to wildtype hearts (Fan et al., 2005c). Finally, Hsp20
transgenic mice displayed an attenuated hypertrophic response to -
AR agonist-induced cardiac hypertrophy, while retaining enhanced
cardiac function. Transgenic expression also prevented the -AR
agonist-induced increase in fetal genes (ANP, BNP), brosis, and
reduced apoptosis (Fan et al., 2006). In vitro experiments from this
study suggest that Hsp20 is able to inhibit apoptosis signal regulating
kinase 1, which in turn leads to protection from-AR agonist-induced
cardiac hypertrophy/remodeling (Fan et al., 2006). Together, this
suggests that proteins that play unique roles in mediating exercise-
induced physiological hypertrophy may be attractive targets for the
treatment of heart disease.
6.3. microRNAs role in cardiac hypertrophy
MicroRNAs (miRNAs) are a relatively recently discovered family of
small, endogenous, single-stranded RNAs that are approximately 20
25 nucleotides in length. miRNAs have emerged rapidly as a major
new direction in several different elds of research. miRNAs are
partially complementary to their mRNA targets and have an
important role in the regulation of target genes by hybridizing to 3
untranslated regions of messenger transcripts to repress their
211 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
translation or regulate degradation (Bartel, 2004; Grifths-Jones et al.,
2006). An average miRNA is estimated to affect expression of
hundreds of mRNA targets, and up to 30% of human protein coding
genes may be regulated by miRNAs (Lewis et al., 2005; Rajewsky
2006).
Although miRNAs were rst described in 1993 by Victor Ambrose
and colleagues (Lee et al., 1993), since then, 5001000 different
miRNAs have been identied, however, an understanding of their
diverse biological functions remains in its infancy. It is known that
miRNAs play a role in a number of biological processes including
development, cell proliferation, differentiation, and apoptosis
(reviewed in Sassen et al., 2008; Cordes & Srivastava, 2009).
Furthermore, changes in miRNA levels have been correlated with
disease processes, which has sparked numerous academic research
and commercial interest, particularly for cancer (reviewed by Sassen
et al., 2008), viral diseases (Jopling et al., 2005; Lecellier et al., 2005),
and more recently kidney disease (Kato et al., 2009; Liang et al., 2009).
A number of studies have demonstrated the involvement of several
miRNAs in cardiac development (reviewed in (Thum et al., 2008a;
Catalucci et al., 2009; Cordes & Srivastava, 2009)) and in cardiac
pathology (van Rooij et al., 2006; Ikeda et al., 2007; Tatsuguchi et al.,
2007; van Rooij & Olson, 2007; Divakaran & Mann, 2008; Thum et al.,
2008b; Rao et al., 2009; Ren et al., 2009; Tang et al., 2009). More
recently, cardiac specic over-expression of miRNA-208a was shown
to cause pathological cardiac hypertrophy, as shown by thickening of
ventricular walls, depressed cardiac function and increased expres-
sion of -MHC (Callis et al., 2009). In contrast, miRNA biology of
physiological cardiac hypertrophy has received little attention.
miRNA-1 and miRNA-133 were decreased in two models of
physiological cardiac hypertrophy (exercised trained rats and cardiac
specic Akt transgenic mice), however, these miRNAs were also
decreased in a model of pathological hypertrophy (pressure over-
load), and in patients with heart disease (Care et al., 2007). Thus, it is
recognized that further studies outlining similarities and differences
between the regulation of miRNAs in physiological and pathological
hypertrophy are needed (Latronico et al., 2008). To address this gap in
miRNA biology, we recently identied miRNAs that are differentially
regulated in a setting of physiological hypertrophy and cardiac
protection versus a model of cardiac stress associated with patholog-
ical growth (Lin et al., 2010). We rst selected miRNAs that were
differentially expressed in the dnPI3K and caPI3K transgenic mouse
models. We then selected those miRNAs that were differentially
regulated in a setting of physiological hypertrophy and cardiac
protection (caPI3K model) and cardiac stress (myocardial infarction
model).
Silencing of miRNAs in vivo with antagomiRNAs (chemically
modied, cholesterol-conjugated single-stranded RNA analogues
complementary to miRNAs) is considered a powerful approach that
may represent a new therapeutic strategy for targeting cardiac
disease. These molecules have been shown to be effective in vivo,
capable of passing through cellular membranes to inhibit miRNA
action by sequestering it from its targets (Krutzfeldt et al., 2005).
7. Gender differences in cardiac hypertrophy
An area of research that has received increasing attention is that of
differential hypertrophic responses associated with gender (Du et al.,
2006). It is recognized that women typically develop heart disease
later than men. One feature that is considered to help explain the
better prognosis in females than males, is the smaller degree of
cardiac hypertrophy and/or type (concentric versus eccentric) in
females in response to cardiovascular complications such as pressure
overload (demonstrated in animal and human studies) (Du et al.,
2006).
Prior to adolescence, there are no signicant differences in heart
size between males and females, suggesting a similar number of
cardiac myocytes at birth, as myocytes are terminally differentiated
(Zak, 1974; de Simone et al., 1995; Sugden & Clerk, 1998; Luczak &
Leinwand, 2009). Following puberty however, males have 1530%
larger hearts than females, suggesting a signicantly larger degree of
hypertrophy in males (de Simone et al., 1995). Men also lose
approximately 1 g of cardiac mass per year following puberty,
which leads to compensatory hypertrophy to maintain adequate
cardiac mass. Females, however, appear to maintain their myocyte
number and size with aging (Grandi et al., 1992; Olivetti et al., 1995;
Luczak & Leinwand, 2009).
7.1. Pathological hypertrophy
From animal studies there are clear differences in the degree and/
or type of pathological hypertrophy in response to cardiac insults
between males and females (Du et al., 2006; Podesser et al., 2007)
(Table 4). Male spontaneously hypertensive rats developed more
cardiac hypertrophy and left ventricular dysfunction than females.
Subsequent heart failure also occurred earlier in males than females
(Tamura et al., 1999). Similar ndings were reported in response to
pressure overload (aortic-banding) (Douglas et al., 1998; Weinberg et
al., 1999; Skavdahl et al., 2005). Gender differences have also been
reported in humans, though the data are generally less conclusive
owing to confounding factors (e.g. treatments, lifestyle), relatively
small sample sizes, and the limited studies in patients with pure
pressure overload (i.e. in the absence of coronary heart disease).
Furthermore, in human studies there is the question of the best
normalization of left ventricular mass data (body surface area, body
mass index, height
2.7
etc). Animal data is tighter because left
ventricular mass can be accurately assessed at autopsy (rather than
relying on echocardiography). However, in spite of these limitations,
gender differences in cardiac hypertrophy in response to similar
pressure loads (in the absence of coronary heart disease) are
apparent. In normotensive men and women (age matched b69),
normalized left ventricular (LV) mass was higher in men than women
(Levy et al., 1987; Laufer et al., 1989). In a population of pre-
menopausal and post-menopausal women with mild essential
hypertension matched with men (in regard to mean arterial pressure,
age and race), pre-menopausal women had smaller left ventricular
mass and higher LV performance indices than men. These gender
differences were most pronounced before menopause and tended to
disappear after menopause (Garavaglia et al., 1989). Normalized LV
mass was also greater in men than women with aortic stenosis
(Carroll et al., 1992). In other studies, males were found to display
eccentric hypertrophy, while females displayed concentric hypertro-
phy, suggesting the male heart may decompensate quicker (Krumholz
et al., 1993; Douglas et al., 1995). Data from the Framingham Heart
Study suggest that different degrees of hypertrophy in males and
females have an impact on clinical outcomes. Women with LV
hypertrophy (free of coronary heart disease) had better clinical
outcomes than men [age-adjusted prevalence of ventricular arrhyth-
mias higher in men than women (28% versus 17%), 6 year cumulative
mortality higher in men than women (38% versus 22%)] (Bikkina et
al., 1993). Women were also shown to have less cardiac myocyte
apoptosis in normal and in failing hearts compared with males at
autopsy (Guerra et al., 1999; Patten & Karas, 2006). Interestingly,
however, in settings of diabetes or hypertension women have a
greater risk than men to develop cardiovascular disease (Regitz-
Zagrosek, 2006), though the reasons for this remain unclear.
7.2. Physiological hypertrophy
Gender differences in physiological cardiac hypertrophy have not
been examined widely but seem to exist based on animal studies
(Table 4). For example, female rats that underwent chronic swim
training showed an increased hypertrophic response compared with
212 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
their male counterparts (Schaible & Scheuer, 1979, 1981; Luczak &
Leinwand, 2009). Female mice have increased exercise capacity for
both voluntary wheel or treadmill running, with females running
more on a cage wheel than males independent of the strain or age.
Furthermore, female mice performed better in endurance tests,
indicative of increased cardiovascular performance (Konhilas et al.,
2004; Luczak & Leinwand, 2009). While cage wheel running induces
signicant cardiac hypertrophy in both genders, females displayed a
greater percent increase in cardiac mass (Konhilas et al., 2004). To
date, there appears to be no supporting data in humans.
7.3. Molecular mechanisms associated
with gender differences in cardiac physiology
Based on clinical trials in which both genders have been included
in signicant numbers, it is clear that males and females respond
differently to cardiovascular drugs. For example, ACEI and -blockers
are less effective in women and show more side effects (Regitz-
Zagrosek, 2006). Digitalis also causes more deaths in women (Regitz-
Zagrosek, 2006). Furthermore, the prevalence of cardiac disease is
declining in men, but not in women (Zipes et al., 2005). Thus, it is of
great importance to understand the molecular mechanisms respon-
sible for the different hypertrophic responses in males and females.
The molecular mechanisms underlying gender dimorphism are
complex and are still not well understood (Babiker et al., 2002;
Turgeon et al., 2004; Edwards, 2005; Mendelsohn & Karas, 2005;
Luczak & Leinwand, 2009). In general, pre-menopausal women tend
to be protected against cardiovascular disease compared with age-
matched men, but this protection is abolished following menopause.
Thus, it has been suggested that estrogen has protective properties
and activation of signaling cascades downstream of estrogen may
explain gender-related differences in the heart (Kannel, 2002;
Mikkola & Clarkson, 2002; Wenger, 2002; Sullivan, 2003; Zipes et
al., 2005; Luczak & Leinwand, 2009).
The sex steroid hormones (estrogen, progesterone and testoster-
one) and their respective receptors are thought to mediate, at least in
part, gender differences in the heart (Mendelsohn & Karas, 2005; Du
et al., 2006; Luczak & Leinwand, 2009). Estrogen and estrogen
receptors (ERs) have been most extensively characterized. Both men
and women produce estrogen, but circulating levels of estrogen are
1020 fold lower in men (Luczak & Leinwand, 2009). Cardiac
expression of ER is similar in both genders, whereas ER expression
is signicantly higher in males (Mahmoodzadeh et al., 2006).
Downregulation of ERs in response to ovariectomy was associated
with adverse cardiac remodeling and cardiac enlargement in animal
studies, suggesting ERs are important for inhibition of pathological
hypertrophy associated with aging (Xu et al., 2003). Early studies in
knockout mice suggested that ER rather than ER was important for
the development of pathological hypertrophy induced by pressure
overload and cardioprotection in ischemiareperfusion studies (Gabel
et al., 2005; Pelzer et al., 2005; Skavdahl et al., 2005). However, these
ndings have been questioned and require further investigation.
Estrogen is able to initiate both 1) genomic responses through
binding to steroid hormone nuclear receptors, which when bound to
estrogen, modulate the transcriptional activity of target genes; as well
as 2) rapid membrane-initiated, estrogen-triggered signaling
responses (non-genomic) via a plasma membrane-associated form
of the receptor (Fig. 11) (de Jager et al., 2001; Konhilas et al., 2004;
Deroo & Korach, 2006; Du et al., 2006; Moriarty et al., 2006). The
protective properties of estrogen on hypertrophy appear to be
mediated in part by its ability to activate Akt, and inhibit GPCR, PKC,
p38-MAPK, and degrade calcineurin (Fig. 11) (Simoncini et al., 2000;
Konhilas et al., 2004; Du et al., 2006; Liu et al., 2006; Donaldson et al.,
2009). Estrogen has also been shown to initiate anti-hypertrophic
Table 4
Differential hypertrophic responses to pathological and physiological stimuli in males and females.
Pathological/
physiological
Model Subject Age Female Male Reference
Pathological C57BL/6 miceTransverse
aortic constriction
Mice 2 months Hypertrophy Hypertrophy Skavdahl et al., 2005
HW/BW 31%
increase
HW/BW 64%
increase
Pathological Aortic banding (ascending aortic
constriction 4 months)
Rats 6 months Hypertrophy Hypertrophy Douglas et al., 1998
Progression to
heart failure
Chamber
dilation
Pathological Ascending aortic stenosis Rats Not specied Hypertrophy Hypertrophy Weinberg et al., 1999
Fetal gene
expression
Fetal gene
expression
ANP ANP
-MHC -MHC
SERCA2a SERCA2a
Contractile
reserve
Contractile
reserve
Pathological Dahl salt-sensitive hypertensive
rats
Rats 34 months Hypertrophy
(concentric)
Hypertrophy
(eccentric)
Podesser et al., 2007
Septal thickness LVPW
Pathological Spontaneously hypertensive
heart failure rats
Rats 6 months, and failing hearts (18 months
male; 24 months female)
Myocyte cross-
sectional area
Myocyte cross-
sectional area
Tamura et al., 1999
Cardiac
function
Cardiac function
Progression to
heart failure
Mortality
Physiological Voluntary cage wheel running Mice 4 months Hypertrophy Hypertrophy Konhilas et al., 2004.
Exercise
capacity
Physiological Swim training Rats 4 months Hypertrophy Hypertrophy Schaible, T.F. and Scheuer, J.,
1979. [male animals]
Schaible, T.F. and Scheuer, J.,
1981. [female animals]
213 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
signaling, via increased ANP and BNP expression (Fig. 11) (van Eickels
et al., 2001; Jankowski et al., 2005). Activation of Akt is reported to
occur through a direct, non-nuclear pathway involving the regulatory
subunit of PI3K (Fig. 11) (Simoncini et al., 2000). Finally, estrogen is
also thought to have a positive impact on energy metabolism (up-
regulate lipid utilization and down regulate glucose oxidation), in
part, through interaction with PPARs and PPAR-activated
coactivator-1 (Keller et al., 1995; Nunez et al., 1997; Ma et al., 1998;
Tcherepanova et al., 2000; Kamei et al., 2003; Schreiber et al., 2003;
Schreiber et al., 2004; Bourdoncle et al., 2005; Du et al., 2006).
Testosterone is also able to initiate genomic responses via nuclear
androgen receptors that modulate transcription, as well as non-
genomic responses. Adult men have approximately 10 fold higher
circulating testosterone levels than females but androgen receptors
are present in both male and female hearts, suggesting testosterone
has a role in both sexes (see Luczak & Leinwand, 2009). In contrast to
estrogen, testosterone is reported to have pro-hypertrophic proper-
ties (Ikeda et al., 2005; Du et al., 2006). Signaling cascades activated by
testosterone in the heart requires further examination (Du et al.,
2006).
7.4. Gender differences in cardiac
hypertrophy molecular mechanisms based on genetic mouse models
A further understanding of the different hypertrophic responses in
males and females at the molecular level has come from studies of
both male and female genetic mouse models. In the past, studies on
genetically modied mouse models have typically focused on males
though there have been some studies which have examined both
genders. Cardiac-specic over-expression of the 2-adrenergic re-
ceptor, TNF-, and phospholamban was associated with greater
pathological hypertrophy in males compared to females (Kadokami et
al., 2000; Dash et al., 2003; Gao et al., 2003). By contrast, cardiac
specic transgenic expression of dn-p38 leads to more hypertrophy
in females than in males under basal conditions and in response to
pressure overload (Liu et al., 2006). Disruption of FKBP12.6 caused
pathological hypertrophy in male but not in female mice (Xin et al.,
2002), and loss of CD38 (regulator of calcium homeostasis) led to
hypertrophy only in male mice (Takahashi et al., 2003). In contrast,
only male mice null for both
1A/C
and
1B
adrenergic receptors or
muscle specic (cardiac and skeletal) transgenic mice over-expres-
sing myostatin showed reduced cardiac size (O'Connell et al., 2003;
Reisz-Porszasz et al., 2003). Mutations of cardiac troponin T are
associated with different hypertrophic responses that are dependent
on gender under basal conditions and in response to pathological
stimuli (e.g. Ang II and isoproterenol) (Maass et al., 2004).
8. Therapeutic strategies for the treatment of heart failure
Current therapies for heart failure include drug therapy, implan-
tation of devices and surgery (see Krum& Abraham, 2009 for detailed
review). Palliative care and exercise training regimes are important
non-pharmacological approaches that can be implemented to
alleviate symptoms and improve quality of life (Flynn et al., 2009;
Goodlin, 2009). This section gives an overview of current drug
therapies and the signaling proteins they target, followed by new
potential therapeutic strategies that may be implemented in the
future based on differences between physiological and pathological
cardiac hypertrophy.
8.1. Current drug therapy
Traditionally, research and therapy has focused on identifying and
inhibiting processes associated with pathological hypertrophy, cardi-
ac dysfunction, and the transition to heart failure. It is not uncommon
for heart failure patients to be prescribed between two to seven
medications alone, including medications to treat the side effects.
Heart failure medications include ACEI, diuretics, -blockers, ARBs,
hydrolazine, and nitrates (McMurray & Pfeffer, 2005). Side effects can
include coughing, uid retention, joint pain, hypotension, fatigue,
depression, renal insufcient, anemia and headaches. ACEI and ARBs
are typically the rst line of defence for patients with heart failure
(Hunt et al., 2009). As outlined previously, Ang II signaling via GPCRs
is a stimulus for pathological hypertrophy and brosis. ACEI and ARBs
reduce blood pressure and attenuate LV remodeling by reducing
signaling via the Ang II receptors.
-blockers are another class of drug that are used in combination
with ACEI/ARBs to treat patients with heart failure (Hunt et al., 2009).
Fig. 11. Signaling cascades involved in estrogen-mediated cardiac hypertrophic responses. Estrogen acts via the estrogen receptor which interacts with the p85 regulatory subunit of
PI3K to induce physiological hypertrophy and attenuate pathological signaling cascades. Nuclear estrogen receptors alter gene expression to regulate cardiac hypertrophic
responses. Estrogen is also able to induce ubiquitination of calcineurin, which leads to degradation of calcineurin by the proteasome. Ang II: angiotensin II, ANP: Atrial natriuretic
peptide, BNP: B-type natriuretic peptide, CN: calcineurin, ER: estrogen receptor, ERK: extracellular regulated kinase, ET-1: endothelin-1, GPCR: G protein-coupled receptor, IGF1:
insulin-like growth factor 1, IRS1: insulin receptor substrate 1, PI3K: phosphoinositide 3-kinase, PKC: protein kinase C, Ub: ubiquitin.
214 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Table 5
Mouse models which highlight the protective effects of IGF1-PI3K-Akt signaling in settings of heart disease.
Protective mechanism Mouse model Expression/activity Evidence of protective effect in settings of cardiovascular disease
IGF1 signaling IGF1 transgenic (Reiss et al., 1996)
Cardiac-specic
84% increase in circulating IGF1 due
to increased secretion from cardiac myocytes
Dilated cardiomyopathy
Attenuated LV remodeling and cardiac dysfunction, prevented apoptosis (Welch et al., 2002)
Diabetic cardiomyopathy
Reduced cardiac dysfunction (Kajstura et al., 2001)
Eccentric cardiac hypertrophy
Attenuated cardiomyocyte necrosis (Li et al., 1999)
Myocardial infarction
Attenuated ventricular dilation and cell death (Li et al., 1997)
IGF1R transgenic (McMullen et al., 2004b)
Cardiac-specic
20-fold increase in IGF1R expression
in cardiac myocytes
Pressure overload
Blunted pathological hypertrophy and brosis (McMullen et al., 2004b)
Diabetic cardiomyopathy
Reduced cadiac dysfunction and brosisi (Huynh et al., 2010)
PI3K activity caPI3K transgenic (Shioi et al., 2000)
Cardiac-specic
6.5-fold increase in PI3K(p110)
activity in cardiac myocytes
Pressure overload
Prevented cardiac dysfunction, blunted pathological hypertrophy and brosis (McMullen et al., 2007)
Dilated cardiomyopathy
Prolonged lifespan (McMullen et al., 2007)
Pathological hypertrophy/heart failure
Prevented brosis and improved cardiac function in mice overexpressing PKC2 (Rigor et al., 2009)
dnPI3K transgenic (Shioi et al., 2000)
Cardiac-specic
77% reduction in PI3K(p110)
activity in cardiac myocytes
Pressure overload
Reduced PI3K(p110) activity exacerbated brosis and cardiac dysfunction, indicating that
PI3K(p110) is important for maintaining cardiac structure and function in settings of cardiac stress
(McMullen et al., 2003; McMullen et al., 2007)
Dilated cardiomyopathy
Reduced lifespan (McMullen et al., 2007)
Induced atrial brillation, reduced cardiac function, increased brosis (Pretorius et al., 2009a)
Akt activity Akt-nuc (Shiraishi et al., 2004) Increased accumulation and activation
of Akt in cardiac myocyte nuclei
Pressure overload
Attenuated LV remodeling and cardiac dysfunction, improved survival (Tsujita et al., 2006)
Akt1
/
(Cho et al., 2001) Loss of Akt1 mRNA and protein
expression due to targeted disruption
of Akt1 gene (whole body)
Pressure overload
Loss of Akt1 exacerbated hypertrophy and cardiac dysfunction, indicating that Akt is
important for protecting the heart in settings of cardiac stress (DeBosch et al., 2006b)
Exercise (activates
IGF1-PI3K signaling)
Voluntary cage wheel running Hypertrophic cardiomyopathy
Reversed expression of hypertrophic markers, reduced brosis, inhibited apoptosis
(Konhilas et al., 2006)
Swim training Dilated cardiomyopathy
Prolonged lifespan (McMullen et al., 2007)
2
1
5
B
.
C
.
B
e
r
n
a
r
d
o
e
t
a
l
.
