You are on page 1of 19

Exercise in Cardiovascular Disease

Cardiovascular Effects of Exercise Training


Molecular Mechanisms
Stephan Gielen, MD; Gerhard Schuler, MD; Volker Adams, PhD

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

n the natural habitat of our ancestors, physical activity


was not a preventive intervention but a matter of
survival. In this hostile environment with scarce food and
ubiquitous dangers, human genes were selected to optimize aerobic metabolic pathways and conserve energy for
potential future famines.1 Cardiac and vascular functions
were continuously challenged by intermittent bouts of
high-intensity physical activity and adapted to meet the
metabolic demands of the working skeletal muscle under
these conditions.
When speaking about molecular cardiovascular effects of
exercise, we should keep in mind that most of the changes
from baseline are probably a return to normal values. The
statistical average of physical activity in Western societies is
so much below the levels normal for our genetic background
that sedentary lifestyle in combination with excess food
intake has surpassed smoking as the No. 1 preventable cause
of death in the United States.2
Physical activity has been shown to have beneficial
effects on glucose metabolism, skeletal muscle function,
ventilator muscle strength, bone stability, locomotor coordination, psychological well-being, and other organ functions. However, in the context of this review, we will focus
entirely on important molecular effects on the cardiovascular system. The aim of this review is to provide a
birds-eye view on what is known and unknown about the
physiological and biochemical mechanisms involved in
mediating exercise-induced cardiovascular effects. The
resulting map is surprisingly detailed in some areas (ie,
endothelial function), whereas other areas, such as direct
cardiac training effects in heart failure, are still incompletely understood.
For practical purposes, we have decided to use primarily an
anatomic approach to present key data on exercise effects on
cardiac and vascular function. For the cardiac effects, the left
ventricle and the cardiac valves will be described separately;
for the vascular effects, we will follow the arterial vascular
tree, addressing changes in the aorta, the large conduit
arteries, the resistance vessels, and the microcirculation
before turning our attention toward the venous and the
pulmonary circulation (Figure 1).

Cardiac Effects of Exercise


Left Ventricular Myocardium and
Ventricular Arrhythmias
The maintenance of left ventricular (LV) mass and function
depends on regular exercise. Prolonged periods of physical
inactivity, as studied in bed rest trials, lead to significant
reductions in LV mass and impaired cardiac compliance,
resulting in reduced upright stroke volume and orthostatic
intolerance.3 In contrast, a group of bed rest subjects randomized to regular supine lower-body negative pressure treadmill
exercise showed an increase in LV mass and a preserved LV
stoke volume.4 In previously sedentary healthy subjects, a
12-week moderate exercise program induced a mild cardiac
hypertrophic response as measured by cardiac magnetic
resonance imaging.5 These findings highlight the plasticity of
LV mass and function in relation to the current level of
physical activity.
Normal LV Contractile Function
In individuals with normal LV function, exercise-related
cardiac effects may be subdivided into 3 entities: (1) prevention of cardiac pathologies associated with aging; (2) cardiac
adaptation to strenuous regular exercise training resulting in
physiological cardiac hypertrophy commonly known as athletes heart; and (3) prevention of impaired systolic and/or
diastolic cardiac function associated with short-term challenges (eg, ischemia/reperfusion [I/R], doxorubicin).
Prevention of Age-Related Cardiac Pathologies
Oxidative Stress. Theoretically, tissues with fewer cell divisions are more susceptible to cumulative damage by reactive
oxygen species (ROS) than tissues with high replication rates.
In animal models of aging, increased malondialdehyde levels
were observed in sedentary old rats. Although the expression
of antioxidative enzymes such as superoxide dismutase
(SOD) was unchanged,6,7 its enzymatic activity was reduced,8
resulting in a net decline of antioxidative protection. Regular
physical exercise seems to delay the accumulation of ROSmediated cell damage by improving the antioxidative protection in the myocardium. In trained old rats, Rinaldi et al7
found increased myocardial protein expression of MnSOD
and Cu/Zn SOD.

From the Department of Internal Medicine/Cardiology, University of Leipzig, Heart Center, Leipzig, Germany.
Correspondence to Volker Adams, PhD, Department of Internal Medicine/Cardiology, University of Leipzig, Heart Center, Strumpellstrae 39, 04289
Leipzig, Germany. E-mail adav@medizin.uni-leipzig.de
(Circulation. 2010;122:1221-1238.)
2010 American Heart Association, Inc.
Circulation is available at http://circ.ahajournals.org

DOI: 10.1161/CIRCULATIONAHA.110.939959

1221

1222

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Figure 1. Summary of exercise-mediated effects on different parts of the cardiovascular system. Max. indicates maximum; HFNEF,
heart failure with normal ejection fraction; ANP, atrial natriuretic peptide; and BNP, brain natriuretic peptide.

Heat Shock Proteins. Heat shock proteins (HSP) are highly


conserved chaperones that are involved in the folding of
newly synthesized or damaged proteins. The overexpression
of HSP70 in cardiomyocytes protects against ischemic damage and is associated with increased cell survival.9 Stressinduced synthesis of HSP70 is reduced in the aged cardiovascular system, consistent with a diminished stress
resistance.10 HSP70 and HSP27 were also elevated in trained
old rats versus sedentary old rats, indicating improved resistance to ischemic insults.7
Cardiomyocyte Apoptosis and Regeneration. Despite the
increase in LV mass generally observed with aging, the
number of cardiomyocytes declines by 30%.11 The exponential increase of cardiomyocyte death by apoptosis or
necrosis and the loss of cardiac regeneration by stem cells
have been discussed as key factors for this decline in cell
number.12,13 The increase in apoptosis may be a consequence
of decreased mitochondrial membrane stability and permeability transition pore formation, which leads to cytochrome c
release into the cytosol, resulting in activation of caspase-3
and -9.12 Exercise training seems to be highly protective
against cardiomyocyte apoptosis by modulating proapoptotic
and antiapoptotic genes.14
Human Functional Data. On the functional level in humans,
cardiac aging is associated with a significant increased wall
stress, decreased elasticity, impaired early LV diastolic relaxation, increased end-systolic LV volume, and reduced contractile reserve.15 Arbab-Zadeh et al16 were able to show that
lifelong physical activity prevents the development of diastolic LV dysfunction in older age. This finding also indicates

that so-called healthy aging is in fact primarily sedentary


aging and that the lack of physical activity may be more
relevant to the observed pathologies than aging by itself.
Cardiac Adaptation to Strenuous Exercise and
Cardiac Fatigue
Athletes Heart. In normal individuals, regular high-intensity
physical activity for 5 to 6 hours per week may result in
cardiac adaptations known as athletes heart, resulting in
compensatory myocardial hypertrophy, which is more eccentric in endurance training and more concentric in resistance
training.17 Endurance-trained athletes show mean LV end-diastolic diameters of 53.7 mm compared with 49.6 mm in
normal subjects17; however, 14% of 1309 athletes examined
by Pelliccia et al18 showed LV end-diastolic diameters between 66 and 70 mm. The diagnostic strategies to distinguish
normal LV hypertrophy from hypertrophic cardiomyopathy
are a matter of continuing discussions (reviewed by Fagard19
and Maron20).
LV physiological hypertrophy is caused by a proportional
increase in myocardial cell length and width without evidence
of myocardial hyperplasia in the majority of cases and is
mediated via increased cardiac insulin-like growth factor-1
(IGF-1) expression and activation of phophoinositide-3 kinase (PI3K) (p110).21 Transgenic mice with decreased
cardiac PI3K activity (dnPI3K mice)22 and mice lacking the
p85 subunit of PI3K23 show a reduced basal cardiac size and
an attenuation of exercise-induced myocardial hypertrophy.
Studies by Care` et al24 in exercised rats suggest that
downregulation of cardiac-specific microRNAs (miRNA,
miR-133, and miR-1) is regularly involved in exercise-

Gielen et al

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

induced cardiac hypertrophy. Reduction of miR-133 levels in


cardiac myocytes caused marked hypertrophy with increased
protein synthesis and fetal gene expression. MiR-1 overexpression, on the other hand, significantly reduced protein
synthesis and fetal gene expression.24 Its hypertrophic effects
are probably mediated via RhoA, Cdc42, and Whsc2, whose
RNA 3 untranslated regions comprise flanking nucleotides
matching miR-133.
The role of miR-1 in physiological hypertrophy was
further elucidated by Elia et al,25 who documented that IGF-1
and IGF-1 receptor are targets of miR-1 and that miR-1 and
IGF-1 are inversely correlated in cardiac hypertrophy.25IGF-1
activation, on the other hand, inhibits the miR-1 transcription
factor Foxo3a via an AKT-dependent pathway, further amplifying its hypertrophic effects.
Using a transgenic mouse model with overexpression of a
constitutively activated AKT (AKT-E40K), Catalucci et al26
confirmed that IGF-1/AKT also improved LV contractility by
phosphorylation of phospholamban at threonine. By translocating to the sarcoplasmatic reticulum, activated AKT can
mediate improved calcium handling by disinhibition of the
sarcoplasmic reticulum Ca2 ATPase pump (SERCA2a)
through phospholamban phosphorylation.
Cardiac plasticity in response to exercise training is
surprisingly rapid. In a rat model of high-intensity interval
treadmill running for 2, 4, 8, and 13 weeks followed by 2 or
4 weeks of detraining, Kemi and colleagues27 identified
cardiomyocyte cell length, diastolic relaxation, and calcium
decay as main factors for the training-induced increase in
VO2max by multiple regression analysis. These trainingrelated cardiac adaptations seem to be dependent on training
intensity.28
Cardiac Fatigue. Prolonged strenuous activity, as in the
Ironman Triathlon, can lead to transient reductions of LV
systolic and diastolic function.29 Several mechanisms are
involved in cardiac fatigue, including changes in preload
conditions, myocardial stunning, 1-receptor desensitization,
and altered cardiac autonomic regulation.30
Preload conditions were highlighted as a potential mechanism by Dawson et al, 31 who demonstrated the lack of
changes in LV function when central venous pressure was
well maintained. However, other researchers with different
training protocols found depressed LV contractile function
unrelated to preload conditions, which confirms the relevance
of intrinsic myocardial damage.32 In a dog model, Vatner and
Hittinger33 demonstrated that prolonged strenuous exercise
can indeed lead to ischemic myocardial dysfunction and
postischemic myocardial stunning in the absence of epicardial coronary stenoses.
Prolonged strenuous exercise results in a 5-fold increase in
circulating catecholamines34 with consecutive desensitization
of 1-adrenoreceptors.35 This results in a blunted chronotropic and inotropic response to dobutamine.32