/
P
h
a
r
m
a
c
o
l
o
g
y
&
T
h
e
r
a
p
e
u
t
i
c
s
1
2
8
(
2
0
1
0
)
1
9
1

2
2
7
The use of -blockers to treat heart failure patients was rst
introduced by Waagstein et al. (1975). This was subsequently
conrmed in later studies (Swedberg et al., 1979, 1980; Waagstein
et al., 1989). Since these discoveries, several randomized, controlled
clinical trials (e.g. CIBIS-II, MERIT-HF) and smaller studies have shown
that administration of -blockers improved survival rates and/or was
associated with improvement in left ventricular ejection fraction
(reviewed by Molenaar & Parsonage, 2005). -blockers reduce
symptoms and improve survival by reducing signal transduction via
-ARs (see Bristow, 2000).
As heart failure progresses, administration of diuretics is essential
for alleviating uid retention (Hunt et al., 2009). Despite the success
of pharmacological agents that reduce pathological remodeling,
mortality remains high, with approximately one third of patients
with heart failure dying within a year of diagnosis (McMurray &
Pfeffer, 2005). Thus, novel therapeutic strategies are needed to further
reduce mortality due to heart failure.
8.2. Novel therapeutic strategies for the treatment of heart failure
Current drug treatments usually delay heart failure progression
rather than regressing it. Thus, there is an urgent need for the
development of therapies that have the potential to improve function
of the failing heart. Identication of the molecular distinction between
pathological and physiological cardiac hypertrophy has provided a
newavenue for tackling this problem. An alternate strategy to directly
inhibiting pathological heart growth is to activate regulators of
physiological heart growth.
8.3. Benets of activating physiological
signaling cascades in a setting of cardiac disease
Induction of physiological cardiac hypertrophy may be a potential
therapeutic strategy for the treatment of heart failure (see Pretorius et
al., 2008; Owen et al., 2009). Regular physical activity protects against
cardiovascular disease, and exercise training in stable chronic heart
failure patient groups is safe and benecial (Scheuer et al., 1982;
Jennings et al., 1986; Nelson et al., 1986; Schaible et al., 1986;
Jennings, 1995; Orenstein et al., 1995; Coats, 2000; Konhilas et al.,
2006). In animal studies, exercise training was shown to reverse
functional and molecular abnormalities associated with cardiac
pathology (Scheuer et al., 1982; Schaible et al., 1986; Orenstein et
al., 1995; McMullen et al., 2003; Konhilas et al., 2006; McMullen et al.,
2007). As described in Section 3.1.1, the IGF1-PI3K(p110)-Akt
pathway is a critical mediator of exercise-induced physiological
hypertrophy. Studies from genetic mouse models have highlighted
the benecial effects of this pathway in settings of cardiac stress
(Table 5).
Over-expression of IGF1 or IGF1R was benecial in models of
pressure overload (McMullen et al., 2004b), dilated cardiomyopathy
(Welch et al., 2002), myocardial infarction (Li et al., 1997), decom-
pensated eccentric hypertrophy (Li et al., 1999), and diabetic
cardiomyopathy (Kajstura et al., 2001; Huynh et al., 2010). caPI3K
mice maintain normal cardiac function and are protected from
developing pathological hypertrophy and brosis when subjected to
pressure overload (McMullen et al., 2007). In contrast, dnPI3K mice
with decreased cardiac PI3K(p110) activity had depressed cardiac
function and increased brosis after pressure overload (McMullen et
al., 2003; McMullen et al., 2007). Further supporting evidence of a
protective role of PI3K(p110) came fromgenetically crossing caPI3K
and dnPI3K mice with a transgenic mouse model of dilated
cardiomyopathy (DCM-Tg) (McMullen et al., 2007). Under basal
conditions, DCM-Tg mice had ventricular dilation, impaired systolic
function, congestive heart failure, and premature death (Buerger et al.,
2006; McMullen et al., 2007). Increasing PI3K(p110) activity in
hearts of this model by crossing DCM-Tg with caPI3K transgenic mice,
delayed the onset of heart failure, and improved lifespan. Conversely,
decreasing PI3K(p110) activity dramatically accelerated the pro-
gression of heart failure and lifespan was shortened (McMullen et al.,
2007). In a more recent study, reduction of PI3K(p110) activity in
another transgenic mouse model of dilated cardiomyopathy (mam-
malian sterile 20-like kinase 1; (Yamamoto et al., 2003)) caused atrial
brillation that was associated with heart failure and premature death
(Pretorius et al., 2009a). Furthermore, PI3K(p110) activity was
reduced in atrial appendages from patients who developed atrial
brillation, but not patients in sinus rhythm (Pretorius et al., 2009a).
Akt1 has also been shown to protect the heart in a setting of pressure
overload-induced hypertrophy. Akt1
/
displayed more hypertrophy
and cardiac dysfunction in response to aortic-banding (DeBosch et al.,
2006b). Finally, the PI3K-Akt pathway may also play a role in
protecting the heart against dysfunction in a setting of pressure
overload via an interaction with an intercalated disc protein, nectin-2
(McMullen 2009; Satomi-Kobayashi et al., 2009). Taken together,
these studies suggest that the IGF1-PI3K(p110)-Akt1 pathway
protects the heart against cardiac insults (summarized in Table 5).
Mechanisms via which activation of the IGF1-PI3K(p110)-Akt
cascade leads to protection against cardiac dysfunction and progres-
sion to heart failure include: anti-brotic properties, anti-apoptotic
properties, maintenance of proteins associated with cardiac contrac-
tile function (e.g. SERCA2a), and inhibition of pathological signaling
cascades (see Section 8.4) (McMullen, 2008).
Other novel therapeutic approaches that are currently under
investigation include statin therapy, vasopressin-receptor antago-
nists, inhibition of oxidative stress, improved Ca
2+
handling,
prevention of apoptosis, and cardiac regeneration (Landmesser and
Drexler, 2005).
8.4. Possible advantages of activating physiological signaling cascades
A dual action of physiological signaling was recently identied.
Studies utilizing PI3K(p110) transgenic and Akt1 knockout mice
demonstrated that the PI3K(p110)-Akt1 pathway can inhibit
pathological growth in addition to promoting physiological growth.
dnPI3K transgenic mice and Akt1 knockout mice showed an
exaggerated hypertrophic response to pressure overload whereas
IGF1R and caPI3K showed a blunted response (McMullen et al., 2004a;
DeBosch et al., 2006b; McMullen et al., 2007). The PI3K(p110)-Akt1
pathways appear to inhibit pathological growth by inhibiting
signaling proteins downstream of GPCR, including ERK1/2 and PKC
(Fig. 5) (DeBosch et al., 2006b; McMullen et al., 2007; Rigor et al.,
2009).
8.5. Activating the IGF1-PI3K pathway in a clinical setting
Epidemiological studies in the general population suggest that
serum IGF1 levels in the lower normal range are associated with
increased risk of acute myocardial infarction, ischemic heart disease
and heart failure (Juul et al., 2002; Vasan et al., 2003; Laughlin et al.,
2004). This is consistent with ndings from patients with growth
hormone deciency (see review (Colao, 2008)). Patients with
hypopituitarism had a 2-fold higher risk of dying from cardiovascular
disease compared with healthy controls (Rosen and Bengtsson, 1990;
Tomlinson et al., 2001). Furthermore, growth hormone deciency has
been associated with decreased cardiac size and impairment in
cardiac function (Colao et al., 2002; Shulman et al., 2003; Salerno et
al., 2004). In support of the suggestion that these cardiac effects are
due to growth hormone deciency, multiple studies have reported
that growth hormone replacement in these patients increases heart
size and preserves/improves cardiac function (see review(Shulman et
al., 2003; Salerno et al., 2004; Colao, 2008)). However, of note,
patients with acromegaly (growth hormone excess) develop cardiac
hypertrophy that has been associated with cardiac dysfunction.
216 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Furthermore, while growth hormone and IGF1 have been considered
as potential therapeutic agents in patients with heart failure, results
have been conicting (see Colao et al., 2001).
While substantial evidence shows the benets of activating the
IGF1-PI3K(p110) pathway in the heart (see Section 8.3), there are
challenges in targeting PI3K(p110) directly because of its numerous
actions in various cell types. Of particular note, PI3K(p110) permits
cancer cells to bypass normal growth-limiting controls (McMullen &
Jay, 2007). Thus, it is of interest to identify downstream targets
regulated by PI3K(p110) that may represent more specic thera-
peutic targets. We have recently identied cardiac-selective miRNAs
and mRNAs regulated by PI3K(p110) (Lin et al., 2010). Modulation
of these targets are likely to be better tolerated in patients than
activating PI3K(p110) directly.
9. Summary
The generation and characterization of transgenic and knockout
mice have allowed investigators to use a reductionist approach to
delineate molecular mechanisms that are responsible for mediating
distinct forms of heart growth. Further progress has been accom-
plished by subjecting these models to pathological and physiological
stimuli. Recognition of distinct mechanisms responsible for the
induction of pathological and physiological cardiac hypertrophy has
provided new possibilities for drug discovery. Identication of unique
regulators of physiological heart growth may lead to the development
of innovative pharmacotherapies in the clinical management of heart
failure. A better understanding of the mechanisms responsible for the
development of concentric and eccentric hypertrophy, gender
differences, and the key mechanisms responsible for the transition
from hypertrophy to heart failure will also be important for the
development of new and improved therapeutics with an ability to
improve function of the failing heart as opposed to delaying disease
progression.
Acknowledgments
We acknowledge funding support from the National Health and
Medical Research Council of Australia (NHMRC) and the National
Heart Foundation of Australia. KLW and LP are supported by
Australian Postgraduate Awards and Baker IDI Foundation Postgrad-
uate Awards. JRM is supported by an Australian Research Council
Future Fellowship and holds an Honorary NHMRC Research Fellow-
ship. We also thank Nelly Cemerlang and Joon Win Tan for
administrative support.
References
Adams, J. W., Sakata, Y., Davis, M. G., Sah, V. P., Wang, Y., Liggett, S. B., et al. (1998).
Enhanced Galphaq signaling: a common pathway mediates cardiac hypertrophy
and apoptotic heart failure. Proc Natl Acad Sci U S A 95(17), 1014010145.
Adams, T. E., Epa, V. C., Garrett, T. P., & Ward, C. W. (2000). Structure and function of the
type 1 insulin-like growth factor receptor. Cell Mol Life Sci 57(7), 10501093.
AIHW. (2004). Heart, Stroke and Vascular DiseasesAustralian Facts 2004. Canberra,
Australian Institute for Health and Welfare and National Heart Foundation of
Australia (Cardiovascular Disease Series No. 22).
Aikawa, R., Komuro, I., Yamazaki, T., Zou, Y., Kudoh, S., Zhu, W., et al. (1999). Rho family
small G proteins play critical roles in mechanical stress-induced hypertrophic
responses in cardiac myocytes. Circ Res 84(4), 458466.
Akazawa, H., & Komuro, I. (2003). Roles of cardiac transcription factors in cardiac
hypertrophy. Circ Res 92(10), 10791088.
Akhter, S. A., Luttrell, L. M., Rockman, H. A., Iaccarino, G., Lefkowitz, R. J., & Koch, W. J.
(1998). Targeting the receptor-Gq interface to inhibit in vivo pressure overload
myocardial hypertrophy. Science 280(5363), 574577.
Akki, A., Smith, K., & Seymour, A. M. (2008). Compensated cardiac hypertrophy is
characterised by a decline in palmitate oxidation. Mol Cell Biochem 311(12),
215224.
Allard, M. F., Schonekess, B. O., Henning, S. L., English, D. R., & Lopaschuk, G. D. (1994).
Contribution of oxidative metabolism and glycolysis to ATP production in
hypertrophied hearts. Am J Physiol 267(2 Pt 2), H742H750.
Anand, I., McMurray, J., Cohn, J. N., Konstam, M. A., Notter, T., Quitzau, K., et al. (2004).
Long-term effects of darusentan on left-ventricular remodelling and clinical
outcomes in the EndothelinA Receptor Antagonist Trial in Heart Failure (EARTH):
randomised, double-blind, placebo-controlled trial. Lancet 364(9431), 347354.
Anderson, N. G., Maller, J. L., Tonks, N. K., & Sturgill, T. W. (1990). Requirement for
integration of signals from two distinct phosphorylation pathways for activation of
MAP kinase. Nature 343(6259), 651653.
Antos, C. L., McKinsey, T. A., Frey, N., Kutschke, W., McAnally, J., Shelton, J. M., et al.
(2002). Activated glycogen synthase-3 beta suppresses cardiac hypertrophy in
vivo. Proc Natl Acad Sci U S A 99(2), 907912.
Anversa, P., & Nadal-Ginard, B. (2002). Myocyte renewal and ventricular remodelling.
Nature 415(6868), 240243.
Anversa, P., Leri, A., Kajstura, J., & Nadal-Ginard, B. (2002). Myocyte growth and cardiac
repair. J Mol Cell Cardiol 34(2), 91105.
Aoki, H. and Izumo, S. (2001). Signal transduction of cardiac myocyte hypertrophy.
Heart Physiology and Pathology. Sperelakis, N., Kurachi, Y., Terzic, A. and Cohen, M.V.
San Diego, Academic Press: 10651086.
Aoki, H., Izumo, S., & Sadoshima, J. (1998). Angiotensin II activates RhoA in cardiac
myocytes: a critical role of RhoA in angiotensin II-induced premyobril formation.
Circ Res 82(6), 666676.
Arai, M., Yoguchi, A., Iso, T., Takahashi, T., Imai, S., Murata, K., et al. (1995). Endothelin-1
and its binding sites are upregulated in pressure overload cardiac hypertrophy. Am
J Physiol 268(5 Pt 2), H2084H2091.
Arber, S., Hunter, J. J., Ross, J., Jr., Hongo, M., Sansig, G., Borg, J., et al. (1997). MLP-
decient mice exhibit a disruption of cardiac cytoarchitectural organization, dilated
cardiomyopathy, and heart failure. Cell 88(3), 393403.
Aronow, B. J., Toyokawa, T., Canning, A., Haghighi, K., Delling, U., Kranias, E., et al.
(2001). Divergent transcriptional responses to independent genetic causes of
cardiac hypertrophy. Physiol Genomics 6(1), 1928.
Babiker, F. A., De Windt, L. J., van Eickels, M., Grohe, C., Meyer, R., Doevendans, P. A., et al.
(2002). Estrogenic hormone action in the heart: regulatory network and function.
Cardiovasc Res 53(3), 709719.
Backs, J., & Olson, E. N. (2006). Control of cardiac growth by histone acetylation/
deacetylation. Circ Res 98(1), 1524.
Backs, J., Song, K., Bezprozvannaya, S., Chang, S., & Olson, E. N. (2006). CaM kinase II
selectively signals to histone deacetylase 4 during cardiomyocyte hypertrophy.
J Clin Invest 116(7), 18531864.
Backs, J., Backs, T., Neef, S., Kreusser, M. M., Lehmann, L. H., & Patrick, D. M. (2009). The
delta isoform of CaM kinase II is required for pathological cardiac hypertrophy and
remodeling after pressure overload. Proc Natl Acad Sci U S A 106(7), 23422347.
Badorff, C., Ruetten, H., Mueller, S., Stahmer, M., Gehring, D., Jung, F., et al. (2002). Fas
receptor signaling inhibits glycogen synthase kinase 3 beta and induces cardiac
hypertrophy following pressure overload. J Clin Invest 109(3), 373381.
Baines, C. P., & Molkentin, J. D. (2005). STRESS signaling pathways that modulate
cardiac myocyte apoptosis. J Mol Cell Cardiol 38(1), 4762.
Balakumar, P., & Singh, M. (2006). Differential role of rho-kinase in pathological and
physiological cardiac hypertrophy in rats. Pharmacology 78(2), 9197.
Baltas, L. G., Karczewski, P., & Krause, E. G. (1995). The cardiac sarcoplasmic reticulum
phospholambankinase is a distinct delta-CaMkinase isozyme. FEBS Lett 373(1), 7175.
Barger, P. M., & Kelly, D. P. (2000). PPAR signaling in the control of cardiac energy
metabolism. Trends Cardiovasc Med 10(6), 238245.
Barger, P. M., Brandt, J. M., Leone, T. C., Weinheimer, C. J., & Kelly, D. P. (2000).
Deactivation of peroxisome proliferator-activated receptor-alpha during cardiac
hypertrophic growth. J Clin Invest 105(12), 17231730.
Barki-Harrington, L., Perrino, C., & Rockman, H. A. (2004). Network integration of the
adrenergic system in cardiac hypertrophy. Cardiovasc Res 63(3), 391402.
Barrans, J. D., Stamatiou, D., & Liew, C. (2001). Construction of a human cardiovascular
cDNA microarray: portrait of the failing heart. Biochem Biophys Res Commun 280(4),
964969.
Bartel, D. P. (2004). MicroRNAs: genomics, biogenesis, mechanism, and function. Cell
116(2), 281297.
Bassett, J. H., Harvey, C. B., & Williams, G. R. (2003). Mechanisms of thyroid hormone
receptor-specic nuclear and extra nuclear actions. Mol Cell Endocrinol 213(1), 111.
Bassi, R., Heads, R., Marber, M. S., & Clark, J. E. (2008). Targeting p38-MAPK in the
ischaemic heart: kill or cure? Curr Opin Pharmacol 8(2), 141146.
Bedotto, J. B., Gay, R. G., Graham, S. D., Morkin, E., & Goldman, S. (1989). Cardiac
hypertrophy induced by thyroid hormone is independent of loading conditions and
beta adrenoceptor blockade. J Pharmacol Exp Ther 248(2), 632636.
Bendall, J. K., Cave, A. C., Heymes, C., Gall, N., & Shah, A. M. (2002). Pivotal role of a gp91
(phox)-containing NADPH oxidase in angiotensin II-induced cardiac hypertrophy
in mice. Circulation 105(3), 293296.
Benjamin, I. J., Jalil, J. E., Tan, L. B., Cho, K., Weber, K. T., & Clark, W. A. (1989).
Isoproterenol-induced myocardial brosis in relation to myocyte necrosis. Circ Res
65(3), 657670.
Berenji, K., Drazner, M. H., Rothermel, B. A., & Hill, J. A. (2005). Does load-induced
ventricular hypertrophy progress to systolic heart failure? Am J Physiol Heart Circ
Physiol 289(1), H8H16.
Bi, L., Okabe, I., Bernard, D. J., Wynshaw-Boris, A., & Nussbaum, R. L. (1999). Proliferative
defect and embryonic lethality in mice homozygous for a deletion in the p110alpha
subunit of phosphoinositide 3-kinase. J Biol Chem 274(16), 1096310968.
Bikkina, M., Larson, M. G., & Levy, D. (1993). Asymptomatic ventricular arrhythmias and
mortality risk in subjects with left ventricular hypertrophy. J Am Coll Cardiol 22(4),
11111116.
Billet, S., Bardin, S., Verp, S., Baudrie, V., Michaud, A., Conchon, S., et al. (2007). Gain-of-
function mutant of angiotensin II receptor, type 1A, causes hypertension and
cardiovascular brosis in mice. J Clin Invest 117(7), 19141925.
217 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Billet, S., Aguilar, F., Baudry, C., & Clauser, E. (2008). Role of angiotensin II AT1 receptor
activation in cardiovascular diseases. Kidney Int 74(11), 13791384.
Bisognano, J. D., Weinberger, H. D., Bohlmeyer, T. J., Pende, A., Raynolds, M. V., &
Sastravaha, A. (2000). Myocardial-directed overexpression of the human beta(1)-
adrenergic receptor in transgenic mice. J Mol Cell Cardiol 32(5), 817830.
Bisping, E., Ikeda, S., Kong, S. W., Tarnavski, O., Bodyak, N., McMullen, J. R., et al. (2006).
Gata4 is required for maintenance of postnatal cardiac function and protection
from pressure overload-induced heart failure. Proc Natl Acad Sci U S A 103(39),
1447114476.
Bleumink, G. S., Knetsch, A. M., Sturkenboom, M. C., Straus, S. M., Hofman, A., Deckers, J. W.,
et al. (2004). Quantifying the heart failure epidemic: prevalence, incidence rate, lifetime
riskandprognosis of heart failureTheRotterdamStudy. Eur Heart J 25(18), 16141619.
Blinderman, C. D., Homel, P., Billings, J. A., Portenoy, R. K., & Tennstedt, S. L. (2008).
Symptom distress and quality of life in patients with advanced congestive heart
failure. J Pain Sympt Manage 35(6), 594603.
Boengler, K., Hilker-Kleiner, D., Drexler, H., Heusch, G., &Schulz, R. (2008). Themyocardial
JAK/STAT pathway: from protection to failure. Pharmacol Ther 120(2), 172185.
Bogoyevitch, M. A., Gillespie-Brown, J., Ketterman, A. J., Fuller, S. J., Ben-Levy, R.,
Ashworth, A., et al. (1996). Stimulation of the stress-activated mitogen-activated
protein kinase subfamilies in perfused heart. p38/RK mitogen-activated protein
kinases and c-Jun N-terminal kinases are activated by ischemia/reperfusion. Circ
Res 79(2), 162173.
Boluyt, M. O., Brevick, J. L., Rogers, D. S., Randall, M. J., Scalia, A. F., & Li, Z. B. (2006).
Changes in the rat heart proteome induced by exercise training: increased
abundance of heat shock protein hsp20. Proteomics 6(10), 31543169.
Bossuyt, J., Helmstadter, K., Wu, X., Clements-Jewery, H., Haworth, R. S., Avkiran, M.,
et al. (2008). Ca2+/calmodulin-dependent protein kinase IIdelta and protein
kinase D overexpression reinforce the histone deacetylase 5 redistribution in heart
failure. Circ Res 102(6), 695702.
Boulton, T. G., Nye, S. H., Robbins, D. J., Ip, N. Y., Radziejewska, E., Morgenbesser, S. D.,
et al. (1991). ERKs: a family of proteinserine/threonine kinases that are activated
and tyrosine phosphorylated in response to insulin and NGF. Cell 65(4), 663675.
Bourdoncle, A., Labesse, G., Margueron, R., Castet, A., Cavailles, V., & Royer, C. A. (2005).
The nuclear receptor coactivator PGC-1alpha exhibits modes of interaction with the
estrogen receptor distinct from those of SRC-1. J Mol Biol 347(5), 921934.
Bowman, J. C., Steinberg, S. F., Jiang, T., Geenen, D. L., Fishman, G. I., & Buttrick, P. M.
(1997). Expression of protein kinase C beta in the heart causes hypertrophy in adult
mice and sudden death in neonates. J Clin Invest 100(9), 21892195.
Boya, P., Gonzalez-Polo, R. A., Casares, N., Perfettini, J. L., Dessen, P., Larochette, N., et al. (2005).
Inhibition of macroautophagy triggers apoptosis. Mol Cell Biol 25(3), 10251040.
Brancaccio, M., Fratta, L., Notte, A., Hirsch, E., Poulet, R., Guazzone, S., et al. (2003).
Melusin, a muscle-specic integrin beta1-interacting protein, is required to prevent
cardiac failure in response to chronic pressure overload. Nat Med 9(1), 6875.
Braz, J. C., Bueno, O. F., De Windt, L. J., & Molkentin, J. D. (2002). PKC alpha regulates the
hypertrophic growth of cardiomyocytes through extracellular signal-regulated
kinase1/2 (ERK1/2). J Cell Biol 156(5), 905919.
Braz, J. C., Bueno, O. F., Liang, Q., Wilkins, B. J., Dai, Y. S., Parsons, S., et al. (2003).
Targeted inhibition of p38 MAPK promotes hypertrophic cardiomyopathy through
upregulation of calcineurin-NFAT signaling. J Clin Invest 111(10), 14751486.
Braz, J. C., Gregory, K., Pathak, A., Zhao, W., Sahin, B., Klevitsky, R., et al. (2004). PKC-
alpha regulates cardiac contractility and propensity toward heart failure. Nat Med
10(3), 248254.
Bristow, M. R. (2000). Beta-adrenergic receptor blockade in chronic heart failure.
Circulation 101(5), 558569.
Bristow, M. (2003). Antiadrenergic therapy of chronic heart failure: surprises and new
opportunities. Circulation 107(8), 11001102.
Bristow, M. R., Ginsburg, R., Minobe, W., Cubicciotti, R. S., Sageman, W. S., Lurie, K., et al.
(1982). Decreased catecholamine sensitivity and beta-adrenergic-receptor density
in failing human hearts. N Engl J Med 307(4), 205211.
Brooks, W. W., & Conrad, C. H. (2000). Myocardial brosis in transforming growth
factor beta(1)heterozygous mice. J Mol Cell Cardiol 32(2), 187195.
Brower, G. L., Gardner, J. D., Forman, M. F., Murray, D. B., Voloshenyuk, T., Levick, S. P.,
et al. (2006). The relationship between myocardial extracellular matrix remodeling
and ventricular function. Eur J Cardiothorac Surg 30(4), 604610.
Brunner, F., Bras-Silva, C., Cerdeira, A. S., & Leite-Moreira, A. F. (2006). Cardiovascular
endothelins: essential regulators of cardiovascular homeostasis. Pharmacol Ther
111(2), 508531.
Bueno, O. F., De Windt, L. J., Tymitz, K. M., Witt, S. A., Kimball, T. R., Klevitsky, R., et al.
(2000). The MEK1-ERK1/2 signaling pathway promotes compensated cardiac
hypertrophy in transgenic mice. EMBO J 19(23), 63416350.
Bueno, O. F., Wilkins, B. J., Tymitz, K. M., Glascock, B. J., Kimball, T. F., Lorenz, J. N., et al.
(2002). Impaired cardiac hypertrophic response in Calcineurin Abeta-decient
mice. Proc Natl Acad Sci U S A 99(7), 45864591.
Buerger, A., Rozhitskaya, O., Sherwood, M. C., Dorfman, A. L., Bisping, E., Abel, E. D., et al.
(2006). Dilated cardiomyopathy resulting from high-level myocardial expression
of Cre-recombinase. J Card Fail 12(5), 392398.
Buitrago, M., Lorenz, K., Maass, A. H., Oberdorf-Maass, S., Keller, U., Schmitteckert, E. M.,
et al. (2005). The transcriptional repressor Nab1 is a specic regulator of
pathological cardiac hypertrophy. Nat Med 11(8), 837844.
Burelle, Y., Wambolt, R. B., Grist, M., Parsons, H. L., Chow, J. C., Antler, C., et al. (2004).
Regular exercise is associated with a protective metabolic phenotype in the rat
heart. Am J Physiol Heart Circ Physiol 287(3), H1055H1063.
Burniston, J. G. (2009). Adaptation of the rat cardiac proteome in response to intensity-
controlled endurance exercise. Proteomics 9(1), 106115.
Bush, E. W., & McKinsey, T. A. (2009). Targeting histone deacetylases for heart failure.
Expert Opin Ther Targets 13(7), 767784.
Callis, T. E., Pandya, K., Seok, H. Y., Tang, R. -H., Tatsuguchi, M., Huang, Z. -P., et al. (2009).
MicroRNA-208a is a regulator of cardiac hypertrophy and conduction in mice. J Clin
Invest 119(9), 27722786.
Cantley, L. C. (2002). The phosphoinositide 3-kinase pathway. Science 296(5573),
16551657.
Cardiogenomics. (20012003). An NHLBI Sponsored Programfor Genomic Applications
2003, from http://www.cardiogenomics.org.
Care, A., Catalucci, D., Felicetti, F., Bonci, D., Addario, A., Gallo, P., et al. (2007).
MicroRNA-133 controls cardiac hypertrophy. Nat Med 13(5), 613618.
Carley, A. N., & Severson, D. L. (2005). Fatty acid metabolism is enhanced in type 2
diabetic hearts. Biochim Biophys Acta 1734(2), 112126.
Carroll, J. D., Carroll, E. P., Feldman, T., Ward, D. M., Lang, R. M., McGaughey, D., et al.
(1992). Sex-associated differences in left ventricular function in aortic stenosis of
the elderly. Circulation 86(4), 10991107.
Catalucci, D., Gallo, P., & Condorelli, G. (2009). MicroRNAs in cardiovascular biology and
heart disease. Circ Cardiovasc Genet 2(4), 402408.
Chan, H. W., Smith, N. J., Hannan, R. D., & Thomas, W. G. (2006). Tackling the EGFR in
pathological tissue remodelling. Pulm Pharmacol Ther 19(1), 7478.
Chang, S., McKinsey, T. A., Zhang, C. L., Richardson, J. A., Hill, J. A., & Olson, E. N. (2004).
Histone deacetylases 5 and 9 govern responsiveness of the heart to a subset of
stress signals and play redundant roles in heart development. Mol Cell Biol 24(19),
84678476.
Chen, L., Hahn, H., Wu, G., Chen, C. H., Liron, T., Schechtman, D., et al. (2001a). Opposing
cardioprotective actions and parallel hypertrophic effects of delta PKC and epsilon
PKC. Proc Natl Acad Sci U S A 98(20), 1111411119.
Chen, Z., Gibson, T. B., Robinson, F., Silvestro, L., Pearson, G., Xu, B., et al. (2001b). MAP
kinases. Chem Rev 101(8), 24492476.
Chesley, A., Lundberg, M. S., Asai, T., Xiao, R. P., Ohtani, S., Lakatta, E. G., et al. (2000). Thebeta
(2)-adrenergic receptor delivers anantiapoptotic signal tocardiac myocytes throughG
(i)-dependent couplingtophosphatidylinositol 3-kinase. Circ Res 87(12), 11721179.
Chien, K. R. (2001). To Cre or not to Cre: the next generation of mouse models of human
cardiac diseases. Circ Res 88(6), 546549.
Chien, K. R., Knowlton, K. U., Zhu, H., & Chien, S. (1991). Regulation of cardiac gene
expression during myocardial growth and hypertrophy: molecular studies of an
adaptive physiologic response. FASEB J 5(15), 30373046.
Chiloeches, A., Paterson, H. F., Marais, R., Clerk, A., Marshall, C. J., & Sugden, P. H. (1999).
Regulation of Ras.GTP loading and Ras-Raf association in neonatal rat ventricular
myocytes by G protein-coupled receptor agonists and phorbol ester. Activation of
the extracellular signal-regulated kinase cascade by phorbol ester is mediated by
Ras. J Biol Chem 274(28), 1976219770.
Ching, G. W., Franklyn, J. A., Stallard, T. J., Daykin, J., Sheppard, M. C., & Gammage, M. D.
(1996). Cardiac hypertrophy as a result of long-term thyroxine therapy and
thyrotoxicosis. Heart 75(4), 363368.
Chiu, H. C., Kovacs, A., Ford, D. A., Hsu, F. F., Garcia, R., Herrero, P., et al. (2001). A novel
mouse model of lipotoxic cardiomyopathy. J Clin Invest 107(7), 813822.
Cho, H., Thorvaldsen, J. L., Chu, Q., Feng, F., & Birnbaum, M. J. (2001). Akt1/PKBalpha is
required for normal growth but dispensable for maintenance of glucose
homeostasis in mice. J Biol Chem 276(42), 3834938352.
Choukroun, G., Hajjar, R., Kyriakis, J. M., Bonventre, J. V., Rosenzweig, A., & Force, T.
(1998). Role of the stress-activated protein kinases in endothelin-induced
cardiomyocyte hypertrophy. J Clin Invest 102(7), 13111320.
Choukroun, G., Hajjar, R., Fry, S., del Monte, F., Haq, S., Guerrero, J. L., et al. (1999).
Regulation of cardiac hypertrophy in vivo by the stress-activated protein kinases/c-
Jun NH(2)-terminal kinases. J Clin Invest 104(4), 391398.
Christe, M. E., & Rodgers, R. L. (1994). Altered glucose and fatty acid oxidation in hearts
of the spontaneously hypertensive rat. J Mol Cell Cardiol 26(10), 13711375.
Christensen, N. J., & Galbo, H. (1983). Sympathetic nervous activity during exercise.
Annu Rev Physiol 45, 139153.
Chu, G., Egnaczyk, G. F., Zhao, W., Jo, S. -H., Fan, G. -C., Maggio, J. E., et al. (2004).
Phosphoproteome analysis of cardiomyocytes subjected to {beta}-adrenergic
stimulation: identication and characterization of a cardiac heat shock protein
p20. Circ Res 94(2), 184193.
Clark, J. E., Sarafraz, N., & Marber, M. S. (2007). Potential of p38-MAPK inhibitors in the
treatment of ischaemic heart disease. Pharmacol Ther 116(2), 192206.
Clerk, A., & Sugden, P. H. (1999). Activation of protein kinase cascades in the heart by
hypertrophic G protein-coupled receptor agonists. Am J Cardiol 83(12A), 64H69H.
Clerk, A., & Sugden, P. H. (2000). Small guanine nucleotide-binding proteins and
myocardial hypertrophy. Circ Res 86(10), 10191023.
Clerk, A., Pham, F. H., Fuller, S. J., Sahai, E., Aktories, K., Marais, R., et al. (2001).
Regulation of mitogen-activated protein kinases in cardiac myocytes through the
small G protein Rac1. Mol Cell Biol 21(4), 11731184.
Clerk, A., Aggeli, I. K., Stathopoulou, K., & Sugden, P. H. (2006). Peptide growth factors
signal differentially through protein kinase C to extracellular signal-regulated
kinases in neonatal cardiomyocytes. Cell Signal 18(2), 225235.
Coats, A. J. (2000). Exercise training in heart failure. Curr Control Trials Cardiovasc Med 1(3),
155160.
Cohn, J. N., Bristow, M. R., Chien, K. R., Colucci, W. S., Frazier, O. H., Leinwand, L. A., et al.
(1997). Report of the national heart, lung, and blood institute special emphasis
panel on heart failure research. Circulation 95(4), 766770.
Colao, A. (2008). The GH-IGF-I axis and the cardiovascular system: clinical implications.
Clin Endocrinol (Oxf) 69(3), 347358.
Colao, A., Marzullo, P., Di Somma, C., & Lombardi, G. (2001). Growth hormone and the
heart. Clin Endocrinol (Oxf) 54(2), 137154.
Colao, A., Di Somma, C., Salerno, M., Spinelli, L., Orio, F., & Lombardi, G. (2002). The
cardiovascular risk of GH-decient adolescents. J Clin Endocrinol Metab 87(8),
36503655.
218 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Condorelli, G., Drusco, A., Stassi, G., Bellacosa, A., Roncarati, R., Iaccarino, G., et al.
(2002). Akt induces enhanced myocardial contractility and cell size in vivo in
transgenic mice. Proc Natl Acad Sci U S A 99(19), 1233312338.
Conlon, I., & Raff, M. (1999). Size control in animal development. Cell 96(2), 235244.
Cook, S. A., Sugden, P. H., & Clerk, A. (1999). Activation of c-Jun N-terminal kinases and
p38-mitogen-activated protein kinases in human heart failure secondary to
ischaemic heart disease. J Mol Cell Cardiol 31(8), 14291434.
Cooper, G. T. (1987). Cardiocyte adaptation to chronically altered load. Annu Rev Physiol
49, 501518.
Cordes, K. R., & Srivastava, D. (2009). MicroRNA regulation of cardiovascular
development. Circ Res 104(6), 724732.
Cowie, M. R., Wood, D. A., Coats, A. J., Thompson, S. G., Suresh, V., Poole-Wilson, P. A.,
et al. (2000). Survival of patients with a newdiagnosis of heart failure: a population
based study. Heart 83(5), 505510.
Crackower, M. A., Oudit, G. Y., Kozieradzki, I., Sarao, R., Sun, H., Sasaki, T., et al. (2002).
Regulation of myocardial contractility and cell size by distinct PI3K-PTEN signaling
pathways. Cell 110(6), 737749.
D'Angelo, D. D., Sakata, Y., Lorenz, J. N., Boivin, G. P., Walsh, R. A., Liggett, S. B., et al.
(1997). Transgenic Galphaq overexpression induces cardiac contractile failure in
mice. Proc Natl Acad Sci U S A 94(15), 81218126.
Danzi, S., & Klein, I. (2002). Thyroid hormone-regulated cardiac gene expression and
cardiovascular disease. Thyroid 12(6), 467472.
Dash, R., Schmidt, A. G., Pathak, A., Gerst, M. J., Biniakiewicz, D., Kadambi, V. J., et al. (2003).
Differential regulation of p38 mitogen-activated protein kinase mediates gender-
dependent catecholamine-induced hypertrophy. Cardiovasc Res 57(3), 704714.
Davies, M., Hobbs, F., Davis, R., Kenkre, J., Roalfe, A. K., Hare, R., et al. (2001). Prevalence of
left-ventricular systolic dysfunction and heart failure in the Echocardiographic Heart
of England Screening study: a population based study. Lancet 358(9280), 439444.
Davila-Roman, V. G., Vedala, G., Herrero, P., de las Fuentes, L., Rogers, J. G., Kelly, D. P.,
et al. (2002). Altered myocardial fatty acid and glucose metabolism in idiopathic
dilated cardiomyopathy. J Am Coll Cardiol 40(2), 271277.
De Acetis, M., Notte, A., Accornero, F., Selvetella, G., Brancaccio, M., Vecchione, C., et al.
(2005). Cardiac overexpression of melusin protects from dilated cardiomyopathy
due to long-standing pressure overload. Circ Res 96(10), 10871094.
de Gasparo, M., Husain, A., Alexander, W., Catt, K. J., Chiu, A. T., Drew, M., et al. (1995).
Proposed update of angiotensin receptor nomenclature. Hypertension 25(5), 924927.
de Gasparo, M., Catt, K. J., Inagami, T., Wright, J. W., & Unger, T. (2000). International
union of pharmacology. XXIII. The angiotensin II receptors. Pharmacol Rev 52(3),
415472.
de Jager, T., Pelzer, T., Muller-Botz, S., Imam, A., Muck, J., & Neyses, L. (2001).
Mechanisms of estrogen receptor action in the myocardium. Rapid gene activation
via the ERK1/2 pathway and serum response elements. J Biol Chem 276(30),
2787327880.
de Simone, G., Devereux, R. B., Daniels, S. R., & Meyer, R. A. (1995). Gender differences in
left ventricular growth. Hypertension 26(6 Pt 1), 979983.
De Windt, L. J., Lim, H. W., Taigen, T., Wencker, D., Condorelli, G., Dorn, G. W., Jr., et al.
(2000). Calcineurin-mediated hypertrophy protects cardiomyocytes from apopto-
sis in vitro and in vivo: an apoptosis-independent model of dilated heart failure.
Circ Res 86(3), 255263.
De Windt, L. J., Lim, H. W., Bueno, O. F., Liang, Q., Delling, U., Braz, J. C., et al. (2001).
Targeted inhibition of calcineurin attenuates cardiac hypertrophy in vivo. Proc Natl
Acad Sci U S A 98(6), 33223327.
DeBosch, B., Sambandam, N., Weinheimer, C., Courtois, M., & Muslin, A. J. (2006a). Akt2
regulates cardiac metabolism and cardiomyocyte survival. J Biol Chem 281(43),
3284132851.
DeBosch, B., Treskov, I., Lupu, T. S., Weinheimer, C., Kovacs, A., Courtois, M., et al.
(2006b). Akt1 is required for physiological cardiac growth. Circulation 113(17),
20972104.
Delaughter, M. C., Taffet, G. E., Fiorotto, M. L., Entman, M. L., & Schwartz, R. J. (1999).
Local insulin-like growth factor I expression induces physiologic, then pathologic,
cardiac hypertrophy in transgenic mice. FASEB J 13(14), 19231929.
Deroo, B. J., & Korach, K. S. (2006). Estrogen receptors and human disease. J Clin Invest
116(3), 561570.
Devereux, R. B. (2000). Therapeutic options in minimizing left ventricular hypertrophy.
Am Heart J 139(1 Pt 2), S9S14.
Diffee, G. M., Seversen, E. A., Stein, T. D., & Johnson, J. A. (2003). Microarray expression
analysis of effects of exercise training: increase in atrial MLC-1 in rat ventricles. AmJ
Physiol Heart Circ Physiol 284(3), H830H837.
Dillmann, W. H. (2002). Cellular action of thyroid hormone on the heart. Thyroid 12(6),
447452.
Ding, G., Fu, M., Qin, Q., Lewis, W., Kim, H. W., Fukai, T., et al. (2007). Cardiac peroxisome
proliferator-activated receptor gamma is essential in protecting cardiomyocytes
from oxidative damage. Cardiovasc Res 76(2), 269279.
Divakaran, V., & Mann, D. L. (2008). The emerging role of microRNAs in cardiac
remodeling and heart failure. Circ Res 103(10), 10721083.
Donaldson, C., Eder, S., Baker, C., Aronovitz, M. J., Weiss, A. D., Hall-Porter, M., et al.
(2009). Estrogen attenuates left ventricular and cardiomyocyte hypertrophy by an
estrogen receptor-dependent pathway that increases calcineurin degradation. Circ
Res 104(2), 265275 211p following 275.
Dorn, G. W., Jr. (2007). The fuzzy logic of physiological cardiac hypertrophy.
Hypertension 49(5), 962970.
Dorn, G. W., Jr., & Force, T. (2005). Protein kinase cascades in the regulation of cardiac
hypertrophy. J Clin Invest 115(3), 527537.
Dorn, G. W., Jr., Robbins, J., Ball, N., & Walsh, R. A. (1994). Myosin heavy chain regulation
and myocyte contractile depression after LV hypertrophy in aortic-banded mice.
Am J Physiol 267(1 Pt 2), H400H405.
Dorn, G. W., Jr., Robbins, J., & Sugden, P. H. (2003). Phenotyping hypertrophy: eschew
obfuscation. Circ Res 92(11), 11711175.
Douglas, P. S., Otto, C. M., Mickel, M. C., Labovitz, A., Reid, C. L., & Davis, K. B. (1995).
Gender differences in left ventricle geometry and function in patients undergoing
balloon dilatation of the aortic valve for isolated aortic stenosis. NHLBI Balloon
Valvuloplasty Registry. Br Heart J 73(6), 548554.
Douglas, P. S., Katz, S. E., Weinberg, E. O., Chen, M. H., Bishop, S. P., & Lorell, B. H. (1998).
Hypertrophic remodeling: gender differences in the early response to left
ventricular pressure overload. J Am Coll Cardiol 32(4), 11181125.
Du, X. J. (2008). Distinct role of adrenoceptor subtypes in cardiac adaptation to chronic
pressure overload. Clin Exp Pharmacol Physiol 35(3), 355360.
Du, X. J., Fang, L., & Kiriazis, H. (2006). Sex dimorphism in cardiac pathophysiology:
experimental ndings, hormonal mechanisms, and molecular mechanisms.
Pharmacol Ther 111(2), 434475.
Edman, C. F., & Schulman, H. (1994). Identication and characterization of delta B-CaM
kinase and delta C-CaM kinase from rat heart, two new multifunctional Ca2+/
calmodulin-dependent proteinkinase isoforms. BiochimBiophys Acta 1221(1), 89101.
Edwards, D. P. (2005). Regulation of signal transduction pathways by estrogen and
progesterone. Annu Rev Physiol 67, 335376.
Eghbali, M., Deva, R., Alioua, A., Minosyan, T. Y., Ruan, H., Wang, Y., et al. (2005).
Molecular and functional signature of heart hypertrophy during pregnancy. Circ Res
96(11), 12081216.
Eghbali, M., Wang, Y., Toro, L., & Stefani, E. (2006). Heart hypertrophy during
pregnancy: a better functioning heart? Trends Cardiovasc Med 16(8), 285291.
Engelhardt, S., Hein, L., Wiesmann, F., & Lohse, M. J. (1999). Progressive hypertrophy
and heart failure in beta1-adrenergic receptor transgenic mice. Proc Natl Acad Sci
U S A 96(12), 70597064.
Esposito, G., Prasad, S. V., Rapacciuolo, A., Mao, L., Koch, W. J., & Rockman, H. A. (2001).
Cardiac overexpression of a G(q) inhibitor blocks induction of extracellular signal-
regulated kinase and c-Jun NH(2)-terminal kinase activity in in vivo pressure
overload. Circulation 103(10), 14531458.