Cardiovascular Effects of Exercise Training

1223

approach to I/R protection to exercise is to differentiate


between the mechanisms that are involved in I/R protection
and those that are essential. Mechanisms such as collateral
formation,38 elevated HSP,39,40 increased myocardial cyclooxygenase-2 expression,41 and higher levels of endoplasmic
reticulum stress proteins have all been shown to be not
essential for exercise-mediated I/R protection.42 Therefore,
we will focus on mechanisms found to be a condition sine qua
non for I/R protection. Among these are improved antioxidative protection, changes in mitochondrial metabolism and
protein expression, increased expression of sarcolemmal/
mitochondrial KATP channels, and attenuation of I/Rinduced calpain activation (Figure 2).
Antioxidative Protection. Production of ROS, accumulation
of hydrogen ions, and generation of reactive nitrogen species
play a major role in the complex pathophysiology of I/R
injury,43,44 and the quantity of ROS production critically
depends on the duration of anoxia and reoxygenation.44 The
primary source of ROS in I/R situations seem to be the
mitochondria.45 In animal models, exercise training protects
cardiac myocytes against I/R-induced oxidative stress and the
mitochondria against I/R-associated structural damage.46,47
Several mechanisms for training-induced cardioprotection
are under discussion (Figure 2). There is broad consensus that
increased antioxidative protection by training-induced upregulation of key antioxidative enzymes such as MnSOD,
glutathione peroxidase, and catalase plays an important
role48 50; however, myocardial MnSOD is only increased at
high training intensities. Recent studies have confirmed that
MnSOD is essential for full protection against I/R-induced
myocardial infarction (ie, ischemia 20 minutes) and related
ventricular arrhythmias; however, protection against myocardial stunning seems to work even in the absence of MnSOD.51
Changes in Mitochondrial Metabolism and Protein Expression. Cardiac mitochondria seem to contribute to exerciseinduced protection against I/R injury in several ways, as
follows: (1) Exercise training results in reduced generation of
ROS by cardiac mitochondria52,53; (2) mitochondria isolated
from the myocardium of exercised animals are more resistant
to calcium-induced opening of the mitochondrial permeability transition pore,52 a major proapoptotic stimulus; and (3)
exercise induces a downregulation of mitochondrial monoamine oxidase-A,54 which seems to be a major source of
oxidative stress by H2O2 generation.55 Monoamine oxidase-A
knockout animals are protected from I/R-induced myocardial
damage, highlighting the importance of this pathway for I/R
injury.55

Prevention of Functional Decline After Exposure


to Ischemia
I/R injury is a classic model of an acute cardiac injury with
great clinical relevance in the setting of myocardial infarction
with interventional reperfusion therapy. Human epidemiological data indicate that regular endurance exercise reduces the
risk of death during clinical I/R injury.36

Role of Sarcolemmal/Mitochondrial KATP Channels. Recently,


the sarcolemmal KATP channel has been described as a
potential mechanism for training-induced I/R protection.
Opening the sarcolemmal KATP channel accelerates cardiomyocyte repolarization by increasing K outflow and shortens the action potential.56 As a result of the shorter action
potential, Ca2 overloading is prevented by reducing opening
rates of the L-type Ca2 channel.56

Key Mechanisms for Protection Against I/R Injury. As


summarized by Powers et al, 37 the problem of a mechanistic

Attenuation of I/R-Induced Calpain Activation. Calpain is a


calcium-dependent protease with 2 isoforms (calpain I [milli]