Estigoy, C. B., Ponten, F., Odeberg, J., Herbert, B., Guilhaus, M., Charleston, M., et al.
(2009). Intercalated discs: multiple proteins perform multiple functions in non-
failing and failing human hearts. Biophys Rev 1, 4349.
Exton, J. H. (1985). Mechanisms involved in alpha-adrenergic phenomena. Am J Physiol
248(6 Pt 1), E633E647.
Fagard, R. H. (1997). Impact of different sports and training on cardiac structure and
function. Cardiol Clin 15(3), 397412.
Fan, G. -C., Chu, G., & Kranias, E. G. (2005a). Hsp20 and its cardioprotection. Trends
Cardiovasc Med 15(4), 138141.
Fan, G., Jiang, Y. -P., Lu, Z., Martin, D. W., Kelly, D. J., Zuckerman, J. M., et al. (2005b). A
transgenic mouse model of heart failure using inducible Galpha q. J Biol Chem 280(48),
4033740346.
Fan, G. -C., Yuan, Q., Song, G., Wang, Y., Chen, G., Qian, J., et al. (2006). Small heat-shock
protein Hsp20 attenuates {beta}-agonist-mediated cardiac remodeling through
apoptosis signal-regulating kinase 1. Circ Res 99(11), 12331242.
Farach-Carson, M. C., & Davis, P. J. (2003). Steroid hormone interactions with target
cells: cross talk between membrane and nuclear pathways. J Pharmacol Exp Ther
307(3), 839845.
Feldman, T., Borow, K. M., Sarne, D. H., Neumann, A., & Lang, R. M. (1986). Myocardial
mechanics in hyperthyroidism: importance of left ventricular loading conditions,
heart rate and contractile state. J Am Coll Cardiol 7(5), 967974.
Ferkey, D. M., & Kimelman, D. (2000). GSK-3: new thoughts on an old enzyme. Dev Biol
225(2), 471479.
Ferrans, V. J. (1984). Cardiac hypertrophy: morphological aspects. Growth of the Heart in
Health and Disease (pp. 187239). New York: Raven Press.
Fielitz, J., Kim, M. S., Shelton, J. M., Qi, X., Hill, J. A., Richardson, J. A., et al. (2008).
Requirement of protein kinase D1 for pathological cardiac remodeling. Proc Natl
Acad Sci U S A 105(8), 30593063.
Finck, B. N., Han, X., Courtois, M., Aimond, F., Nerbonne, J. M., Kovacs, A., et al. (2003). Acritical
role for PPARalpha-mediated lipotoxicity in the pathogenesis of diabetic cardiomyop-
athy: modulation by dietary fat content. Proc Natl Acad Sci U S A 100(3), 12261231.
Fiorotto, M. L., Schwartz, R. J., & Delaughter, M. C. (2003). Persistent IGF-I over-
expression in skeletal muscle transiently enhances DNA accretion and growth.
FASEB J 17(1), 5960.
Fischer, P., & Hilker-Kleiner, D. (2007). Survival pathways in hypertrophy and heart
failure: the gp130-STAT3 axis. Basic Res Cardiol 102(4), 279297.
Fischer, P., &Hilker-Kleiner, D. (2008). Role of gp130-mediated signalling pathways in the
heart and its impact on potential therapeutic aspects. Br J Pharmacol 153(Suppl 1),
S414427.
Flynn, K. E., Pina, I. L., Whellan, D. J., Lin, L., Blumenthal, J. A., Ellis, S. J., et al. (2009).
Effects of exercise training on health status in patients with chronic heart failure:
HF-ACTION randomized controlled trial. JAMA 301(14), 14511459.
Forfar, J. C., Muir, A. L., Sawers, S. A., & Toft, A. D. (1982). Abnormal left ventricular
function in hyperthyroidism: evidence for a possible reversible cardiomyopathy.
N Engl J Med 307(19), 11651170.
Frey, N., & Olson, E. N. (2003). Cardiac hypertrophy: the good, the bad, and the ugly.
Annu Rev Physiol 65, 4579.
Frey, N., McKinsey, T. A., & Olson, E. N. (2000). Decoding calcium signals involved in
cardiac growth and function. Nat Med 6(11), 12211227.
Friddle, C. J., Koga, T., Rubin, E. M., & Bristow, J. (2000). Expression proling reveals
distinct sets of genes altered during induction and regression of cardiac
hypertrophy. Proc Natl Acad Sci U S A 97(12), 67456750.
Gabel, S. A., Walker, V. R., London, R. E., Steenbergen, C., Korach, K. S., & Murphy, E.
(2005). Estrogen receptor beta mediates gender differences in ischemia/reperfu-
sion injury. J Mol Cell Cardiol 38(2), 289297.
219 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Gao, X. -M., Agrotis, A., Autelitano, D. J., Percy, E., Woodcock, E. A., Jennings, G. L., et al.
(2003). Sex hormones and cardiomyopathic phenotype induced by cardiac beta 2-
adrenergic receptor overexpression. Endocrinology 144(9), 40974105.
Garavaglia, G. E., Messerli, F. H., Schmieder, R. E., Nunez, B. D., & Oren, S. (1989). Sex
differences in cardiac adaptation to essential hypertension. Eur Heart J 10(12),
11101114.
Garcia-Sainz, J. A., Vazquez-Prado, J., & Villalobos-Molina, R. (1999). Alpha 1-
adrenoceptors: subtypes, signaling, and roles in health and disease. Arch Med Res
30(6), 449458.
Gembardt, F., Heringer-Walther, S., van Esch, J. H., Sterner-Kock, A., van Veghel, R., Le, T. H.,
et al. (2008). Cardiovascular phenotype of mice lacking all three subtypes of
angiotensin II receptors. FASEB J 22(8), 30683077.
Geng, Y. J., Ishikawa, Y., Vatner, D. E., Wagner, T. E., Bishop, S. P., Vatner, S. F., et al.
(1999). Apoptosis of cardiac myocytes in Gsalpha transgenic mice. Circ Res 84(1),
3442.
Gertz, E. W., Wisneski, J. A., Stanley, W. C., & Neese, R. A. (1988). Myocardial substrate
utilization during exercise in humans. Dual carbon-labeled carbohydrate isotope
experiments. J Clin Invest 82(6), 20172025.
Goodlin, S. J. (2009). Palliative care in congestive heart failure. J Am Coll Cardiol 54(5),
386396.
Gosselin, H., Beliveau, L., Burelle, Y., Clement, R., Lajoie, C., El-Helou, V., et al. (2006).
Disparate regulation of signaling proteins after exercise and myocardial infarction.
Med Sci Sports Exerc 38(3), 455462.
Gottlieb, R. A., Finley, K. D., &Mentzer, R. M., Jr. (2009). Cardioprotection requires taking
out the trash. Basic Res Cardiol 104(2), 169180.
Grandi, A. M., Venco, A., Barzizza, F., Scalise, F., Pantaleo, P., & Finardi, G. (1992).
Inuence of age and sex on left ventricular anatomy and function in normals.
Cardiology 81(1), 813.
Grifths-Jones, S., Grocock, R. J., van Dongen, S., Bateman, A., & Enright, A. J. (2006).
miRBase: microRNA sequences, targets and gene nomenclature.Nucleic Acids Res
34, D140144 Database issue.
Grossman, W., Jones, D., & McLaurin, L. P. (1975). Wall stress and patterns of
hypertrophy in the human left ventricle. J Clin Invest 56(1), 5664.
Guerra, S., Leri, A., Wang, X., Finato, N., Di Loreto, C., Beltrami, C. A., et al. (1999).
Myocyte death in the failing human heart is gender dependent. Circ Res 85(9),
856866.
Gunasinghe, S. K., & Spinale, F. G. (2004). Myocardial basis for heart failure. In D. L. Mann
(Ed.), Role of the Cardiac InterstitiumHeart Failure. (pp. 5770) Philadelphia: Saunders.
Gupta, M. P., Samant, S. A., Smith, S. H., & Shroff, S. G. (2008). HDAC4 and PCAF bind to
cardiac sarcomeres and play a role in regulating myolament contractile activity.
J Biol Chem 283(15), 1013510146.
Gutkind, J. S. (1998a). Cell growth control by G protein-coupled receptors: from signal
transduction to signal integration. Oncogene 17(11 Reviews), 13311342.
Gutkind, J. S. (1998b). The pathways connecting G protein-coupled receptors to the
nucleus through divergent mitogen-activated protein kinase cascades. J Biol Chem
273(4), 18391842.
Hahn, H. S., Marreez, Y., Odley, A., Sterbling, A., Yussman, M. G., Hilty, K. C., et al. (2003).
Protein kinase Calpha negatively regulates systolic and diastolic function in
pathological hypertrophy. Circ Res 93(11), 11111119.
Hamawaki, M., Coffman, T. M., Lashus, A., Koide, M., Zile, M. R., Oliverio, M. I., et al.
(1998). Pressure-overload hypertrophy is unabated in mice devoid of AT1A
receptors. Am J Physiol 274(3 Pt 2), H868873.
Hamdani, N., Kooij, V., van Dijk, S., Merkus, D., Paulus, W. J., Remedios, C. D., et al.
(2008). Sarcomeric dysfunction in heart failure. Cardiovasc Res 77(4), 649658.
Haq, S., Choukroun, G., Kang, Z. B., Ranu, H., Matsui, T., Rosenzweig, A., et al. (2000).
Glycogen synthase kinase-3beta is a negative regulator of cardiomyocyte
hypertrophy. J Cell Biol 151(1), 117130.
Haq, S., Choukroun, G., Lim, H., Tymitz, K. M., del Monte, F., Gwathmey, J., et al. (2001).
Differential activation of signal transduction pathways in human hearts with
hypertrophy versus advanced heart failure. Circulation 103(5), 670677.
Harada, K., Komuro, I., Zou, Y., Kudoh, S., Kijima, K., Matsubara, H., et al. (1998). Acute
pressure overload could induce hypertrophic responses in the heart of angiotensin
II type 1a knockout mice. Circ Res 82(7), 779785.
Hardt, S. E., & Sadoshima, J. (2002). Glycogen synthase kinase-3beta: a novel regulator
of cardiac hypertrophy and development. Circ Res 90(10), 10551063.
Harris, I. S., Zhang, S., Treskov, I., Kovacs, A., Weinheimer, C., & Muslin, A. J. (2004). Raf-1
kinase is required for cardiac hypertrophy and cardiomyocyte survival in response
to pressure overload. Circulation 110(6), 718723.
Harrison, B. C., Kim, M. S., van Rooij, E., Plato, C. F., Papst, P. J., Vega, R. B., et al. (2006).
Regulation of cardiac stress signaling by protein kinase d1. Mol Cell Biol 26(10),
38753888.
Harvey, C. B., & Williams, G. R. (2002). Mechanism of thyroid hormone action. Thyroid
12(6), 441446.
Harwood, A. J. (2001). Regulation of GSK-3: a cellular multiprocessor. Cell 105(7),
821824.
Haunstetter, A., & Izumo, S. (1998). Apoptosis: basic mechanisms and implications for
cardiovascular disease. Circ Res 82(11), 11111129.
Hayakawa, Y., Chandra, M., Miao, W., Shirani, J., Brown, J. H., & Dorn, G. W., Jr. (2003).
Inhibition of cardiac myocyte apoptosis improves cardiac function and abolishes
mortality in the peripartum cardiomyopathy of Galpha(q) transgenic mice.
Circulation 108(24), 30363041.
Hein, L., Barsh, G. S., Pratt, R. E., Dzau, V. J., & Kobilka, B. K. (1995). Behavioural and
cardiovascular effects of disrupting the angiotensin II type-2 receptor in mice.
Nature 377(6551), 744747.
Hein, S., Kostin, S., Heling, A., Maeno, Y., & Schaper, J. (2000). The role of the
cytoskeleton in heart failure. Cardiovasc Res 45(2), 273278.
Henderson, B. C., Sen, U., Reynolds, C., Moshal, K. S., Ovechkin, A., & Tyagi, N. (2007).
Reversal of systemic hypertension-associated cardiac remodeling in chronic
pressure overload myocardium by ciglitazone. Int J Biol Sci 3(6), 385392.
Hill, J. A., Rothermel, B., Yoo, K. -D., Cabuay, B., Demetroulis, E., Weiss, R. M., et al.
(2002). Targeted inhibition of calcineurin in pressure-overload cardiac hypertro-
phy. Preservation of systolic function. J Biol Chem 277(12), 1025110255.
Hirota, H., Chen, J., Betz, U. A., Rajewsky, K., Gu, Y., Ross, J., et al. (1999). Loss of a gp130
cardiac muscle cell survival pathway is a critical event in the onset of heart failure
during biomechanical stress. Cell 97(2), 189198.
Hoch, B., Meyer, R., Hetzer, R., Krause, E. G., & Karczewski, P. (1999). Identication and
expression of delta-isoforms of the multifunctional Ca2+/calmodulin-dependent
protein kinase in failing and nonfailing human myocardium. Circ Res 84(6), 713721.
Hoesl, E., Stieber, J., Herrmann, S., Feil, S., Tybl, E., Hofmann, F., et al. (2008). Tamoxifen-
inducible gene deletion in the cardiac conduction system. J Mol Cell Cardiol 45(1),
6269.
Horiuchi, M., Yoshida, H., Kobayashi, K., Kuriwaki, K., Yoshimine, K., Tomomura, M.,
et al. (1993). Cardiac hypertrophy in juvenile visceral steatosis (jvs) mice with
systemic carnitine deciency. FEBS Lett 326(13), 267271.
Hoshijima, M. (2006). Mechanical stressstrain sensors embedded in cardiac
cytoskeleton: Z disk, titin, and associated structures. Am J Physiol Heart Circ Physiol
290(4), H13131325.
Hoshijima, M., Sah, V. P., Wang, Y., Chien, K. R., & Brown, J. H. (1998). The lowmolecular
weight GTPase Rho regulates myobril formation and organization in neonatal rat
ventricular myocytes. Involvement of Rho kinase. J Biol Chem 273(13), 77257730.
Hudlicka, O., & Brown, M. D. (1996). Postnatal growth of the heart and its blood vessels.
J Vasc Res 33(4), 266287.
Hunt, S. A., Abraham, W. T., Chin, M. H., Feldman, A. M., Francis, G. S., Ganiats, T. G., et al.
(2009). 2009 focused update incorporated into the ACC/AHA 2005 Guidelines for
the Diagnosis and Management of Heart Failure in Adults: a report of the American
College of Cardiology Foundation/American Heart Association Task Force on
Practice Guidelines: developed in collaboration with the International Society for
Heart and Lung Transplantation. Circulation 119(14), e391e479.
Hunter, J. J., & Chien, K. R. (1999). Signaling pathways for cardiac hypertrophy and
failure. N Engl J Med 341(17), 12761283.
Hunter, J. J., Tanaka, N., Rockman, H. A., Ross, J., Jr., & Chien, K. R. (1995). Ventricular
expression of a MLC-2v-ras fusion gene induces cardiac hypertrophy and selective
diastolic dysfunction in transgenic mice. J Biol Chem 270(39), 2317323178.
Huynh, K., McMullen, J. R., Julius, T. L., Tan, J. W., Love, J. E., Cemerlang, N., Kiriazis, H.,
Du, X. J., & Ritchie, R. H. (2010). Cardiac-specic IGF-1 receptor transgenic
expression protects against cardiac brosis and diastolic dysfunction in a mouse
model of diabetic cardiomyopathy. Diabetes 59(6), 15121520.
Hwang, D. M., Dempsey, A. A., Lee, C. Y., & Liew, C. C. (2000). Identication of
differentially expressed genes in cardiac hypertrophy by analysis of expressed
sequence tags. Genomics 66(1), 114.
Hwang, D., Schmitt, W. A., & Stephanopoulos, G. (2002a). Determination of minimum
sample size and discriminatory expression patterns in microarray data. Bioinfor-
matics 18(9), 11841193.
Hwang, J. -J., Allen, P. D., Tseng, G. C., Lam, C. -W., Fananapazir, L., Dzau, V. J., et al.
(2002b). Microarray gene expression proles in dilated and hypertrophic
cardiomyopathic end-stage heart failure. Physiol Genomics 10(1), 3144.
Ichihara, S., Obata, K., Yamada, Y., Nagata, K., Noda, A., & Ichihara, G. (2006). Attenuation
of cardiac dysfunction by a PPAR-alpha agonist is associated with down-regulation
of redox-regulated transcription factors. J Mol Cell Cardiol 41(2), 318329.
Ichiki, T., Labosky, P. A., Shiota, C., Okuyama, S., Imagawa, Y., & Fogo, A. (1995). Effects
on blood pressure and exploratory behaviour of mice lacking angiotensin II type-2
receptor. Nature 377(6551), 748750.
Iemitsu, M., Miyauchi, T., Maeda, S., Sakai, S., Kobayashi, T., & Fujii, N. (2001).
Physiological and pathological cardiac hypertrophy induce different molecular
phenotypes in the rat. Am J Physiol Regul Integr Comp Physiol 281(6), R20292036.
Iemitsu, M., Miyauchi, T., Maeda, S., Sakai, S., Fujii, N., Miyazaki, H., et al. (2003). Cardiac
hypertrophy by hypertension and exercise training exhibits different gene
expression of enzymes in energy metabolism. Hypertens Res 26(10), 829837.
Ikeda, Y., Aihara, K., Sato, T., Akaike, M., Yoshizumi, M., & Suzaki, Y. (2005). Androgen
receptor gene knockout male mice exhibit impaired cardiac growth and exacerbation
of angiotensin II-induced cardiac brosis. J Biol Chem 280(33), 2966129666.
Ikeda, S., Kong, S. W., Lu, J., Bisping, E., Zhang, H., & Allen, P. D. (2007). Altered microRNA
expression in human heart disease. Physiol Genomics 31(3), 367373.
Ito, H., Hirata, Y., Adachi, S., Tanaka, M., Tsujino, M., Koike, A., et al. (1993). Endothelin-1
is an autocrine/paracrine factor in the mechanism of angiotensin II-induced
hypertrophy in cultured rat cardiomyocytes. J Clin Invest 92(1), 398403.
Izumo, S., Nadal-Ginard, B., & Mahdavi, V. (1986). All members of the MHC multigene
family respond to thyroid hormone in a highly tissue-specic manner. Science 231
(4738), 597600.
Izumo, S., Lompre, A. M., Matsuoka, R., Koren, G., Schwartz, K., Nadal-Ginard, B., et al.
(1987). Myosin heavy chain messenger RNA and protein isoformtransitions during
cardiac hypertrophy. Interaction between hemodynamic and thyroid hormone-
induced signals. J Clin Invest 79(3), 970977.
Izumo, S., Nadal-Ginard, B., & Mahdavi, V. (1988). Protooncogene induction and
reprogramming of cardiac gene expression produced by pressure overload. Proc
Natl Acad Sci U S A 85(2), 339343.
Jalil, M. A., Horiuchi, M., Wakamatsu, M., Li, M. X., Begum, L., Suzuki, K., et al. (2006).
Attenuation of cardiac hypertrophy in carnitine-decient juvenile visceral steatosis
(JVS) mice achieved by lowering dietary lipid. J Biochem 139(2), 263270.
Jankowski, M., Wang, D., Mukaddam-Daher, S., & Gutkowska, J. (2005). Pregnancy
alters nitric oxide synthase and natriuretic peptide systems in the rat left ventricle.
J Endocrinol 184(1), 209217.
220 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Jennings, G. L. (1995). Mechanisms for reduction of cardiovascular risk by regular
exercise. Clin Exp Pharmacol Physiol 22(3), 209211.
Jennings, G., Nelson, L., Nestel, P., Esler, M., Korner, P., Burton, D., et al. (1986). The
effects of changes in physical activity on major cardiovascular risk factors,
hemodynamics, sympathetic function, and glucose utilization in man: a controlled
study of four levels of activity. Circulation 73(1), 3040.
Jiang, Y., Gram, H., Zhao, M., New, L., Gu, J., Feng, L., et al. (1997). Characterization of the
structure and function of the fourth member of p38 group mitogen-activated
protein kinases, p38delta. J Biol Chem 272(48), 3012230128.
Jones, E. S., Vinh, A., McCarthy, C. A., Gaspari, T. A., & Widdop, R. E. (2008). AT2 receptors:
functional relevance in cardiovascular disease. Pharmacol Ther 120(3), 292316.
Jopling, C. L., Yi, M., Lancaster, A. M., Lemon, S. M., & Sarnow, P. (2005). Modulation of
hepatitis C virus RNA abundance by a liver-specic MicroRNA. Science 309(5740),
15771581.
Juul, A., Scheike, T., Davidsen, M., Gyllenborg, J., & Jorgensen, T. (2002). Low serum
insulin-like growth factor I is associated with increased risk of ischemic heart
disease: a population-based casecontrol study. Circulation 106(8), 939944.
Kadokami, T., McTiernan, C. F., Kubota, T., Frye, C. S., & Feldman, A. M. (2000). Sex-
related survival differences in murine cardiomyopathy are associated with
differences in TNF-receptor expression. J Clin Invest 106(4), 589597.
Kagiyama, S., Eguchi, S., Frank, G. D., Inagami, T., Zhang, Y. C., & Phillips, M. I. (2002).
Angiotensin II-induced cardiac hypertrophy and hypertension are attenuated by
epidermal growth factor receptor antisense. Circulation 106(8), 909912.
Kajstura, J., Fiordaliso, F., Andreoli, A. M., Li, B., Chimenti, S., & Medow, M. S. (2001). IGF-
1 overexpression inhibits the development of diabetic cardiomyopathy and
angiotensin II-mediated oxidative stress. Diabetes 50(6), 14141424.
Kalra, P. R., Moon, J. C., & Coats, A. J. (2002). Do results of the ENABLE (Endothelin
Antagonist Bosentan for Lowering Cardiac Events in Heart Failure) study spell the
end for non-selective endothelin antagonism in heart failure? Int J Cardiol 85(23),
195197.
Kaluski, E., Cotter, G., Leitman, M., Milo-Cotter, O., Krakover, R., Kobrin, I., et al. (2008).
Clinical and hemodynamic effects of bosentan dose optimization in symptomatic
heart failure patients with severe systolic dysfunction, associated with secondary
pulmonary hypertensiona multi-center randomized study. Cardiology 109(4),
273280.
Kamei, Y., Ohizumi, H., Fujitani, Y., Nemoto, T., Tanaka, T., & Takahashi, N. (2003).
PPARgamma coactivator 1beta/ERR ligand 1 is an ERR protein ligand, whose
expression induces a high-energy expenditure and antagonizes obesity. Proc Natl
Acad Sci U S A 100(21), 1237812383.
Kang, P. M., & Izumo, S. (2000). Apoptosis and heart failure: a critical review of the
literature. Circ Res 86(11), 11071113.
Kang, P. M., Yue, P., & Izumo, S. (2002). New insights into the role of apoptosis in
cardiovascular disease. Circ J 66(1), 19.
Kang, P. M., Yue, P., Liu, Z., Tarnavski, O., Bodyak, N., & Izumo, S. (2004). Alterations in
apoptosis regulatory factors during hypertrophy and heart failure. Am J Physiol
Heart Circ Physiol.
Kankaanpaa, M., Lehto, H. R., Parkka, J. P., Komu, M., Viljanen, A., & Ferrannini, E. (2006).
Myocardial triglyceride content and epicardial fat mass in human obesity:
relationship to left ventricular function and serum free fatty acid levels. J Clin
Endocrinol Metab 91(11), 46894695.
Kannel, W. B. (2002). The Framingham Study: historical insight on the impact of
cardiovascular risk factors in men versus women. J Gend Specif Med 5(2), 2737.
Kasahara, H., Ueyama, T., Wakimoto, H., Liu, M. K., Maguire, C. T., Converso, K. L., et al. (2003).
Nkx2.5 homeoprotein regulates expression of gap junction protein connexin 43 and
sarcomere organizationinpostnatal cardiomyocytes. J Mol Cell Cardiol 35(3), 243256.
Kato, M., Arce, L., & Natarajan, R. (2009). MicroRNAs and their role in progressive
kidney diseases. Clin J Am Soc Nephrol 4(7), 12551266.
Kaumann, A. J., &Molenaar, P. (2008). The low-afnitysite of thebeta1-adrenoceptor andits
relevance to cardiovascular pharmacology. Pharmacol Ther 118(3), 303336.
Kaumann, A., Bartel, S., Molenaar, P., Sanders, L., Burrell, K., & Vetter, D. (1999).
Activation of beta2-adrenergic receptors hastens relaxation and mediates
phosphorylation of phospholamban, troponin I, and C-protein in ventricular
myocardium from patients with terminal heart failure. Circulation 99(1), 6572.
Kedzierski, R. M., & Yanagisawa, M. (2001). Endothelin system: the double-edged
sword in health and disease. Annu Rev Pharmacol Toxicol 41, 851876.
Kedzierski, R. M., Grayburn, P. A., Kisanuki, Y. Y., Williams, C. S., Hammer, R. E.,
Richardson, J. A., et al. (2003). Cardiomyocyte-specic endothelin A receptor
knockout mice have normal cardiac function and an unaltered hypertrophic
response to angiotensin II and isoproterenol. Mol Cell Biol 23(22), 82268232.
Kelland, N. F., & Webb, D. J. (2006). Clinical trials of endothelin antagonists in heart
failure: a question of dose? Exp Biol Med (Maywood) 231(6), 696699.
Keller, H., Givel, F., Perroud, M., & Wahli, W. (1995). Signaling cross-talk between
peroxisome proliferator-activated receptor/retinoid X receptor and estrogen
receptor through estrogen response elements. Mol Endocrinol 9(7), 794804.
Kenessey, A., & Ojamaa, K. (2006). Thyroid hormone stimulates protein synthesis in the
cardiomyocyte by activating the Akt-mTOR and p70S6K pathways. J Biol Chem.
Kerfant, B. G., Zhao, D., Lorenzen-Schmidt, I., Wilson, L. S., Cai, S., Chen, S. R., et al.
(2007). PI3Kgamma is required for PDE4, not PDE3, activity in subcellular
microdomains containing the sarcoplasmic reticular calcium ATPase in cardio-
myocytes. Circ Res 101(4), 400408.
Khan, R., & Sheppard, R. (2006). Fibrosis in heart disease: understanding the role of
transforming growth factor-beta in cardiomyopathy, valvular disease and
arrhythmia. Immunology 118(1), 1024.
Kim, Y., Phan, D., van Rooij, E., Wang, D. Z., McAnally, J., Qi, X., et al. (2008a). The MEF2D
transcription factor mediates stress-dependent cardiac remodeling in mice. J Clin
Invest 118(1), 124132.
Kim, J., Wende, A. R., Sena, S., Theobald, H. A., Soto, J., Sloan, C., et al. (2008b). Insulin-
like growth factor I receptor signaling is required for exercise-induced cardiac
hypertrophy. Mol Endocrinol 22(11), 25312543.
Kingma, J. G. J. (1999). Cardiac adaptation to ischemiareperfusion injury. Ann NY Acad
Sci 874(HEART IN STRESS), 8399.
Kirchhefer, U., Schmitz, W., Scholz, H., & Neumann, J. (1999). Activity of cAMP-
dependent protein kinase and Ca2+/calmodulin-dependent protein kinase in
failing and nonfailing human hearts. Cardiovasc Res 42(1), 254261.
Kiriazis, H., Wang, K., Xu, Q., Gao, X. M., Ming, Z., Su, Y., et al. (2008). Knockout
of beta(1)- and beta(2)-adrenoceptors attenuates pressure overload-induced
cardiac hypertrophy and brosis. Br J Pharmacol 153(4), 684692.
Kirkby, N. S., Hadoke, P. W., Bagnall, A. J., & Webb, D. J. (2008). The endothelin systemas
a therapeutic target in cardiovascular disease: great expectations or bleak house?
Br J Pharmacol 153(6), 11051119.
Klee, C. B., Ren, H., & Wang, X. (1998). Regulation of the calmodulin-stimulated protein
phosphatase, calcineurin. J Biol Chem 273(22), 1336713370.
Knoll, R., Hoshijima, M., & Chien, K. (2003). Cardiac mechanotransduction and
implications for heart disease. J Mol Med 81(12), 750756.
Koch, W. J., Rockman, H. A., Samama, P., Hamilton, R. A., Bond, R. A., Milano, C. A., et al.
(1995). Cardiac function in mice overexpressing the beta-adrenergic receptor
kinase or a beta ARK inhibitor. Science 268(5215), 13501353.
Kodama, H., Fukuda, K., Pan, J., Makino, S., Baba, A., Hori, S., et al. (1997). Leukemia
inhibitory factor, a potent cardiac hypertrophic cytokine, activates the JAK/STAT
pathway in rat cardiomyocytes. Circ Res 81(5), 656663.
Koizumi, A., Nozaki, J., Ohura, T., Kayo, T., Wada, Y., Nezu, J., et al. (1999). Genetic
epidemiology of the carnitine transporter OCTN2 gene in a Japanese population and
phenotypic characterization in Japanese pedigrees with primary systemic carnitine
deciency. Hum Mol Genet 8(12), 22472254.
Kok, K., Geering, B., & Vanhaesebroeck, B. (2009). Regulation of phosphoinositide 3-
kinase expression in health and disease. Trends Biochem Sci 34(3), 115127.
Komuro, I., Kudo, S., Yamazaki, T., Zou, Y., Shiojima, I., & Yazaki, Y. (1996). Mechanical
stretch activates the stress-activated protein kinases in cardiac myocytes. FASEB J
10(5), 631636.
Kong, S. W., Bodyak, N., Yue, P., Liu, Z., Brown, J., Izumo, S., et al. (2005). Genetic
expression proles during physiological and pathological cardiac hypertrophy and
heart failure in rats. Physiol Genomics 21(1), 3442.
Konhilas, J. P., Maass, A. H., Luckey, S. W., Stauffer, B. L., Olson, E. N., Leinwand, L. A., et al.
(2004). Sex modies exercise and cardiac adaptation in mice. AmJ Physiol Heart Circ
Physiol 287(6), H27682776.
Konhilas, J. P., Watson, P. A., Maass, A., Boucek, D. M., Horn, T., & Stauffer, B. L. (2006).
Exercise can prevent and reverse the severity of hypertrophic cardiomyopathy. Circ
Res 98(4), 540548.
Koonen, D. P., Febbraio, M., Bonnet, S., Nagendran, J., Young, M. E., Michelakis, E. D., et al.
(2007). CD36 expression contributes to age-induced cardiomyopathy in mice.
Circulation 116(19), 21392147.
Kovacic-Milivojevic, B., Roediger, F., Almeida, E. A., Damsky, C. H., Gardner, D. G., & Ilic,
D. (2001). Focal adhesion kinase and p130Cas mediate both sarcomeric
organization and activation of genes associated with cardiac myocyte hypertrophy.
Mol Biol Cell 12(8), 22902307.
Koziris, L. P., Hickson, R. C., Chatterton, R. T., Jr., Groseth, R. T., Christie, J. M., Goldies, D. G.,
et al. (1999). Serum levels of total and free IGF-I and IGFBP-3 are increased and
maintained in long-term training. J Appl Physiol 86(4), 14361442.
Krishnan, J., Suter, M., Windak, R., Krebs, T., Felley, A., Montessuit, C., et al. (2009).
Activationof a HIF1alpha-PPARgamma axis underlies the integration of glycolytic and
lipid anabolic pathways inpathologic cardiac hypertrophy. Cell Metab 9(6), 512524.
Krum, H., & Abraham, W. T. (2009). Heart failure. Lancet 373(9667), 941955.
Krumholz, H. M., Larson, M., & Levy, D. (1993). Sex differences in cardiac adaptation to
isolated systolic hypertension. Am J Cardiol 72(3), 310313.
Krutzfeldt, J., Rajewsky, N., Braich, R., Rajeev, K. G., Tuschl, T., Manoharan, M., et al. (2005).
Silencing of microRNAs in vivo with antagomirs. Nature 438(7068), 685689.
Kunisada, K., Negoro, S., Tone, E., Funamoto, M., Osugi, T., Yamada, S., et al. (2000).
Signal transducer and activator of transcription 3 in the heart transduces not only a
hypertrophic signal but a protective signal against doxorubicin-induced cardio-
myopathy. Proc Natl Acad Sci U S A 97(1), 315319.
Kuwajima, M., Lu, K., Sei, M., Ono, A., Hayashi, M., Ishiguro, K., et al. (1998).
Characteristics of cardiac hypertrophy in the juvenile visceral steatosis mouse
with systemic carnitine deciency. J Mol Cell Cardiol 30(4), 773781.
Landmesser, U., & Drexler, H. (2005). Chronic heart failure: an overviewof conventional
treatment versus novel approaches. Nat Clin Pract Cardiovasc Med 2(12), 628638.
Lang, H., & Fekete, D. M. (2001). Lineage analysis in the chicken inner ear shows
differences in clonal dispersion for epithelial, neuronal, and mesenchymal cells. Dev
Biol 234(1), 120137.
Laskowski, A., Woodman, O. L., Cao, A. H., Drummond, G. R., Marshall, T., Kaye, D. M.,
et al. (2006). Antioxidant actions contribute to the antihypertrophic effects of atrial
natriuretic peptide in neonatal rat cardiomyocytes. Cardiovasc Res 72(1), 112123.
Latronico, M. V. G., Catalucci, D., & Condorelli, G. (2008). MicroRNA and cardiac
pathologies. Physiol Genomics 34(3), 239242.
Laufer, E., Jennings, G. L., Korner, P. I., & Dewar, E. (1989). Prevalence of cardiac
structural and functional abnormalities in untreated primary hypertension.
Hypertension 13(2), 151162.
Laughlin, G. A., Barrett-Connor, E., Criqui, M. H., & Kritz-Silverstein, D. (2004). The
prospective association of serum insulin-like growth factor I (IGF-I) and IGF-
binding protein-1 levels with all cause and cardiovascular disease mortality in
older adults: the Rancho Bernardo Study. J Clin Endocrinol Metab 89(1), 114120.
Lazar, M. A. (1993). Thyroid hormone receptors: multiple forms, multiple possibilities.
Endocr Rev 14(2), 184193.
221 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Lazar, M. A., & Chin, W. W. (1990). Nuclear thyroid hormone receptors. J Clin Invest 86(6),
17771782.
Lebrasseur, N. K., Duhaney, T. A., De Silva, D. S., Cui, L., Ip, P. C., & Joseph, L. (2007).
Effects of fenobrate on cardiac remodeling in aldosterone-induced hypertension.
Hypertension 50(3), 489496.
Lecellier, C. H., Dunoyer, P., Arar, K., Lehmann-Che, J., Eyquem, S., Himber, C., et al.
(2005). A cellular microRNA mediates antiviral defense in human cells. Science 308
(5721), 557560.
Lee, R. C., Feinbaum, R. L., & Ambros, V. (1993). The C. elegans heterochronic gene lin-4
encodes small RNAs with antisense complementarity to lin-14. Cell 75(5),
843854.
Leevers, S. J., Weinkove, D., MacDougall, L. K., Hafen, E., & Watereld, M. D. (1996). The
Drosophila phosphoinositide 3-kinase Dp110 promotes cell growth. EMBO J 15(23),
65846594.
Lefkowitz, R. J., Rockman, H. A., & Koch, W. J. (2000). Catecholamines, cardiac beta-
adrenergic receptors, and heart failure. Circulation 101(14), 16341637.
Lehman, J. J., & Kelly, D. P. (2002). Gene regulatory mechanisms governing energy
metabolism during cardiac hypertrophic growth. Heart Fail Rev 7(2), 175185.
Lemke, L. E., Bloem, L. J., Fouts, R., Esterman, M., Sandusky, G., & Vlahos, C. J. (2001).
Decreased p38 MAPK activity in end-stage failing human myocardium: p38 MAPK
alpha is the predominant isoform expressed in human heart. J Mol Cell Cardiol 33(8),
15271540.
Levy, D., Savage, D. D., Garrison, R. J., Anderson, K. M., Kannel, W. B., & Castelli, W. P.
(1987). Echocardiographic criteria for left ventricular hypertrophy: the Framing-
ham Heart Study. Am J Cardiol 59(9), 956960.
Levy, D., Garrison, R. J., Savage, D. D., Kannel, W. B., & Castelli, W. P. (1990). Prognostic
implications of echocardiographically determined left ventricular mass in the
Framingham Heart Study. N Engl J Med 322(22), 15611566.
Lewis, B. P., Burge, C. B., & Bartel, D. P. (2005). Conserved seed pairing, often anked by
adenosines, indicates that thousands of human genes are microRNA targets. Cell
120(1), 1520.
Li, Q., Li, B., Wang, X., Leri, A., Jana, K. P., Liu, Y., et al. (1997). Overexpression of insulin-
like growth factor-1 in mice protects from myocyte death after infarction,
attenuating ventricular dilation, wall stress, and cardiac hypertrophy. J Clin Invest
100(8), 19911999.
Li, W. G., Zaheer, A., Coppey, L., &Oskarsson, H. J. (1998). Activation of JNK in the remote
myocardium after large myocardial infarction in rats. Biochem Biophys Res Commun
246(3), 816820.
Li, B., Setoguchi, M., Wang, X., Andreoli, A. M., Leri, A., Malhotra, A., et al. (1999). Insulin-
like growth factor-1 attenuates the detrimental impact of nonocclusive coronary
artery constriction on the heart. Circ Res 84(9), 10071019.
Li, J., Wang, J., Russell, F. D., & Molenaar, P. (2005). Activation of calcineurin in human
failing heart ventricle by endothelin-1, angiotensin II and urotensin II. Br J
Pharmacol 145(4), 432440.
Liang, Q., Bueno, O. F., Wilkins, B. J., Kuan, C. Y., Xia, Y., & Molkentin, J. D. (2003). c-Jun N-
terminal kinases (JNK) antagonize cardiac growth through cross-talk with
calcineurin-NFAT signaling. EMBO J 22(19), 50795089.
Liang, M., Liu, Y., Mladinov, D., Cowley, A. W., Jr., Trivedi, H., Fang, Y., et al. (2009).
MicroRNA: a new frontier in kidney and blood pressure research. Am J Physiol Ren
Physiol 297(3), F553558.
Liao, P., Georgakopoulos, D., Kovacs, A., Zheng, M., Lerner, D., Pu, H., et al. (2001). The in
vivo role of p38 MAP kinases in cardiac remodeling and restrictive cardiomyopathy.
Proc Natl Acad Sci U S A 98(21), 1228312288.
Lijnen, P., & Petrov, V. (1999). Reninangiotensin system, hypertrophy and gene
expression in cardiac myocytes. J Mol Cell Cardiol 31(5), 949970.
Lim, H. W., De Windt, L. J., Steinberg, L., Taigen, T., Witt, S. A., Kimball, T. R., et al. (2000).
Calcineurin expression, activation, and function in cardiac pressure-overload
hypertrophy. Circulation 101(20), 24312437.
Lin, R. C. Y., Weeks, K. L., Gao, X. M., Williams, R. B. H., Bernardo, B. C., Kiriazis, H., et al.
(2010). PI3K(p110alpha) protects against myocardial infarction-induced heart
failure: identication of PI3K-regulated miRNAs and mRNAs. Arterioscler Thromb
Vasc Biol 30(4), 724732.
Lind, J. M., Chiu, C., & Semsarian, C. (2006). Genetic basis of hypertrophic
cardiomyopathy. Expert Rev Cardiovasc Ther 4(6), 927934.
Ling, H., Zhang, T., Pereira, L., Means, C. K., Cheng, H., Gu, Y., et al. (2009). Requirement for
Ca2+/calmodulin-dependent kinase II in the transition from pressure overload-
induced cardiac hypertrophy to heart failure in mice. J Clin Invest 119(5),
12301240.
Linz, W., Wohlfart, P., Baader, M., Breitschopf, K., Falk, E., Schafer, H. L., et al. (2009). The
peroxisome proliferator-activated receptor-alpha (PPAR-alpha) agonist, AVE8134,
attenuates the progression of heart failure and increases survival in rats. Acta
Pharmacol Sin 30(7), 935946.
Liu, J. C., Chan, P., Chen, J. J., Lee, H. M., Lee, W. S., Shih, N. L., et al. (2004). The inhibitory
effect of trilinolein on norepinephrine-induced beta-myosin heavy chain promoter
activity, reactive oxygen species generation, and extracellular signal-regulated
kinase phosphorylation in neonatal rat cardiomyocytes. J Biomed Sci 11(1), 1118.
Liu, J., Sadoshima, J., Zhai, P., Hong, C., Yang, G., Chen, W., et al. (2006). Pressure
overload induces greater hypertrophy and mortality in female mice with p38alpha
MAPK inhibition. J Mol Cell Cardiol 41(4), 680688.
Liu, W., Zi, M., Jin, J., Prehar, S., Oceandy, D., Kimura, T. E., et al. (2009). Cardiac-specic
deletion of mkk4 reveals its role in pathological hypertrophic remodeling but not in
physiological cardiac growth. Circ Res 104(7), 905914.
Lloyd-Jones, D., Adams, R., Carnethon, M., De Simone, G., Ferguson, T. B., & Flegal, K.
(2009). Heart Disease and Stroke Statistics2009 Update: a report from the
American Heart Association Statistics Committee and Stroke Statistics Subcom-
mittee. Circulation 119(3), 480486.
Lomasney, J. W., &Allen, L. F. (2001). Adrenergic receptors in the cardiovascular system.
In N. Sperelakis, Y. Kurachi, A. Terzic, & M. V. Cohen (Eds.), Heart Physiology and
Pathophysiology,San Diego: Academic Press.
Lorell, B. H. (1999). Role of angiotensin AT1, and AT2 receptors in cardiac hypertrophy
and disease. Am J Cardiol 83(12A), 48H52H.
Lorenz, K., Schmitt, J. P., Schmitteckert, E. M., & Lohse, M. J. (2009). A new type of
ERK1/2 autophosphorylation causes cardiac hypertrophy. Nat Med 15(1), 7583.
Louis, A., Cleland, J. G., Crabbe, S., Ford, S., Thackray, S., Houghton, T., et al. (2001). Clinical
Trials Update: CAPRICORN, COPERNICUS, MIRACLE, STAF, RITZ-2, RECOVER and
RENAISSANCE and cachexia and cholesterol in heart failure. Highlights of the Scientic
Sessions of the American College of Cardiology, 2001. Eur J Heart Fail 3(3), 381387.
Lu, Z., Jiang, Y. -P., Wang, W., Xu, X. -H., Mathias, R. T., & Entcheva, E. (2009). Loss of
cardiac phosphoinositide 3-kinase p110{alpha} results in contractile dysfunction.
Circulation 120(4), 318325.
Luczak, E. D., &Leinwand, L. A. (2009). Sex-based cardiac physiology. Ann Rev Physiol 71,
118.
Luo, J., McMullen, J. R., Sobkiw, C. L., Zhang, L., Dorfman, A. L., & Sherwood, M. C. (2005).
Class IA phosphoinositide 3-kinase regulates heart size and physiological cardiac
hypertrophy. Mol Cell Biol 25(21), 94919502.
Luscher, T. F., Enseleit, F., Pacher, R., Mitrovic, V., Schulze, M. R., Willenbrock, R., et al.
(2002). Hemodynamic and neurohumoral effects of selective endothelin A (ET(A))
receptor blockade in chronic heart failure: the Heart Failure ET(A) Receptor
Blockade Trial (HEAT). Circulation 106(21), 26662672.
Ma, H., Sprecher, H. W., & Kolattukudy, P. E. (1998). Estrogen-induced production of a
peroxisome proliferator-activated receptor (PPAR) ligand in a PPARgamma-
expressing tissue. J Biol Chem 273(46), 3013130138.
Maass, A. H., Ikeda, K., Oberdorf-Maass, S., Maier, S. K. G., & Leinwand, L. A. (2004).
Hypertrophy, brosis, and sudden cardiac death in response to pathological stimuli
in mice with mutations in cardiac troponin T. Circulation 110(15), 21022109.