1224

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Figure 2. Exercise training has the potential to prevent myocardial damage related to I/R injury. Exercise training induces an upregulation of
key antioxidative enzymes such as MnSOD, glutathione peroxidase (GPX), and catalase. Opening the sarcolemmal (sarco) KATP channel
accelerates cardiomyocyte repolarization by increasing K outflow. As a result of the shorter action potential, Ca2 overloading is prevented
by reducing opening rates of the L-type Ca2 channel. Although not essential for I/R protection, increased levels of HSP72 were observed
after training. HSP72 can prevent apoptotic cell death, likely by interaction with caspase-3. Adapted from Powers et al 37 with kind permission of Springer Science and Business Media. Copyright 2008, Springer Science and Business Media

and calpain II [micro]), which are activated by prolonged


exposure to elevated cytosolic calcium levels and contribute
to myocardial I/R injury.57 French and colleagues58 showed
that calpain inhibition attenuates I/R-induced contractile
myocardial dysfunction. Exercise training reduces I/Rassociated calpain activation, most likely by improved antioxidative protection and prevention of I/R-induced SERCA2a
and phospholamban degradation.58
Thresholds and Time Course of I/R Protection
Training Intensity. In a rat model of 40 minutes per day of
treadmill exercise at 55% to 60% of VO2max for a total time
of 16 weeks, Starnes et al59 were unable to confirm a
significant level of I/R protection. This study therefore
provided evidence for a threshold effect (ie, that a minimum
intensity and duration of exercise have to be exceeded to
benefit from I/R protection). Lennon et al,60 on the other
hand, compared 2 training protocols at 55% and 75% of
VO2max for 60 minutes per day on 3 consecutive days and
did not find any difference in training-induced I/R protection between both groups, indicating that there may not be
a linear dose-response relation between training intensity
and cardioprotection.
Training Duration. Surprisingly, the protective effect is
similar after short-term (3 to 5 days) and long-term (weeks to
months) exercise training.40,46,61 In a trial specifically de-

signed to assess the loss of cardioprotective effects after


training cessation, Lennon et al62 were able to show a
persistence of protection against myocardial stunning for up
to 9 days, but cardioprotection was completely lost at 18 days
after exercise termination.
Systolic Heart Failure
Reduced exercise capacity and early fatigue in chronic heart
failure (CHF) are unrelated to the degree of LV systolic
dysfunction but are determined primarily by peripheral alterations with impaired endothelial function, reduced aerobic
oxidative metabolism in the skeletal muscle, skeletal muscle
atrophy, ventilatory dysfunction, and neurohumoral changes.63 In the context of this review, we will focus on the
cardiac effects of exercise only.
Training Effects on LV Function and Reverse Remodeling
In one of the first small prospective studies of aerobic
endurance training in CHF patients (n12), Sullivan et al64
confirmed that 4 to 6 months of training did not worsen LV
ejection fraction and tended to improve maximal cardiac
output. The extent of the cardiac changes did not, however,
explain the large 23% improvement in peak oxygen uptake so
that peripheral changes in limb perfusion and oxidative
metabolism must account for the larger part of the beneficial
symptomatic training effects.
The first larger prospective randomized study to actually
provide evidence for a training-induced reverse remodeling

Gielen et al

Cardiovascular Effects of Exercise Training

1225

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Figure 3. In systolic heart failure, endurance exercise can induce reverse remodeling with improvement of systolic and diastolic LV
function and reduction of LV end-diastolic diameters. It does so by interfering with the vicious cycle of CHF-associated skeletal muscle
hypoperfusion with increased local ROS and cytokine generation resulting in endothelial function and sympathetic activation, which further increases afterload and worsens LV function. The reverse remodeling is a consequence of peripheral training effects (eg, improved
antioxidative protection in the skeletal muscle, increased parasympathetic tone, improved endothelial function as a result of increased
NO production and bioavailability, and direct cardiac effects related to improved Ca handling, increased PI3K [p110] activity, reduction
of fibrosis, and cardiomyocyte apoptosis). NE indicates norepinephrine; AT II, angiotensin II; ecSOD, extracellular SOD; GPX, glutathione peroxidase; and ox. Phos, oxidative phosphorylation.

came from Hambrecht et al,65 who demonstrated that endurance training led to reverse LV remodeling with modest
improvements of ejection fraction from 30% to 35% and
reductions of LV end-diastolic diameter in a mixed population of subjects with ischemic and dilated cardiomyopathy.
These findings were corroborated by the Exercise in Left
Ventricular Dysfunction and Chronic Heart Failure study.66 A
recent study in our institution indicates that the reverse LV
remodeling occurs as early as 4 weeks after the initiation of
endurance training (unpublished data, 2010).
Mechanisms Explaining Reverse Remodeling in CHF
In the absence of myocardial biopsies for molecular analysis
of myocardial changes induced by training, most authors
interpreted this favorable training effect as secondary to
afterload reduction with reduced resting blood pressure due to
improved endothelial function.65 67 However, animal models
reveal that there are clear direct myocardial effects of training
that are related to signaling pathways of myocardial hypertrophy and fibrosis (Figure 3)
Phosphoinositide-3 Kinase (p110). On the molecular basis,
recent animal studies suggest that activation of PI3K (p110)
is involved in exercise-mediated cardioprotection: In mice
with dilated cardiomyopathy, both swim training and consti-

tutive activation of PI3K (p110, caPI3K) improved survival


by 20%.22 In an aortic banding model of pathological
hypertrophy, caPI3K mice showed a blunted increase in heart
size compared with wild-type mice.22 Levels of interstitial
fibrosis were reduced in caPI3K mice, highlighting the
potential of PI3K as an antifibrotic signal. Another important
downstream effect of PI3K is the protection against apoptotic
cardiomyocyte death via AKT.
Anabolic/Catabolic Balance in the Myocardium. Animal
studies in which an left anterior descending artery ligation
model was used demonstrated a significant upregulation of
components of the ubiquitin-proteasome system as well as
myostatin.68,69 Both of these factors were significantly reduced by exercise training over a period of 4 weeks.68,69
Calcium Handling. Alterations in calcium handling are also
associated with pathological hypertrophy and transition from
hypertrophy to failure: SERCA2a protein levels were reduced
in mouse and dog models of heart failure and were normalized by exercise training.70,71 In addition, exercise training
activates Ca2/calmodulin-dependent protein kinase II, leading to a hyperphosphorylation of phospholamban,72 which in
its phosphorylated form no longer inhibits SERCA2a. In
conjunction with an increased expression of Na-Ca2 ex-

1226

Circulation

September 21, 2010

changer,73 this leads to improved calcium cycling and finally


to better cardiomyocyte function. For more detailed information on exercise-induced improvements on the contractile
apparatus and calcium cycling, please see a recent review by
Kemi and Wisloff.74

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Neurohormonal Adaptations. An aerobic training program in


patients with CHF regularly reduces the resting heart rate,
which indicates a reduction in sympathoadrenergic drive.
This has also been confirmed for serum catecholamine levels:
Coats et al75 showed a 16% reduction of radiolabeled norepinephrine secretion after 8 weeks of training. This reduction in
adrenergic tone was accompanied by an increase in heart rate
variability. In addition to the reduction in circulating catecholamines, Braith et al76 described a 25% to 30% reduction
of angiotensin II, aldosterone, arginine vasopeptide, and atrial
natriuretic peptide after 4 months of walking training in
patients with CHF.
In a rat model of ischemic CHF, the beneficial training
effects on local neurohumoral balance were analyzed in the
noninfarcted LV myocardium. Xu and colleagues77 found a
significant reduction of myocardial angiotensin-converting
enzyme mRNA expression and AT1-receptor expression after
8 weeks of treadmill training. This finding is of special
importance given the fact that 90% of angiotensin II is
produced locally in the myocardium and implies that local
angiotensin II levels are significantly reduced by training.
This reduction also translates into reduced fibrogenesis, as
indicated by reduced tissue inhibitor of metalloproteinase-1
expression with unchanged matrix metalloproteinase
(MMP)-1 expression and reduced collagen volume fraction in
the exercised animals.77
Diastolic Heart Failure
Heart failure with normal ejection fraction is a comparatively
new disease category that includes patients who display the
typical clinical, physiological, and neurohormonal heart failure phenotype in the absence of reduced LV ejection fraction.78 Thus far, no specific pharmaceutical therapy for heart
failure with normal ejection fraction has been found. However, a number of clinical and experimental studies indicate a
protective and curative role of exercise training for this
disease entity.
Protective Effects
It is well established in sports physiology that endurance
training improves LV diastolic filling at rest and during
exercise.79 Endurance exercise interventions improve early
diastolic relaxation at all ages in healthy individuals, but in
the presence of an elevated atrial filling rate in elderly
patients, this could be normalized.80 A landmark study for the
preventive effects of lifelong physical activity against agerelated diastolic dysfunction was published by Arbab-Zadeh
et al in 2004.16 They were able to show with invasive
measurement of LV pressure-volume relations that senior
athletes did not exhibit any degree of diastolic dysfunction
compared with young healthy sedentary controls.
Curative Effects
Patients after myocardial infarction and with preexistent
abnormal relaxation patterns benefit from endurance training,