Mackay, K., & Mochly-Rosen, D. (2001). Localization, anchoring, and functions of
protein kinase C isozymes in the heart. J Mol Cell Cardiol 33(7), 13011307.
MacLellan, W. R., & Schneider, M. D. (2000). Genetic dissection of cardiac growth
control pathways. Annu Rev Physiol 62, 289319.
Mahmoodzadeh, S., Eder, S., Nordmeyer, J., Ehler, E., Huber, O., Martus, P., et al. (2006).
Estrogen receptor alpha up-regulation and redistribution in human heart failure.
FASEB J 20(7), 926934.
Manabe, I., Shindo, T., & Nagai, R. (2002). Gene expression in broblasts and brosis:
involvement in cardiac hypertrophy. Circ Res 91(12), 11031113.
Mangelsdorf, D. J., Thummel, C., Beato, M., Herrlich, P., Schutz, G., Umesono, K., et al.
(1995). The nuclear receptor superfamily: the second decade. Cell 83(6), 835839.
Margulies, K. B., & Houser, S. R. (2004). In D. L. Mann (Ed.), Myocyte Abnormalities in
Human Heart FailureHeart Failure. (pp. 4156) Philadelphia: Saunders.
Maric, C., Ryan, G. B., & Alcorn, D. (1997). Embryonic and postnatal development of the
rat renal interstitium. Anat Embryol (Berl) 195(6), 503514.
Maron, B. J. (2003). Sudden death in young athletes. N Engl J Med 349(11), 10641075.
Maron, B. J., & Pelliccia, A. (2006). The heart of trained athletes: cardiac remodeling and
the risks of sports, including sudden death. Circulation 114(15), 16331644.
Maron, B. J., Gohman, T. E., & Aeppli, D. (1998). Prevalence of sudden cardiac death
during competitive sports activities in Minnesota high school athletes. J Am Coll
Cardiol 32(7), 18811884.
Masaki, H., Kurihara, T., Yamaki, A., Inomata, N., Nozawa, Y., Mori, Y., et al. (1998).
Cardiac-specic overexpression of angiotensin II AT2 receptor causes attenuated
response to AT1 receptor-mediated pressor and chronotropic effects. J Clin Invest
101(3), 527535.
Matsuda, T., Zhai, P., Maejima, Y., Hong, C., Gao, S., Tian, B., et al. (2008). Distinct roles of
GSK-3alpha and GSK-3beta phosphorylation in the heart under pressure overload.
Proc Natl Acad Sci U S A 105(52), 2090020905.
Matsui, T., & Rosenzweig, A. (2005). Convergent signal transduction pathways
controlling cardiomyocyte survival and function: the role of PI 3-kinase and Akt.
J Mol Cell Cardiol 38(1), 6371.
Matsui, T., Li, L., Wu, J. C., Cook, S. A., Nagoshi, T., Picard, M. H., et al. (2002). Phenotypic
spectrum caused by transgenic overexpression of activated Akt in the heart. J Biol
Chem 277(25), 2289622901.
Mayer, P., Mohlig, M., Idlibe, D., & Pfeiffer, A. (1995). Novel and uncommon isoforms of
the calcium sensing enzyme calcium/calmodulin dependent protein kinase II in
heart tissue. Basic Res Cardiol 90(5), 372379.
McGavock, J. M., Victor, R. G., Unger, R. H., & Szczepaniak, L. S. (2006). Adiposity of the
heart, revisited. Ann Intern Med 144(7), 517524.
McKinsey, T. A., Zhang, C. L., & Olson, E. N. (2002). Signaling chromatin to make muscle.
Curr Opin Cell Biol 14(6), 763772.
McMullen, J. R. (2008). Role of insulin-like growth factor 1 and phosphoinositide 3-
kinase in a setting of heart disease. Clin Exp Pharmacol Physiol 35(3), 349354.
McMullen, J. R. (2009). Nectin-2: an intercalated disc protein that maintains cardiac
function in a setting of pressure overload. Hypertension 54(4), 713715.
McMullen, J. R., & Jay, P. Y. (2007). PI3K(p110alpha) inhibitors as anti-cancer agents:
minding the heart. Cell Cycle 6(8).
McMullen, J. R., Shioi, T., Zhang, L., Tarnavski, O., Sherwood, M. C., Kang, P. M., et al.
(2003). Phosphoinositide 3-kinase(p110alpha) plays a critical role for the
induction of physiological, but not pathological, cardiac hypertrophy. Proc Natl
Acad Sci U S A 100(21), 1235512360.
McMullen, J. R., Sherwood, M. C., Tarnavski, O., Zhang, L., Dorfman, A. L., Shioi, T., et al.
(2004a). Inhibition of mTOR signaling with rapamycin regresses established cardiac
hypertrophy induced by pressure overload. Circulation 109(24), 30503055.
McMullen, J. R., Shioi, T., Huang, W. -Y., Zhang, L., Tarnavski, O., Bisping, E., et al. (2004b). The
insulin-like growth factor 1 receptor induces physiological heart growth via the
phosphoinositide 3-kinase(p110alpha) pathway. J Biol Chem 279(6), 47824793.
222 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
McMullen, J. R., Shioi, T., Zhang, L., Tarnavski, O., Sherwood, M. C., Dorfman, A. L., et al.
(2004c). Deletion of ribosomal S6 kinases does not attenuate pathological,
physiological, or insulin-like growth factor 1 receptor-phosphoinositide 3-kinase-
induced cardiac hypertrophy. Mol Cell Biol 24(14), 62316240.
McMullen, J. R., Amirahmadi, F., Woodcock, E. A., Schinke-Braun, M., Bouwman, R. D.,
Hewitt, K. A., et al. (2007). Protective effects of exercise and phosphoinositide 3-
kinase(p110alpha) signaling in dilated and hypertrophic cardiomyopathy. Proc
Natl Acad Sci U S A 104(2), 612617.
McMurray, J. J., & Pfeffer, M. A. (2005). Heart failure. Lancet 365(9474), 18771889.
McMurray, J., Chopra, M., Abdullah, I., Smith, W. E., & Dargie, H. J. (1993). Evidence of
oxidative stress in chronic heart failure in humans. Eur Heart J 14(11), 14931498.
Mende, U., Kagen, A., Cohen, A., Aramburu, J., Schoen, F. J., & Neer, E. J. (1998). Transient
cardiac expression of constitutively active Galphaq leads to hypertrophy and
dilated cardiomyopathy by calcineurin-dependent and independent pathways.
Proc Natl Acad Sci U S A 95(23), 1389313898.
Mendelsohn, M. E., & Karas, R. H. (2005). Molecular and cellular basis of cardiovascular
gender differences. Science 308(5728), 15831587.
Mikkola, T. S., & Clarkson, T. B. (2002). Estrogen replacement therapy, atherosclerosis,
and vascular function. Cardiovasc Res 53(3), 605619.
Minamino, T., Yujiri, T., Terada, N., Taffet, G. E., Michael, L. H., Johnson, G. L., et al. (2002).
MEKK1 is essential for cardiac hypertrophy and dysfunction induced by Gq. Proc
Natl Acad Sci U S A 99(6), 38663871.
Mirotsou, M., Dzau, V. J., Pratt, R. E., & Weinberg, E. O. (2006). Physiological genomics of
cardiac disease: quantitative relationships between gene expression and left
ventricular hypertrophy. Physiol Genomics 27(1), 8694.
Miyata, S., &Haneda, T. (1994). Hypertrophic growth of cultured neonatal rat heart cells
mediated by type 1 angiotensin II receptor. Am J Physiol 266(6 Pt 2), H24432451.
Miyazaki, H., Oka, N., Koga, A., Ohmura, H., Ueda, T., & Imaizumi, T. (2006). Comparison
of gene expression proling in pressure and volume overload-induced myocardial
hypertrophies in rats. Hypertens Res 29(12).
Modesti, P. A., Vanni, S., Bertolozzi, I., Cecioni, I., Polidori, G., Paniccia, R., et al. (2000).
Early sequence of cardiac adaptations and growth factor formation in pressure- and
volume-overload hypertrophy. Am J Physiol Heart Circ Physiol 279(3), H976985.
Mogi, M., Iwai, M., & Horiuchi, M. (2009). New insights into the regulation of
angiotensin receptors. Curr Opin Nephrol Hypertens 18(2), 138143.
Molenaar, P., & Parsonage, W. A. (2005). Fundamental considerations of beta-
adrenoceptor subtypes in human heart failure. Trends Pharmacol Sci 26(7), 368375.
Molkentin, J. D., & Dorn, I. G., Jr. (2001). Cytoplasmic signaling pathways that regulate
cardiac hypertrophy. Annu Rev Physiol 63, 391426.
Molkentin, J. D., Lu, J. -R., Antos, C. L., Markham, B., Richardson, J., Robbins, J., et al.
(1998). A calcineurin-dependent transcriptional pathway for cardiac hypertrophy.
Cell 93(2), 215228.
Moriarty, K., Kim, K. H., & Bender, J. R. (2006). Minireview: estrogen receptor-mediated
rapid signaling. Endocrinology 147(12), 55575563.
Morimoto, S. (2008). Sarcomeric proteins and inherited cardiomyopathies. Cardiovasc
Res 77(4), 659666.
Morisco, C., Zebrowski, D., Condorelli, G., Tsichlis, P., Vatner, S. F., & Sadoshima, J.
(2000). The Akt-glycogen synthase kinase 3beta pathway regulates transcription of
atrial natriuretic factor induced by beta-adrenergic receptor stimulation in cardiac
myocytes. J Biol Chem 275(19), 1446614475.
Morisco, C., Seta, K., Hardt, S. E., Lee, Y., Vatner, S. F., & Sadoshima, J. (2001). Glycogen
synthase kinase 3beta regulates GATA4 in cardiac myocytes. J Biol Chem 276(30),
2858628597.
Morissette, M. R., Cook, S. A., Foo, S., McKoy, G., Ashida, N., Novikov, M., et al. (2006).
Myostatin regulates cardiomyocyte growth through modulation of Akt signaling.
Circ Res 99(1), 1524.
Morkin, E. (2000). Control of cardiac myosin heavy chain gene expression. Microsc Res
Tech 50(6), 522531.
Murdoch, C. E., Zhang, M., Cave, A. C., &Shah, A. M. (2006). NADPHoxidase-dependent redox
signalling in cardiac hypertrophy, remodelling and failure. Cardiovasc Res 71(2),
208215.
Muslin, A. J. (2008). MAPK signalling in cardiovascular health and disease: molecular
mechanisms and therapeutic targets. Clin Sci (Lond) 115(7), 203218.
Nag, A. C. (1980). Study of non-muscle cells of the adult mammalian heart: a ne
structural analysis and distribution. Cytobios 28(109), 4161.
Naga Prasad, S. V., Esposito, G., Mao, L., Koch, W. J., & Rockman, H. A. (2000).
Gbetagamma-dependent phosphoinositide 3-kinase activation in hearts with in
vivo pressure overload hypertrophy. J Biol Chem 275(7), 46934698.
NagaPrasad, S. V., Barak, L. S., Rapacciuolo, A., Caron, M. G., &Rockman, H. A. (2001). Agonist-
dependent recruitment of phosphoinositide 3-kinase to the membrane by beta-
adrenergic receptor kinase 1. A role in receptor sequestration. J Biol Chem 276(22),
1895318959.
Nakagawa, M., Hamaoka, K., Hattori, T., & Sawada, T. (1988). Postnatal DNA synthesis in
hearts of mice: autoradiographic and cytouorometric investigations. Cardiovasc
Res 22(8), 575583.
Neer, E. J. (1995). Heterotrimeric G proteins: organizers of transmembrane signals. Cell
80(2), 249257.
Nelson, L., Jennings, G. L., Esler, M. D., & Korner, P. I. (1986). Effect of changing levels of
physical activity on blood-pressure and haemodynamics in essential hypertension.
Lancet 2(8505), 473476.
Neri Serneri, G. G., Boddi, M., Cecioni, I., Vanni, S., Coppo, M., Papa, M. L., et al. (2001a).
Cardiac angiotensin II formation in the clinical course of heart failure and its
relationship with left ventricular function. Circ Res 88(9), 961968.
Neri Serneri, G. G., Boddi, M., Modesti, P. A., Cecioni, I., Coppo, M., Padeletti, L., et al. (2001b).
Increased cardiac sympathetic activity and insulin-like growth factor-I formation are
associated with physiological hypertrophy in athletes. Circ Res 89(11), 977982.
Neubauer, S. (2007). The failing heartan engine out of fuel. N Engl J Med 356(11),
11401151.
Nicol, R. L., Frey, N., Pearson, G., Cobb, M., Richardson, J., & Olson, E. N. (2001). Activated
MEK5 induces serial assembly of sarcomeres and eccentric cardiac hypertrophy.
EMBO J 20(11), 27572767.
Nienaber, J. J., Tachibana, H., Naga Prasad, S. V., Esposito, G., Wu, D., Mao, L., et al. (2003).
Inhibition of receptor-localized PI3K preserves cardiac beta-adrenergic receptor
function and ameliorates pressure overload heart failure. J Clin Invest 112(7),
10671079.
Nishida, K., Yamaguchi, O., Hirotani, S., Hikoso, S., Higuchi, Y., Watanabe, T., et al.
(2004). p38{alpha} mitogen-activated protein kinase plays a critical role in
cardiomyocyte survival but not in cardiac hypertrophic growth in response to
pressure overload. Mol Cell Biol 24(24), 1061110620.
Nishida, K., Yamaguchi, O., & Otsu, K. (2008). Crosstalk between autophagy and
apoptosis in heart disease. Circ Res 103(4), 343351.
Nishida, K., Kyoi, S., Yamaguchi, O., Sadoshima, J., & Otsu, K. (2009). The role of
autophagy in the heart. Cell Death Differ 16(1), 3138.
Nishiyama, A., Kambe, F., Kamiya, K., Seo, H., & Toyama, J. (1998). Effects of thyroid
status on expression of voltage-gated potassium channels in rat left ventricle.
Cardiovasc Res 40(2), 343351.
Nunez, S. B., Medin, J. A., Braissant, O., Kemp, L., Wahli, W., Ozato, K., et al. (1997).
Retinoid X receptor and peroxisome proliferator-activated receptor activate an
estrogen responsive gene independent of the estrogen receptor. Mol Cell Endocrinol
127(1), 2740.
O'Connell, T. D., Ishizaka, S., Nakamura, A., Swigart, P. M., Rodrigo, M. C., Simpson, G. L.,
et al. (2003). The alpha(1A/C)- and alpha(1B)-adrenergic receptors are required
for physiological cardiac hypertrophy in the double-knockout mouse. J Clin
Investig 111(11), 17831791.
O'Connell, T. D., Swigart, P. M., Rodrigo, M. C., Ishizaka, S., Joho, S., Turnbull, L., et al.
(2006). Alpha1-adrenergic receptors prevent a maladaptive cardiac response to
pressure overload. J Clin Invest 116(4), 10051015.
Okere, I. C., Chess, D. J., McElfresh, T. A., Johnson, J., Rennison, J., Ernsberger, P., et al.
(2005). High-fat diet prevents cardiac hypertrophy and improves contractile
function in the hypertensive dahl salt-sensitive rat. Clin Exp Pharmacol Physiol 32
(10), 825831.
Oliverio, M. I., Kim, H. S., Ito, M., Le, T., Audoly, L., Best, C. F., et al. (1998). Reduced
growth, abnormal kidney structure, and type 2 (AT2) angiotensin receptor-
mediated blood pressure regulation in mice lacking both AT1A and AT1B receptors
for angiotensin II. Proc Natl Acad Sci U S A 95(26), 1549615501.
Olivetti, G., Giordano, G., Corradi, D., Melissari, M., Lagrasta, C., Gambert, S. R., et al.
(1995). Gender differences and aging: effects on the human heart. J Am Coll Cardiol
26(4), 10681079.
Olson, E. N., & Williams, R. S. (2000). Calcineurin signaling and muscle remodeling. Cell
101(7), 689692.
Orenstein, T. L., Parker, T. G., Butany, J. W., Goodman, J. M., Dawood, F., Wen, W. H., et al.
(1995). Favorable left ventricular remodeling following large myocardial infarction
by exercise training. Effect on ventricular morphology and gene expression. J Clin
Invest 96(2), 858866.
Osadchii, O. E. (2007). Myocardial phosphodiesterases and regulation of cardiac
contractility in health and cardiac disease. Cardiovasc Drugs Ther 21(3), 171194.
Ostadal, B., Ostadalova, I., & Dhalla, N. S. (1999). Development of cardiac sensitivity to
oxygen deciency: comparative and ontogenetic aspects. Physiol Rev 79(3),
635659.
Oudit, G. Y., & Kassiri, Z. (2007). Role of PI3 kinase gamma in excitationcontraction
coupling and heart disease. Cardiovasc Hematol Disord Drug Targets 7(4), 295304.
Oudit, G. Y., Crackower, M. A., Eriksson, U., Sarao, R., Kozieradzki, I., Sasaki, T., et al.
(2003). Phosphoinositide 3-kinase gamma-decient mice are protected from
isoproterenol-induced heart failure. Circulation 108(17), 21472152.
Oudit, G. Y., Sun, H., Kerfant, B. G., Crackower, M. A., Penninger, J. M., & Backx, P. H.
(2004). The role of phosphoinositide-3 kinase and PTEN in cardiovascular
physiology and disease. J Mol Cell Cardiol 37(2), 449471.
Owen, K. L., Pretorius, L., & McMullen, J. R. (2009). The protective effects of exercise and
phosphoinositide 3-kinase (p110alpha) in the failing heart. Clin Sci (Lond) 116(5),
365375.
Packer, M., McMurray, J., Massie, B. M., Caspi, A., Charlon, V., Cohen-Solal, A., et al.
(2005). Clinical effects of endothelin receptor antagonism with bosentan in
patients with severe chronic heart failure: results of a pilot study. J Card Fail 11(1),
1220.
Paradis, P., Dali-Youcef, N., Paradis, F. W., Thibault, G., & Nemer, M. (2000).
Overexpression of angiotensin II type I receptor in cardiomyocytes induces cardiac
hypertrophy and remodeling. Proc Natl Acad Sci U S A 97(2), 931936.
Park, T. S., Yamashita, H., Blaner, W. S., & Goldberg, I. J. (2007). Lipids in the heart: a
source of fuel and a source of toxins. Curr Opin Lipidol 18(3), 277282.
Parrizas, M., Saltiel, A. R., & LeRoith, D. (1997). Insulin-like growth factor 1 inhibits
apoptosis using the phosphatidylinositol 3-kinase and mitogen-activated protein
kinase pathways. J Biol Chem 272(1), 154161.
Pasumarthi, K. B., & Field, L. J. (2002). Cardiomyocyte cell cycle regulation. Circ Res 90
(10), 10441054.
Patrucco, E., Notte, A., Barberis, L., Selvetella, G., Maffei, A., Brancaccio, M., et al. (2004).
PI3Kgamma modulates the cardiac response to chronic pressure overload by
distinct kinase-dependent and -independent effects. Cell 118(3), 375387.
Patten, R. D., & Karas, R. H. (2006). Estrogen replacement and cardiomyocyte
protection. Trends Cardiovasc Med 16(3), 6975.
Payne, D. M., Rossomando, A. J., Martino, P., Erickson, A. K., Her, J. H., Shabanowitz, J.,
et al. (1991). Identication of the regulatory phosphorylation sites in pp 42/
mitogen-activated protein kinase (MAP kinase). EMBO J 10(4), 885892.
223 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Pearson, G., Robinson, F., Beers Gibson, T., Xu, B. E., Karandikar, M., Berman, K., et al.
(2001). Mitogen-activated protein (MAP) kinase pathways: regulation and
physiological functions. Endocr Rev 22(2), 153183.
Pellegrini, S., & Dusanter-Fourt, I. (1997). The structure, regulation and function of the
Janus kinases (JAKs) and the signal transducers and activators of transcription
(STATs). Eur J Biochem 248(3), 615633.
Pelzer, T., Loza, P. A., Hu, K., Bayer, B., Dienesch, C., Calvillo, L., et al. (2005). Increased
mortality and aggravation of heart failure in estrogen receptor-beta knockout mice
after myocardial infarction. Circulation 111(12), 14921498.
Peng, X., Kraus, M. S., Wei, H., Shen, T. L., Pariaut, R., Alcaraz, A., et al. (2006).
Inactivation of focal adhesion kinase in cardiomyocytes promotes eccentric cardiac
hypertrophy and brosis in mice. J Clin Invest 116(1), 217227.
Perrino, C., Naga Prasad, S. V., Schroder, J. N., Hata, J. A., Milano, C., & Rockman, H. A.
(2005). Restoration of beta-adrenergic receptor signaling and contractile function
in heart failure by disruption of the betaARK1/phosphoinositide 3-kinase complex.
Circulation 111(20), 25792587.
Perrino, C., Naga Prasad, S. V., Mao, L., Noma, T., Yan, Z., Kim, H. -S., et al. (2006).
Intermittent pressure overload triggers hypertrophy-independent cardiac dys-
function and vascular rarefaction. J Clin Invest 116(6), 15471560.
Perrino, C., Schroder, J. N., Lima, B., Villamizar, N., Nienaber, J. J., Milano, C. A., et al.
(2007). Dynamic regulation of phosphoinositide 3-kinase-gamma activity and
beta-adrenergic receptor trafcking in end-stage human heart failure. Circulation
116(22), 25712579.
Petrich, B. G., & Wang, Y. (2004). Stress-activated MAP kinases in cardiac remodeling and
heart failure; newinsights fromtransgenic studies. Trends Cardiovasc Med 14(2), 5055.
Petrich, B. G., Molkentin, J. D., & Wang, Y. (2003). Temporal activation of c-Jun N-
terminal kinase in adult transgenic heart via cre-loxP-mediated DNA recombina-
tion. FASEB J 17(6), 749751.
Petrich, B.G., Eloff, B. C., Lerner, D. L., Kovacs, A., Saftz, J. E., Rosenbaum, D. S., et al.
(2004). Targeted activation of c-Jun N-terminal kinase in vivo induces restrictive
cardiomyopathy and conduction defects. J Biol Chem 279(15), 1533015338.