as shown by Yu et al.81 Although data on training effects in


patients with diabetes mellitus derived from small studies are
still controversial,82,83 data on patients with established heart
failure with normal ejection fraction are convincingly
positive.84
CHF with systolic dysfunction is associated primarily with
significant diastolic dysfunction. Prospective randomized
training studies confirmed that as early as after 4 weeks of
endurance training, diastolic dysfunction is significantly improved with reduced E/E ratio (unpublished data, 2010).
Molecular Mechanisms
In diastolic heart failure, calcium homeostasis is affected by
a reduced expression of SERCA2a,85 which is the key player
for calcium reuptake into the sarcoplasmic reticulum and is
regulated in its activity by phospholamban and sarcolipin and
increased diastolic calcium leakage via the ryanodine receptors.86,87 Phospholamban affects both the calcium affinity of
SERCA and its calcium reuptake rates. Animal studies in
healthy rats indicate that endurance training can increase both
SERCA and phospholamban expression,73 a finding that was
confirmed in a mouse model of sympathetic hyperactivityinduced heart failure (congenic 2A/2C-adrenoceptor knockout mice).88 No data are available with regard to trainingassociated effects on calcium leakage. On the structural level,
12 weeks of treadmill training in old Fischer 344 rats did not
affect collagen volume fraction but significantly reduced
collagen cross-links by advanced glycation end-products.89

Cardiac Valves
Aortic Valve Sclerosis/Stenosis
Currently, no effective therapy to prevent calcified aortic
valve disease is established. Calcified aortic valve disease
confers significant morbidity and mortality as the severity of
disease progresses.90
Degenerative calcified aortic valve disease and atherosclerosis share similar mechanisms (ie, clinical risk factors and
histopathological features). Pathological examinations of diseased aortic valves showed areas of subendothelial thickening
with the disruption of the basement membrane, accumulation
of inflammatory infiltrates, and calcium deposits.91,92 It is
well accepted that regular physical exercise prevents the
progression of atherosclerosis by modulating endothelial
function or oxidative stress.93,94 Using low-density lipoprotein receptor knockout mice, we recently demonstrated that
regular exercise training as primary prevention prevents the
development of aortic valve sclerosis.95 It has been proposed
that apoptosis of endothelial cells and myofibroblasts in the
valve,96 elastin degradation and collagen synthesis by MMPs
and/or cathepsin,97 increased oxidative stress,98 and enhanced
osteogenesis99 contribute to aortic valve degeneration. In
consequence, procalcific gene expression such as bone morphogenetic protein 2 (BMP2), CBFA1/Runx2 (runt-related
transcription factor 2), and osteocalcin was amplified.98 The
activation of BMP2 and CBFA1/Runx2, particularly in the
noncalcified valve tissue, reflects an early stage of transdifferentiation of myofibroblast to a more osteoblast-like
phenotype.100

Gielen et al
In the low-density lipoprotein receptor knockout mouse
model, regular exercise attenuated the deleterious elevation of
oxidative stress and reduced expression of BMP2, Runx2,
and osteoblast differentiation marker. Additionally, the disruption of the aortic valve endothelium was prevented.95

Cardiovascular Effects of Exercise Training

1227

metabolic control is dominant in small resistance arteries, in


which the myogenic response to increased perfusion pressure
is also most pronounced. Although the effects of exercise on
the microcirculation and the capillary bed have been studied
extensively, little is known about the training-induced mechanism of vascular changes in the venous circulation.

Vascular Effects of Exercise

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

It has been said that one is as old as ones arteries. In view


of the supreme importance of endothelium in arterial function, I should like to modify this statement by saying that one
is as old as ones endothelium.101 Perhaps Rudolf Altschul
himself would be surprised to see to what extent his prophetic
words have become scientific consensus half a century later.
There are several reasons why endothelial research became so
significant for todays mechanistic concepts of exercisemediated cardiovascular effects: (1) Endothelial dysfunction
has been found to be a condition sine qua non for atherogenesis (ie, an obligatory initial step in early atherosclerosis).102
Thus, prevention of endothelial dysfunction will necessarily
prevent atherosclerosis development. (2) The presence of
endothelial dysfunction predicts future cardiovascular
events.103,104 Hence, endothelial function measurements were
proposed as a surrogate end point in clinical research.105 (3)
Endothelial function can be assessed by a large variety of
methods in both animal and human studies. Surgical/invasive
methods include organ bath measurements of arterial rings
harvested in animals or during aortocoronary bypass surgery
in humans and catheter-based assessment of acetylcholineinduced vasodilatation with angiography of the target vessel.
Noninvasive methods such as brachial/radial ultrasound,
magnetic resonance imaging, and venous occlusion plethysmography permit serial studies comparing endothelial function before and after an exercise intervention. These classic
methodologies are increasingly supplemented by commercial
semiautomatic devices using, for example, digital pulse
amplitude tonometry to measure microvascular function.106
It is beyond the scope of this article to provide in-depth
review of endothelial function in health and disease. We will
focus on 2 important aspects of vascular research in exercise
physiology: (1) the heterogeneity of molecular mechanisms
involved in vasomotor function along the vascular bed and
(2) the molecular mechanisms for exercise-mediated improvements of vascular endothelial function with special
emphasis on clinical studies in overt heart disease (coronary
artery disease and heart failure).