Pluim, B. M., Zwinderman, A. H., van der Laarse, A., &van der Wall, E. E. (2000). The athlete's
heart. A meta-analysis of cardiac structure and function. Circulation 101(3), 336344.
Podesser, B. K., Jain, M., Ngoy, S., Apstein, C. S., & Eberli, F. R. (2007). Unveiling gender
differences in demand ischemia: a study in a rat model of genetic hypertension. Eur
J Cardio-Thor Surg 31(2), 298304.
Popescu, L. M., Gherghiceanu, M., Hinescu, M. E., Cretoiu, D., Ceafalan, L., Regalia, T., et al.
(2006). Insights into the interstitium of ventricular myocardium: interstitial Cajal-
like cells (ICLC). J Cell Mol Med 10(2), 429458.
Pracyk, J. B., Tanaka, K., Hegland, D. D., Kim, K. S., Sethi, R., Rovira, I. I., et al. (1998). A
requirement for the rac1 GTPase in the signal transduction pathway leading to
cardiac myocyte hypertrophy. J Clin Invest 102(5), 929937.
Pretorius, L., Owen, K. L., Jennings, G. L., & McMullen, J. R. (2008). Promoting
physiological hypertrophy in the failing heart. Clin Exp Pharmacol Physiol 35(4),
438441.
Pretorius, L., Du, X. J., Woodcock, E. A., Kiriazis, H., Lin, R. C., Marasco, S., et al. (2009a).
Reduced phosphoinositide 3-kinase (p110alpha) activation increases the suscep-
tibility to atrial brillation. Am J Pathol 175(3), 9981009.
Pretorius, L., Owen, K. L., & McMullen, J. R. (2009b). Role of phosphoinositide 3-kinases
in regulating cardiac function. Front Biosci 14, 22212229.
Pu, W. T., Ma, Q., & Izumo, S. (2003). NFAT transcription factors are critical survival
factors that inhibit cardiomyocyte apoptosis during phenylephrine stimulation in
vitro. Circ Res 92(7), 725731.
Purcell, N. H., Wilkins, B. J., York, A., Saba-El-Leil, M. K., Meloche, S., Robbins, J., et al.
(2007). Genetic inhibition of cardiac ERK1/2 promotes stress-induced apoptosis
and heart failure but has no effect on hypertrophy in vivo. Proc Natl Acad Sci U S A
104(35), 1407414079.
Raake, P. W., Vinge, L. E., Gao, E., Boucher, M., Rengo, G., Chen, X., et al. (2008). G
protein-coupled receptor kinase 2 ablation in cardiac myocytes before or after
myocardial infarction prevents heart failure. Circ Res 103(4), 413422.
Radak, Z., Chung, H. Y., & Goto, S. (2008). Systemic adaptation to oxidative challenge
induced by regular exercise. Free Radic Biol Med 44(2), 153159.
Rajewsky, N. (2006). MicroRNA target predictions in animals. Nat Genet 38(Suppl),
S813.
Ramirez, M. T., Sah, V. P., Zhao, X. L., Hunter, J. J., Chien, K. R., & Brown, J. H. (1997a). The
MEKK-JNK pathway is stimulated by alpha1-adrenergic receptor and ras activation
and is associated with in vitro and in vivo cardiac hypertrophy. J Biol Chem 272(22),
1405714061.
Ramirez, M. T., Zhao, X. L., Schulman, H., & Brown, J. H. (1997b). The nuclear deltaB
isoform of Ca2+/calmodulin-dependent protein kinase II regulates atrial natri-
uretic factor gene expression in ventricular myocytes. J Biol Chem 272(49),
3120331208.
Rao, P. K., Toyama, Y., Chiang, H. R., Gupta, S., Bauer, M., Medvid, R., et al. (2009). Loss of
cardiac microRNA-mediated regulation leads to dilated cardiomyopathy and heart
failure. Circ Res 105(6), 585594.
Rapacciuolo, A., Esposito, G., Caron, K., Mao, L., Thomas, S. A., & Rockman, H. A. (2001).
Important role of endogenous norepinephrine and epinephrine in the development
of in vivo pressure-overload cardiac hypertrophy. J Am Coll Cardiol 38(3), 876882.
Regitz-Zagrosek, V. (2006). Therapeutic implications of the gender-specic aspects of
cardiovascular disease. Nature Reviews. Drug Discovery 5(5), 425438.
Reiss, K., Cheng, W., Ferber, A., Kajstura, J., Li, P., Li, B., et al. (1996). Overexpression of
insulin-like growth factor-1 in the heart is coupled with myocyte proliferation in
transgenic mice. Proc Natl Acad Sci U S A 93(16), 86308635.
Reisz-Porszasz, S., Bhasin, S., Artaza, J. N., Shen, R., Sinha-Hikim, I., Hogue, A., et al.
(2003). Lower skeletal muscle mass in male transgenic mice with muscle-specic
overexpression of myostatin. Am J Physiol Endocrinol Metab 285(4), E876888.
Ren, J., Samson, W. K., & Sowers, J. R. (1999). Insulin-like growth factor I as a cardiac
hormone: physiological and pathophysiological implications in heart disease. J Mol
Cell Cardiol 31(11), 20492061.
Ren, X. P., Wu, J., Wang, X., Sartor, M. A., Qian, J., Jones, K., et al. (2009). MicroRNA-320 is
involved in the regulation of cardiac ischemia/reperfusion injury by targeting heat-
shock protein 20. Circulation 119(17), 23572366.
Rich, S., & McLaughlin, V. V. (2003). Endothelin receptor blockers in cardiovascular
disease. Circulation 108(18), 21842190.
Rigor, D. L., Bodyak, N., Bae, S., Choi, J. H., Zhang, L., Ter-Ovanesyan, D., et al. (2009).
Phosphoinositide 3-kinase Akt signaling pathway interacts with protein kinase C
{beta}2 in the regulation of physiologic developmental hypertrophy and heart
function. Am J Physiol Heart Circ Physiol 296(3), H566572.
Rimbaud, S., Sanchez, H., Garnier, A., Fortin, D., Bigard, X., Veksler, V., et al. (2009).
Stimulus specic changes of energy metabolism in hypertrophied heart. J Mol Cell
Cardiol 46(6), 952959.
Rintelen, F., Stocker, H., Thomas, G., & Hafen, E. (2001). PDK1 regulates growth through
Akt and S6K in Drosophila. Proc Nat Acad Sci U S A 98(26), 1502015025.
Robbins, D. J., Zhen, E., Owaki, H., Vanderbilt, C. A., Ebert, D., & Geppert, T. D. (1993).
Regulation and properties of extracellular signal-regulated protein kinases 1 and 2
in vitro. J Biol Chem 268(7), 50975106.
Robinson, E., & Grieve, D. J. (2009). Signicance of peroxisome proliferator-activated
receptors in the cardiovascular system in health and disease. Pharmacol Ther 122(3),
246263.
Rockman, H. A., Koch, W. J., & Lefkowitz, R. J. (2002). Seven-transmembrane-spanning
receptors and heart function. Nature 415(6868), 206212.
Rohrer, D., & Dillmann, W. H. (1988). Thyroid hormone markedly increases the mRNA
coding for sarcoplasmic reticulumCa2+-ATPase in the rat heart. J Biol Chem263(15),
69416944.
Roman, B. B., Geenen, D. L., Leitges, M., &Buttrick, P. M. (2001). PKC-beta is not necessary
for cardiac hypertrophy. Am J Physiol Heart Circ Physiol 280(5), H22642270.
Rosen, T., & Bengtsson, B. A. (1990). Premature mortality due to cardiovascular disease
in hypopituitarism. Lancet 336(8710), 285288.
Rosenkranz, S. (2004). TGF-beta1 and angiotensin networking in cardiac remodeling.
Cardiovasc Res 63(3), 423432.
Rosenkranz, S., Flesch, M., Amann, K., Haeuseler, C., Kilter, H., Seeland, U., et al. (2002).
Alterations of beta-adrenergic signaling and cardiac hypertrophy in transgenic
mice overexpressing TGF-beta(1). Am J Physiol Heart Circ Physiol 283(3),
H12531262.
Rothermel, B. A., McKinsey, T. A., Vega, R. B., Nicol, R. L., Mammen, P., Yang, J., et al.
(2001). Myocyte-enriched calcineurin-interacting protein, MCIP1, inhibits cardiac
hypertrophy in vivo. Proc Natl Acad Sci U S A 98(6), 33283333.
Ruan, H., Li, J., Ren, S., Gao, J., Li, G., & Kim, R. (2009). Inducible and cardiac specic PTEN
inactivation protects ischemia/reperfusion injury. J Mol Cell Cardiol 46(2), 193200.
Rupp, H. (1981). The adaptive changes in the isoenzyme pattern of myosin from
hypertrophied rat myocardium as a result of pressure overload and physical
training. Basic Res Cardiol 76(1), 7988.
Sabri, A., & Steinberg, S. F. (2003). Protein kinase C isoform-selective signals that lead to
cardiac hypertrophy and the progression of heart failure. Mol Cell Biochem 251(12),
97101.
Sack, M. N., Disch, D. L., Rockman, H. A., & Kelly, D. P. (1997). A role for Sp and nuclear
receptor transcription factors in a cardiac hypertrophic growth program. Proc Natl
Acad Sci U S A 94(12), 64386443.
Sadoshima, J., & Izumo, S. (1993). Molecular characterization of angiotensin II-induced
hypertrophy of cardiac myocytes and hyperplasia of cardiac broblasts. Critical role
of the AT1 receptor subtype. Circ Res 73(3), 413423.
Sadoshima, J., & Izumo, S. (1997). The cellular and molecular response of cardiac
myocytes to mechanical stress. Annu Rev Physiol 59, 551571.
Sadoshima, J., Qiu, Z., Morgan, J. P., & Izumo, S. (1995). Angiotensin II and other
hypertrophic stimuli mediated by G protein-coupled receptors activate tyrosine
kinase, mitogen-activated protein kinase, and 90-kD S6 kinase in cardiac myocytes.
The critical role of Ca(2+)-dependent signaling. Circ Res 76(1), 115.
Sadoshima, J., Malhotra, R., & Izumo, S. (1996). The role of the cardiac reninangiotensin
system in load-induced cardiac hypertrophy. J Card Fail 2(4 Suppl), S16.
Sadoshima, J., Montagne, O., Wang, Q., Yang, G., Warden, J., & Liu, J. (2002). The MEKK1-
JNK pathway plays a protective role in pressure overload but does not mediate
cardiac hypertrophy. J Clin Invest 110(2), 271279.
Saito, S., Hiroi, Y., Zou, Y., Aikawa, R., Toko, H., Shibasaki, F., et al. (2000). beta-
Adrenergic pathway induces apoptosis through calcineurin activation in cardiac
myocytes. J Biol Chem 275(44), 3452834533.
Saito, T., Fukuzawa, J., Osaki, J., Sakuragi, H., Yao, N., Haneda, T., et al. (2003). Roles of
calcineurin and calcium/calmodulin-dependent protein kinase II in pressure
overload-induced cardiac hypertrophy. J Mol Cell Cardiol 35(9), 11531160.
Sakamoto, M., Minamino, T., Toko, H., Kayama, Y., Zou, Y., Sano, M., et al. (2006).
Upregulation of heat shock transcription factor 1 plays a critical role in adaptive
cardiac hypertrophy. Circ Res 99(12), 14111418.
Salazar, N. C., Chen, J., & Rockman, H. A. (2007). Cardiac GPCRs: GPCR signaling in
healthy and failing hearts. Biochim Biophys Acta 1768(4), 10061018.
Salerno, M., Esposito, V., Spinelli, L., Di Somma, C., Farina, V., Muzzica, S., et al. (2004).
Left ventricular mass and function in children with GHdeciency before and during
12 months GH replacement therapy. Clin Endocrinol (Oxf) 60(5), 630636.
Sanger, J. W., Ayoob, J. C., Chowrashi, P., Zurawski, D., & Sanger, J. M. (2000). Assembly
of myobrils in cardiac muscle cells.Adv Exp Med Biol 481, 89102 discussion
103105.
Sarkar, S., Vellaichamy, E., Young, D., &Sen, S. (2004). Inuence of cytokines and growth
factors in ANG II-mediated collagen upregulation by broblasts in rats: role of
myocytes. Am J Physiol Heart Circ Physiol 287(1), H107117.
224 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Sassen, S., Miska, E. A., & Caldas, C. (2008). MicroRNA implications for cancer.
Virchows Arch 452(1), 110.
Satomi-Kobayashi, S., Ueyama, T., Mueller, S., Toh, R., Masano, T., Sakoda, T., et al.
(2009). Deciency of nectin-2 leads to cardiac brosis and dysfunction under
chronic pressure overload. Hypertension 54(4), 825831.
Schaible, T. F., & Scheuer, J. (1979). Effects of physical training by running or swimming
on ventricular performance of rat hearts. J Appl Physiol 46(4), 854860.
Schaible, T. F., & Scheuer, J. (1981). Cardiac function in hypertrophied hearts from
chronically exercised female rats. J Appl Physiol 50(6), 11401145.
Schaible, T. F., & Scheuer, J. (1984). Response of the heart to exercise training. In R. Zak
(Ed.), Growth of the Heart in Health and Disease. New York: Raven Press.
Schaible, T. F., Malhotra, A., Ciambrone, G. J., & Scheuer, J. (1986). Chronic swimming
reverses cardiac dysfunction and myosin abnormalities in hypertensive rats. J Appl
Physiol 60(4), 14351441.
Scheinowitz, M., Kessler-Icekson, G., Freimann, S., Zimmermann, R., Schaper, W.,
Golomb, E., et al. (2003). Short- and long-term swimming exercise training
increases myocardial insulin-like growth factor-I gene expression. Growth Horm
IGF Res 13(1), 1925.
Scheuer, J. (1999). Catecholamines in cardiac hypertrophy. Am J Cardiol 83(12A),
70H74H.
Scheuer, J., Malhotra, A., Hirsch, C., Capasso, J., & Schaible, T. F. (1982). Physiologic
cardiac hypertrophy corrects contractile protein abnormalities associated with
pathologic hypertrophy in rats. J Clin Invest 70(6), 13001305.
Schreiber, S. N., Knutti, D., Brogli, K., Uhlmann, T., & Kralli, A. (2003). The transcriptional
coactivator PGC-1 regulates the expression and activity of the orphan nuclear
receptor estrogen-related receptor alpha (ERRalpha). J Biol Chem 278(11),
90139018.
Schreiber, S. N., Emter, R., Hock, M. B., Knutti, D., Cardenas, J., Podvinec, M., et al. (2004).
The estrogen-related receptor alpha (ERRalpha) functions in PPARgamma
coactivator 1alpha (PGC-1alpha)-induced mitochondrial biogenesis. Proc Natl
Acad Sci U S A 101(17), 64726477.
Schunkert, H., Dzau, V. J., Tang, S. S., Hirsch, A. T., Apstein, C. S., & Lorell, B. H. (1990).
Increased rat cardiac angiotensin converting enzyme activity and mRNA expression
in pressure overload left ventricular hypertrophy. Effects on coronary resistance,
contractility, and relaxation. J Clin Invest 86(6), 19131920.
Seed, A., Love, M. P., & McMurray, J. J. (2001). Clinical experience with endothelin
receptor antagonists in chronic heart failure. Heart Fail Rev 6(4), 317323.
Seguchi, O., Takashima, S., Yamazaki, S., Asakura, M., Asano, Y., Shintani, Y., et al. (2007).
A cardiac myosin light chain kinase regulates sarcomere assembly in the vertebrate
heart. J Clin Invest 117(10), 28122824.
Senbonmatsu, T., Ichihara, S., Price, E., Jr., Gaffney, F. A., & Inagami, T. (2000). Evidence
for angiotensin II type 2 receptor-mediated cardiac myocyte enlargement during in
vivo pressure overload. J Clin Invest 106(3), R2529.
Shai, S. Y., Harpf, A. E., Babbitt, C. J., Jordan, M. C., Fishbein, M. C., Chen, J., et al. (2002).
Cardiac myocyte-specic excision of the beta1 integrin gene results in myocardial
brosis and cardiac failure. Circ Res 90(4), 458464.
Sharma, S., Adrogue, J. V., Golfman, L., Uray, I., Lemm, J., Youker, K., et al. (2004).
Intramyocardial lipid accumulation in the failing human heart resembles the
lipotoxic rat heart. FASEB J 18(14), 16921700.
Sheng, Z., Knowlton, K., Chen, J., Hoshijima, M., Brown, J. H., & Chien, K. R. (1997).
Cardiotrophin 1 (CT-1) inhibition of cardiac myocyte apoptosis via a mitogen-
activated protein kinase-dependent pathway. Divergence from downstream CT-1
signals for myocardial cell hypertrophy. J Biol Chem 272(9), 57835791.
Shimoyama, M., Hayashi, D., Takimoto, E., Zou, Y., Oka, T., Uozumi, H., et al. (1999).
Calcineurin plays a critical role in pressure overload-induced cardiac hypertrophy.
Circulation 100(24), 24492454.
Shioi, T., Kang, P. M., Douglas, P. S., Hampe, J., Yballe, C. M., Lawitts, J., et al. (2000). The
conserved phosphoinositide 3-kinase pathway determines heart size in mice.
EMBO J 19(11), 25372548.
Shioi, T., McMullen, J. R., Kang, P. M., Douglas, P. S., Obata, T., Franke, T. F., et al. (2002).
Akt/protein kinase B promotes organ growth in transgenic mice. Mol Cell Biol 22(8),
27992809.
Shioi, T., McMullen, J. R., Tarnavski, O., Converso, K., Sherwood, M. C., Manning, W. J.,
et al. (2003). Rapamycin attenuates load-induced cardiac hypertrophy in mice.
Circulation 107(12), 16641670.
Shiojima, I., Sato, K., Izumiya, Y., Schiekofer, S., Ito, M., Liao, R., et al. (2005). Disruption
of coordinated cardiac hypertrophy and angiogenesis contributes to the transition
to heart failure. J Clin Invest 115(8), 21082118.
Shiraishi, I., Melendez, J., Ahn, Y., Skavdahl, M., Murphy, E., Welch, S., et al. (2004).
Nuclear targeting of Akt enhances kinase activity and survival of cardiomyocytes.
Circ Res 94(7), 884891.
Shubeita, H. E., McDonough, P. M., Harris, A. N., Knowlton, K. U., Glembotski, C. C., Brown,
J. H., et al. (1990). Endothelin induction of inositol phospholipid hydrolysis, sarco-
mere assembly, and cardiac gene expression in ventricular myocytes. A paracrine
mechanism for myocardial cell hypertrophy. J Biol Chem 265(33), 2055520562.
Shulman, D. I., Root, A. W., Diamond, F. B., Bercu, B. B., Martinez, R., & Boucek, R. J., Jr.
(2003). Effects of one year of recombinant human growth hormone (GH) therapy
on cardiac mass and function in children with classical GH deciency. J Clin
Endocrinol Metab 88(9), 40954099.
Simon, M. I., Strathmann, M. P., & Gautam, N. (1991). Diversity of G proteins in signal
transduction. Science 252(5007), 802808.
Simoncini, T., Hafezi-Moghadam, A., Brazil, D. P., Ley, K., Chin, W. W., & Liao, J. K. (2000).
Interaction of oestrogen receptor with the regulatory subunit of phosphatidylino-
sitol-3-OH kinase. Nature 407(6803), 538541.
Skavdahl, M., Steenbergen, C., Clark, J., Myers, P., Demianenko, T., Mao, L., et al. (2005).
Estrogen receptor-beta mediates malefemale differences in the development
of pressure overload hypertrophy. Am J Physiol Heart Circ Physiol 288(2),
H469476.
Smeets, P. J., Teunissen, B. E., Willemsen, P. H., van Nieuwenhoven, F. A., Brouns, A. E.,
Janssen, B. J., et al. (2008). Cardiac hypertrophy is enhanced in PPAR alpha/
mice in response to chronic pressure overload. Cardiovasc Res 78(1), 7989.
Solaro, R. J. (2008). Multiplex kinase signaling modies cardiac function at the level of
sarcomeric proteins. J Biol Chem 283(40), 2682926833.
Son, N. H., Park, T. S., Yamashita, H., Yokoyama, M., Huggins, L. A., Okajima, K., et al.
(2007). Cardiomyocyte expression of PPARgamma leads to cardiac dysfunction in
mice. J Clin Invest 117(10), 27912801.
Soonpaa, M. H., Kim, K. K., Pajak, L., Franklin, M., & Field, L. J. (1996). Cardiomyocyte DNA
synthesis and binucleation during murine development. Am J Physiol 271(5 Pt 2),
H21832189.
Sopontammarak, S., Aliharoob, A., Ocampo, C., Arcilla, R. A., Gupta, M. P., & Gupta, M.
(2005). Mitogen-activated protein kinases (p38 and c-Jun NH2-terminal kinase)
are differentially regulated during cardiac volume and pressure overload
hypertrophy. Cell Biochem Biophys 43(1), 6176.
Srinivasan, M., Edman, C. F., & Schulman, H. (1994). Alternative splicing introduces a
nuclear localization signal that targets multifunctional CaM kinase to the nucleus.
J Cell Biol 126(4), 839852.
Strom, C. C., Aplin, M., Ploug, T., Christoffersen, T. E. H., Langfort, J., Viese, M., et al.
(2005). Expression proling reveals differences in metabolic gene expression
between exercise-induced cardiac effects and maladaptive cardiac hypertrophy.
FEBS J 272(11), 26842695.
Subramaniam, A., Jones, W. K., Gulick, J., Wert, S., Neumann, J., & Robbins, J. (1991).
Tissue-specic regulation of the alpha-myosin heavy chain gene promoter in
transgenic mice. J Biol Chem 266(36), 2461324620.
Sugden, P. H. (2001). Signalling pathways in cardiac myocyte hypertrophy. Ann Med 33(9),
611622.
Sugden, P. H., & Clerk, A. (1998). Cellular mechanisms of cardiac hypertrophy. J Mol Med
76(11), 725746.
Sullivan, J. L. (2003). Aremenstruatingwomenprotectedfromheart diseasebecauseof, or in
spite of, estrogen? Relevance to the iron hypothesis. Am Heart J 145(2), 190194.