Differential Effects on Different Parts of the


Vascular Bed
The vasomotor responsiveness is not uniformly distributed
along the arterial tree. Four basic mechanisms have been
identified to regulate vascular tone in response to local
metabolic demand in the downstream vascular territory: (1)
endothelium-mediated flow-induced vasodilatation, (2) myogenic control, (3) metabolic vasodilatation, and (4) sympathetic control.107,108 The relative contribution of these mechanisms to vasodilatation is different in different
microdomains of the coronary vascular tree. Endotheliumdependent vasodilatation is most relevant in small conduit
and resistance arteries down to 150 m of diameter108;

Aorta
Although the aorta is a standard vessel to quantify changes in
endothelial function in animal experiments involving small
rodents, it is rarely a target vessel in human vascular research.
The composition of the extracellular matrix has changed
through the evolution of species and is influenced by pulse
pressure amplitude and wall tension. In fact, there seems to be
a universal elastic modulus that determines the relative
proportion of elastic lamellae, collagen fibers, and smooth
muscle cells in the aortic wall.109 Aging, coronary artery
disease, and myocardial infarction are often accompanied by
increased aortic stiffness and reduced compliance.110,111
Among healthy individuals, high-intensity exercise led to
vascular remodeling with up to 51% enlarged brachial conduit artery area and reduced distal aortic cross-sectional areas
(up to 8%), whereas aortic distensibility remained unchanged.112 However, endurance and strength training have
opposite effects on aortic compliance and vascular stiffness;
in an elegant study, Otsuki et al113 measured plasma
endothelin-1, nitric oxide (NO), and arterial stiffness in
young, healthy endurance-trained versus strength-trained
men. They demonstrated that aortic pulse-wave velocity as an
established index of vascular stiffness was significantly
increased in strength athletes and reduced in endurance
athletes versus healthy controls.113 Strength athletes displayed elevated endothelin-1 levels that correlated with aortic
pulse-wave velocity. Reductions of aortic stiffness after
endurance training were confirmed in patients with hypertension114 and coronary artery disease.115
Large Arterial Conduit Vessels
Conduit vessels 2 to 4 mm in diameter are the preferred target
vessels in clinical research because they may be assessed
readily by either noninvasive of invasive methods. After the
first description of coronary endothelial dysfunction,116 it was
quickly recognized that coronary endothelial function
could serve as a stress index integrating the cardiovascular
effects of risk factors and may provide pivotal diagnostic
and prognostic information in patients with coronary heart
disease.103
Exercise and Conduit Vessel Vasomotor Function in
Coronary Artery Disease
The clinical effects of endurance exercise training on coronary endothelial function were first studied in humans by
Hambrecht et al117: A 4-week endurance training program
was effective in attenuating the paradoxical arterial vasoconstriction in epicardial conduit vessels by 54% and increased
average peak flow velocity by 78% in response to intracoronary acetylcholine infusion. The improvements in coronary endothelial function were partially lost during a consecutive 5-month home-based endurance training program at

1228

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Figure 4. The majority of exercise effects on the vascular endothelium are mediated by intermittent increases of laminar shear stress
associated with increased cardiac output during physical exertion. On the luminal side of the endothelial cells, direct signaling can
occur through deformation of the glycocalyx. This shear stressinduced signal can activate calcium ion channels, phospholipase activity leading to calcium signaling, prostaglandin I2 (PGI2) release, and cAMP-mediated smooth muscle cell relaxation. VEGF receptor 2
(VEGFR2) located at the luminal surface can associate with vascular endothelial cadherin, [beta]-catenin, and PI3K to phosphorylate Akt
and induce AKT-mediated eNOS phosphorylation, which leads to higher NO production. Mechanical forces are also transmitted by the
cytoskeleton to adhesion sites, where transmembrane integrins bound to the extracellular matrix serve as a focus for deformation. This
integrin-mediated pathway leads to the activation of Rasmitogen-activated protein kinase activity. In turn, Ras releases nuclear
factor-B (NFB) from its cytosolic inhibitor and thus enables its translocation to the nucleus, where it binds to the promoters of multiple target genes, including the eNOS gene. Increased NO synthesis in response to shear stress induces extracellular SOD (ecSOD) in a
feed-forward fashion to inhibit the degradation of NO by ROS with consecutive formation of the toxic peroxynitrite. Akt indicates protein kinase B; PECAM-1, platelet endothelial cell adhesion molecule-1; Ras, small GTPase; ONOO, peroxynitrite; and AA, Arachidonic
acid. Reprinted from Davies et al 121 with permission of the publisher. Copyright 2008, Cambridge University Press.