Swedberg, K., Hjalmarson, A., Waagstein, F., &Wallentin, I. (1979). Prolongation of survival
incongestivecardiomyopathybybeta-receptor blockade. Lancet 1(8131), 13741376.
Swedberg, K., Hjalmarson, A., Waagstein, F., & Wallentin, I. (1980). Benecial effects of
long-termbeta-blockade in congestive cardiomyopathy. Br Heart J 44(2), 117133.
Swynghedauw, B. (1986). Developmental and functional adaptation of contractile
proteins in cardiac and skeletal muscles. Physiol Rev 66(3), 710771.
Swynghedauw, B. (1999). Molecular mechanisms of myocardial remodeling. Physiol
Rev 79(1), 215262.
Tachibana, H., Perrino, C., Takaoka, H., Davis, R. J., Naga Prasad, S. V., & Rockman, H. A.
(2006). JNK1 is required to preserve cardiac function in the early response to
pressure overload. Biochem Biophys Res Commun 343(4), 10601066.
Taigen, T., De Windt, L. J., Lim, H. W., & Molkentin, J. D. (2000). Targeted inhibition of
calcineurin prevents agonist-induced cardiomyocyte hypertrophy. Proc Natl Acad
Sci U S A 97(3), 11961201.
Takahashi, J., Kagaya, Y., Kato, I., Ohta, J., Isoyama, S., Miura, M., et al. (2003). Decit of
CD38/cyclic ADP-ribose is differentially compensated in hearts by gender. Biochem
Biophys Res Commun 312(2), 434440.
Takahashi, R., Asai, T., Murakami, H., Murakami, R., Tsuzuki, M., Numaguchi, Y., et al.
(2007). Pressure overload-induced cardiomyopathy in heterozygous carrier mice
of carnitine transporter gene mutation. Hypertension 50(3), 497502.
Takeishi, Y., Ping, P., Bolli, R., Kirkpatrick, D. L., Hoit, B. D., & Walsh, R. A. (2000).
Transgenic overexpression of constitutively active protein kinase C epsilon causes
concentric cardiac hypertrophy. Circ Res 86(12), 12181223.
Takeishi, Y., Huang, Q., Abe, J., Glassman, M., Che, W., Lee, J. D., et al. (2001). Src and
multiple MAP kinase activation in cardiac hypertrophy and congestive heart failure
under chronic pressure-overload: comparison with acute mechanical stretch. J Mol
Cell Cardiol 33(9), 16371648.
Tamura, T., Said, S., & Gerdes, A. M. (1999). Gender-related differences in myocyte
remodeling in progression to heart failure. Hypertension 33(2), 676680.
Tang, Y., Zheng, J., Sun, Y., Wu, Z., Liu, Z., & Huang, G. (2009). MicroRNA-1 regulates
cardiomyocyte apoptosis by targeting Bcl-2. Int Heart J 50(3), 377387.
Tatsuguchi, M., Seok, H. Y., Callis, T. E., Thomson, J. M., Chen, J. F., Newman, M., et al.
(2007). Expression of microRNAs is dynamically regulated during cardiomyocyte
hypertrophy. J Mol Cell Cardiol 42(6), 11371141.
Tcherepanova, I., Puigserver, P., Norris, J. D., Spiegelman, B. M., & McDonnell, D. P.
(2000). Modulation of estrogen receptor-alpha transcriptional activity by the
coactivator PGC-1. J Biol Chem 275(21), 1630216308.
Thom, T., Haase, N., Rosamond, W., Howard, V. J., Rumsfeld, J., Manolio, T., et al. (2006).
Heart Disease and Stroke Statistics2006 Update: a report from the American
Heart Association Statistics Committee and Stroke Statistics Subcommittee.
Circulation 113(6), e85151.
Thomas, W. G., Brandenburger, Y., Autelitano, D. J., Pham, T., Qian, H., & Hannan, R. D.
(2002). Adenoviral-directed expression of the type 1A angiotensin receptor
promotes cardiomyocyte hypertrophy via transactivation of the epidermal growth
factor receptor. Circ Res 90(2), 135142.
Thum, T., Catalucci, D., & Bauersachs, J. (2008a). MicroRNAs: novel regulators in cardiac
development and disease. Cardiovasc Res 79(4), 562570.
Thum, T., Gross, C., Fiedler, J., Fischer, T., Kissler, S., Bussen, M., et al. (2008b). MicroRNA-
21 contributes to myocardial disease by stimulating MAP kinase signalling in
broblasts. Nature 456(7224), 980984.
Tobimatsu, T., & Fujisawa, H. (1989). Tissue-specic expression of four types of rat
calmodulin-dependent protein kinase II mRNAs. J Biol Chem 264(30),
1790717912.
225 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Toker, A., & Cantley, L. C. (1997). Signalling through the lipid products of
phosphoinositide-3-OH kinase. Nature 387(6634), 673676.
Tomlinson, J. W., Holden, N., Hills, R. K., Wheatley, K., Clayton, R. N., Bates, A. S.,
et al. (2001). Association between premature mortality and hypopituitarism.
West Midlands Prospective Hypopituitary Study Group. Lancet 357(9254),
425431.
Traverse, J. H., Nesmelov, Y. E., Crampton, M., Lindstrom, P., Thomas, D. D., & Bache, R. J.
(2006). Measurement of myocardial free radical production during exercise using
EPR spectroscopy. Am J Physiol Heart Circ Physiol 290(6), H24532458.
Trivedi, C. M., Luo, Y., Yin, Z., Zhang, M., Zhu, W., Wang, T., et al. (2007). Hdac2 regulates
the cardiac hypertrophic response by modulating Gsk3 beta activity. Nat Med 13(3),
324331.
Tsoutsman, T., Bagnall, R. D., & Semsarian, C. (2008). Impact of multiple gene mutations
in determining the severity of cardiomyopathy and heart failure. Clin Exp Pharmacol
Physiol 35(11), 13491357.
Tsujita, Y., Muraski, J., Shiraishi, I., Kato, T., Kajstura, J., Anversa, P., et al. (2006). Nuclear
targeting of Akt antagonizes aspects of cardiomyocyte hypertrophy. Proc Natl Acad
Sci U S A 103(32), 1194611951.
Tsutsui, H., Ishihara, K., & Cooper, G. T. (1993). Cytoskeletal role in the contractile
dysfunction of hypertrophied myocardium. Science 260(5108), 682687.
Turgeon, J. L., McDonnell, D. P., Martin, K. A., & Wise, P. M. (2004). Hormone therapy:
physiological complexity belies therapeutic simplicity. Science 304(5675), 12691273.
Uozumi, H., Hiroi, Y., Zou, Y., Takimoto, E., Toko, H., Niu, P., et al. (2001). gp130 plays a
critical role in pressure overload-induced cardiac hypertrophy. J Biol Chem 276(25),
2311523119.
van Bilsen, M., van Nieuwenhoven, F. A., & van der Vusse, G. J. (2009). Metabolic
remodelling of the failing heart: benecial or detrimental? Cardiovasc Res 81(3),
420428.
van der Vusse, G. J., Glatz, J. F., Stam, H. C., & Reneman, R. S. (1992). Fatty acid
homeostasis in the normoxic and ischemic heart. Physiol Rev 72(4), 881940.
van Eickels, M., Grohe, C., Cleutjens, J. P., Janssen, B. J., Wellens, H. J., & Doevendans, P. A.
(2001). 17beta-estradiol attenuates the development of pressure-overload
hypertrophy. Circulation 104(12), 14191423.
van Rooij, E., & Olson, E. N. (2007). MicroRNAs: powerful new regulators of heart
disease and provocative therapeutic targets. J Clin Invest 117(9), 23692376.
van Rooij, E., Sutherland, L. B., Liu, N., Williams, A. H., McAnally, J., Gerard, R. D., et al.
(2006). A signature pattern of stress-responsive microRNAs that can evoke cardiac
hypertrophy and heart failure. Proc Natl Acad Sci U S A 103(48), 1825518260.
Vanhaesebroeck, B., & Watereld, M. D. (1999). Signaling by distinct classes of
phosphoinositide 3-kinases. Experiment Cell Res 253(1), 239254.
Vanhaesebroeck, B., Leevers, S. J., Panayotou, G., & Watereld, M. D. (1997).
Phosphoinositide 3-kinases: a conserved family of signal transducers. Trends
Biochem Sci 22(7), 267272.
Vanhaesebroeck, B., Leevers, S. J., Ahmadi, K., Timms, J., Katso, R., Driscoll, P. C., et al.
(2001). Synthesis and function of 3-phosphorylated inositol lipids. Annu Rev
Biochem 70, 535602.
Vasan, R. S., Sullivan, L. M., D'Agostino, R. B., Roubenoff, R., Harris, T., Sawyer, D. B., et al.
(2003). Serum insulin-like growth factor I and risk for heart failure in elderly
individuals without a previous myocardial infarction: the FraminghamHeart Study.
Ann Intern Med 139(8), 642648.
Vega, R. B., Rothermel, B. A., Weinheimer, C. J., Kovacs, A., Naseem, R. H., Bassel-Duby, R.,
et al. (2003). Dual roles of modulatory calcineurin-interacting protein 1 in cardiac
hypertrophy. Proc Natl Acad Sci U S A 100(2), 669674.
Verdu, J., Buratovich, M. A., Wilder, E. L., & Birnbaum, M. J. (1999). Cell-autonomous
regulation of cell and organ growth in Drosophila by Akt/PKB. Nat Cell Biol 1(8), 500.
Vigoreaux, J. O. (1994). The muscle Z band: lessons in stress management. J Muscle Res
Cell Motil 15(3), 237255.
Villarreal, F. J., & Dillmann, W. H. (1992). Cardiac hypertrophy-induced changes in
mRNA levels for TGF-beta 1, bronectin, and collagen. Am J Physiol 262(6 Pt 2),
H18611866.
Waagstein, F., Hjalmarson, A., Varnauskas, E., & Wallentin, I. (1975). Effect of chronic beta-
adrenergic receptor blockade in congestive cardiomyopathy. Br Heart J 37(10),
10221036.
Waagstein, F., Caidahl, K., Wallentin, I., Bergh, C. H., & Hjalmarson, A. (1989). Long-term
beta-blockade in dilated cardiomyopathy. Effects of short- and long-term
metoprolol treatment followed by withdrawal and readministration of metoprolol.
Circulation 80(3), 551563.
Waetzig, V., & Herdegen, T. (2005). Context-specic inhibition of JNKs: overcoming the
dilemma of protection and damage. Trends Pharmacol Sci 26(9), 455461.
Wakasaki, H., Koya, D., Schoen, F. J., Jirousek, M. R., Ways, D. K., Hoit, B. D., et al. (1997).
Targeted overexpression of protein kinase C beta2 isoform in myocardium causes
cardiomyopathy. Proc Natl Acad Sci U S A 94(17), 93209325.
Wang, Y. (2007). Mitogen-activated protein kinases in heart development and diseases.
Circulation 116(12), 14131423.
Wang, L., & Proud, C. G. (2002). Ras/Erk signaling is essential for activation of protein
synthesis by Gq protein-coupled receptor agonists in adult cardiomyocytes. Circ Res
91(9), 821829.
Wang, Y., Huang, S., Sah, V. P., Ross, J., Jr., Brown, J. H., Han, J., et al. (1998a). Cardiac
muscle cell hypertrophy and apoptosis induced by distinct members of the p38
mitogen-activated protein kinase family. J Biol Chem 273(4), 21612168.
Wang, Y., Su, B., Sah, V. P., Brown, J. H., Han, J., & Chien, K. R. (1998b). Cardiac
hypertrophy induced by mitogen-activated protein kinase kinase 7, a specic
activator for c-Jun NH2-terminal kinase in ventricular muscle cells. J Biol Chem
273(10), 54235426.
Weber, K. T. (1997). Monitoring tissue repair and brosis froma distance. Circulation 96(8),
24882492.
Weber, K. T., & Brilla, C. G. (1991). Pathological hypertrophy and cardiac interstitium.
Fibrosis and reninangiotensinaldosterone system. Circulation 83(6),
18491865.
Weber, K. T., Brilla, C. G., & Janicki, J. S. (1993). Myocardial brosis: functional
signicance and regulatory factors. Cardiovasc Res 27(3), 341348.
Weinberg, E. O., Thienelt, C. D., Katz, S. E., Bartunek, J., Tajima, M., Rohrbach, S., et al.
(1999). Gender differences in molecular remodeling in pressure overload
hypertrophy. J Am Coll Cardiol 34(1), 264273.
Weiss, R. E., Korcarz, C., Chassande, O., Cua, K., Sadow, P. M., Koo, E., et al. (2002).
Thyroid hormone and cardiac function in mice decient in thyroid hormone
receptor-alpha or -beta: an echocardiograph study. Am J Physiol Endocrinol Metab
283(3), E428435.
Welch, S., Plank, D., Witt, S., Glascock, B., Schaefer, E., Chimenti, S., et al. (2002). Cardiac-
specic IGF-1 expression attenuates dilated cardiomyopathy in tropomodulin-
overexpressing transgenic mice. Circ Res 90(6), 641648.
Wencker, D., Chandra, M., Nguyen, K., Miao, W., Garantziotis, S., Factor, S. M., et al.
(2003). A mechanistic role for cardiac myocyte apoptosis in heart failure. J Clin
Invest 111(10), 14971504.
Wenger, N. K. (2002). Coronary heart disease and women: magnitude of the problem.
Cardiol Rev 10(4), 211213.
Wettschureck, N., Rutten, H., Zywietz, A., Gehring, D., Wilkie, T. M., Chen, J., et al. (2001).
Absence of pressure overload induced myocardial hypertrophy after conditional
inactivation of Galphaq/Galpha11 in cardiomyocytes. Nat Med 7(11), 12361240.
Widmann, C., Gibson, S., Jarpe, M. B., & Johnson, G. L. (1999). Mitogen-activated protein
kinase: conservationof a three-kinase module fromyeast tohuman. Physiol Rev 79(1),
143180.
Wikman-Coffelt, J., Parmley, W. W., & Mason, D. T. (1979). The cardiac hypertrophy
process. Analyses of factors determining pathological vs. physiological develop-
ment. Circ Res 45(6), 697707.
Wilkins, B. J., De Windt, L. J., Bueno, O. F., Braz, J. C., Glascock, B. J., Kimball, T. F., et al. (2002).
Targeted disruptionof NFATc3, but not NFATc4, reveals anintrinsic defect incalcineurin-
mediated cardiac hypertrophic growth. Mol Cell Biol 22(21), 76037613.
Wilkins, B. J., Dai, Y. S., Bueno, O. F., Parsons, S. A., Xu, J., Plank, D. M., et al. (2004).
Calcineurin/NFAT coupling participates in pathological, but not physiological,
cardiac hypertrophy. Circ Res 94(1), 110118.
Wilson, K. P., Fitzgibbon, M. J., Caron, P. R., Grifth, J. P., Chen, W., McCaffrey, P. G., et al.
(1996). Crystal structure of p38 mitogen-activated protein kinase. J Biol Chem 271
(44), 2769627700.
Wollert, K. C., Taga, T., Saito, M., Narazaki, M., Kishimoto, T., Glembotski, C. C., et al.
(1996). Cardiotrophin-1 activates a distinct form of cardiac muscle cell hypertro-
phy. Assembly of sarcomeric units in series VIA gp130/leukemia inhibitory factor
receptor-dependent pathways. J Biol Chem 271(16), 95359545.
Woodcock, E. A., Du, X. J., Reichelt, M. E., & Graham, R. M. (2008). Cardiac alpha 1-
adrenergic drive in pathological remodelling. Cardiovasc Res 77(3), 452462.
Woods, R. L. (2004). Cardioprotective functions of atrial natriuretic peptide and B-type
natriuretic peptide: a brief review. Clin Exp Pharmacol Physiol 31(11), 791794.
Wu, G., Yussman, M. G., Barrett, T. J., Hahn, H. S., Osinska, H., Hilliard, G. M., et al. (2001).
Increased myocardial Rab GTPase expression: a consequence and cause of
cardiomyopathy. Circ Res 89(12), 11301137.
Xiang, Y., & Kobilka, B. K. (2003). Myocyte adrenoceptor signaling pathways. Science
300(5625), 15301532.
Xin, H. B., Senbonmatsu, T., Cheng, D. S., Wang, Y. X., Copello, J. A., Ji, G. J., et al. (2002).
Oestrogen protects FKBP12.6 null mice from cardiac hypertrophy. Nature 416
(6878), 334338.
Xu, Y., Arenas, I. A., Armstrong, S. J., & Davidge, S. T. (2003). Estrogen modulation of left
ventricular remodeling in the aged heart. Cardiovasc Res 57(2), 388394.
Yamamoto, S., Yang, G., Zablocki, D., Liu, J., Hong, C., Kim, S. J., et al. (2003). Activation of
Mst1 causes dilated cardiomyopathy by stimulating apoptosis without compen-
satory ventricular myocyte hypertrophy. J Clin Invest 111(10), 14631474.
Yamazaki, T., & Yazaki, Y. (1997). Is there major involvement of the reninangiotensin
system in cardiac hypertrophy? Circ Res 81(5), 639642.
Yamazaki, T., Tobe, K., Hoh, E., Maemura, K., Kaida, T., Komuro, I., et al. (1993).
Mechanical loading activates mitogen-activated protein kinase and S6 peptide
kinase in cultured rat cardiac myocytes. J Biol Chem 268(16), 1206912076.
Yamazaki, T., Komuro, I., & Yazaki, Y. (1999). Role of the reninangiotensin system in
cardiac hypertrophy. Am J Cardiol 83(12A), 53H57H.
Yang, J., Moravec, C. S., Sussman, M. A., DiPaola, N. R., Fu, D., Hawthorn, L., et al. (2000).
Decreased SLIM1 expression and increased gelsolin expression in failing human hearts
measured by high-density oligonucleotide arrays. Circulation 102(25), 30463052.
Yayama, K., Horii, M., Hiyoshi, H., Takano, M., Okamoto, H., Kagota, S., et al. (2004). Up-
regulation of angiotensin II type 2 receptor in rat thoracic aorta by pressure-
overload. J Pharmacol Exp Ther 308(2), 736743.
Yeh, J. K., Aloia, J. F., Chen, M., Ling, N., Koo, H. C., & Millard, W. J. (1994). Effect of growth
hormone administration and treadmill exercise on serum and skeletal IGF-I in rats.
Am J Physiol 266(1 Pt 1), E129135.
Zak, R. (1974). Development and proliferative capacity of cardiac muscle cells. Circ Res
35(suppl II), 1726.
Zak, R. (1984). Growth of the heart in health and disease. New York, Raven Press: 124,
131185, 381420
Zannad, F., Briancon, S., Juilliere, Y., Mertes, P. M., Villemot, J. P., Alla, F., et al. (1999).
Incidence, clinical and etiologic features, and outcomes of advanced chronic heart
failure: the EPICAL Study. Epidemiologie de l'Insufsance Cardiaque Avancee en
Lorraine. J Am Coll Cardiol 33(3), 734742.
Zeisberg, E. M., Tarnavski, O., Zeisberg, M., Dorfman, A. L., McMullen, J. R., Gustafsson, E.,
et al. (2007). Endothelial-to-mesenchymal transition contributes to cardiac
brosis. Nat Med 13(8), 952961.
226 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227
Zhai, P., Gao, S., Holle, E., Yu, X., Yatani, A., Wagner, T., et al. (2007). Glycogen synthase
kinase-3alpha reduces cardiac growth and pressure overload-induced cardiac
hypertrophy by inhibition of extracellular signal-regulated kinases. J Biol Chem 282
(45), 3318133191.
Zhang, C. L., McKinsey, T. A., Chang, S., Antos, C. L., Hill, J. A., & Olson, E. N. (2002a). Class
II histone deacetylases act as signal-responsive repressors of cardiac hypertrophy.
Cell 110(4), 479488.
Zhang, T., Johnson, E. N., Gu, Y., Morissette, M. R., Sah, V. P., Gigena, M. S., et al. (2002b).
The cardiac-specic nuclear delta(B) isoform of Ca2+/calmodulin-dependent
protein kinase II induces hypertrophy and dilated cardiomyopathy associated with
increased protein phosphatase 2A activity. J Biol Chem 277(2), 12611267.
Zhang, S., Weinheimer, C., Courtois, M., Kovacs, A., Zhang, C. E., Cheng, A. M., et al.
(2003a). The role of the Grb2-p38 MAPK signaling pathway in cardiac hypertrophy
and brosis. J Clin Invest 111(6), 833841.
Zhang, T., Maier, L. S., Dalton, N. D., Miyamoto, S., Ross, J., Jr., Bers, D. M., et al. (2003b).
The deltaC isoform of CaMKII is activated in cardiac hypertrophy and induces
dilated cardiomyopathy and heart failure. Circ Res 92(8), 912919.
Zhang, R., Khoo, M. S., Wu, Y., Yang, Y., Grueter, C. E., Ni, G., et al. (2005). Calmodulin
kinase II inhibition protects against structural heart disease. Nat Med 11(4),
409417.
Zhu, Y. C., Zhu, Y. Z., Gohlke, P., Stauss, H. M., & Unger, T. (1997). Effects of angiotensin-
converting enzyme inhibition and angiotensin II AT1 receptor antagonism on
cardiac parameters in left ventricular hypertrophy. Am J Cardiol 80(3A),
110A117A.
Zimmerman, B. G., & Dunham, E. W. (1997). Tissue reninangiotensin system: a site of
drug action? Annu Rev Pharmacol Toxicol 37, 5369.
Zipes, D. P., Libby, P., Bonow, R. O., & Braunwald, E. (2005). Braunwald's Heart Disease: a
Textbook Of Cardiovascular Medicine Seventh Edition. Elsevier Saunders:
Philadelphia.
Zou, Y., Hiroi, Y., Uozumi, H., Takimoto, E., Toko, H., Zhu, W., et al. (2001). Calcineurin
plays a critical role in the development of pressure overload-induced cardiac
hypertrophy. Circulation 104(1), 97101.
227 B.C. Bernardo et al. / Pharmacology & Therapeutics 128 (2010) 191227

Вам также может понравиться