lower intensity, and changes were related to daily training


durations.118
In a subsequent study examining training effects on the
peripheral conduit artery function, Gokce et al119 measured
brachial and femoral artery endothelial function in patients
with coronary artery disease before and after a 10-week leg
exercise program. They documented significant improvement
in endothelium-dependent, flow-mediated dilation in conduit
arteries of the leg but not the arm and hypothesized that
training effects could be limited to the trained limb. This may,
however, also depend on the distribution of endothelial
dysfunction and the intensity of the training program because
Linke et al120 were able to confirm improved radial artery
endothelial function after pure bicycle ergometer training in
patients with CHF.
Mechanosensors in the Vascular Wall
Arteries are exposed to 2 main forces: radial strain as a result
of the blood pressure wave that is propagated along the
arterial tree and laminar shear stress as a result of the

frictional forces caused by antegrade blood flow. Whereas


increased radial strain (eg, as a result of arterial hypertension)
increases the atherogenic risk, laminar shear stress as a result
of pulsatile continuous blood flow is a potent survival signal
for endothelial cells.
On the molecular level, laminar shear stress affects multiple signaling pathways121 (Figure 4), which include the PI3K,
extracellular signal-regulated kinase 5 (also known as
mitogen-activated protein kinase 7), and NO pathways.122
Endothelial cells must therefore express specific mechanotransducers that convert physical stress into biochemical
signals. The cytoskeleton plays an important part in the
sensing of shear stress, and high strain is observed at the
luminal and basal membrane. In addition, cellular adhesion
proteins such as platelet endothelial cell adhesion molecule-1,
vascular endothelial cadherin, and the transmembrane tyrosine kinase vascular endothelial growth factor receptor-2
(KDR) were found to mediate biochemical responses to
flow. Vascular endothelial growth factor receptor-2 activates

Gielen et al
PI3K, which is essential for AKT-mediated phosphorylation
and activation of endothelial NO synthase (eNOS).123 In
contrast to laminar shear stress, the atherogenic flow patterns
include turbulent or disturbed flow, low flow, gradients, and
flow reversal. These conditions are associated with high rates
of endothelial cell proliferation and apoptosis, increased
production of ROS, and increased expression of inflammatory markers inducing a preatherosclerotic vascular phenotype.122 In the presence of additional atherogenic risk factors,
such as hypertension, dyslipidemia, and diabetes mellitus,
atherogenesis is significantly accelerated.

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

NO Generation in the Endothelium


A key component of intact endothelial function is a functional
eNOS. Located in the luminal endothelial cell membrane, it
produces NO from L-arginine. By diffusion, the short-lived
NO reaches the vascular smooth muscle cells in the media
and causes relaxation via cyclic guanosine monophosphatedependent pathways. Potential mechanisms of endothelial
dysfunction are (1) reduced eNOS substrate and cofactor
availability (eg, reduction of L-arginine levels, increase of the
competitive eNOS inhibitor asymmetrical dimethylarginine
[ADMA])124; (2) a reduced quantity and activity of eNOS and
inactivation of NO by ROS125; and (3) vascular regeneration
by endothelial progenitor cells (EPCs).126
Exercise and eNOS Substrate/Cofactor Availability. Another
enzyme also responsible for the generation of ROS in the
vascular is NO synthase. Under a number of pathological
conditions, the enzymatic reduction of molecular oxygen by
eNOS is no longer coupled to L-arginine oxidation, resulting
in production of superoxide rather than NO.127 A number of
mechanisms have been reported to contribute to eNOS
uncoupling, including tetrahydrobiopterin (BH4) deficiency,
shortage of L-arginine or HSP90, eNOS dephosphorylation on
threonine residue 495, or elevated ADMA levels.127 With
respect to the BH4 concentration, cell culture studies provided
the first evidence that increased shear stress dramatically
increases BH4 levels.128 This result was confirmed in an
aortocaval fistula model in the rat, wherein aortic blood flow
is enhanced proximal but decreased distal to the fistula,129 as
well as in a rat exercise training model.130 In addition, for
circulating levels of ADMA, a beneficial effect of exercise
training was documented.131
Exercise, eNOS Activity, and Oxidative Stress. To assess
molecular changes in NO production in human arterial
conduit vessels, we randomized patients with coronary artery
disease scheduled for elective coronary artery bypass graft
surgery to 4 weeks of endurance training or control and
harvested left internal mammary artery tissue not needed for
bypass grafting.132 Patients in the training group had 2-fold
higher eNOS protein expression and 4-fold higher eNOS
Ser1177 phosphorylation levels in left internal mammary
artery endothelium. A linear correlation was confirmed between Akt phosphorylation and phospho-eNOS levels and
between phospho-eNOS and Delta mean peak blood flow
velocity.
NO bioavailability is also influenced by ROS-mediated
breakdown. In animal and human studies, exercise train-

Cardiovascular Effects of Exercise Training

1229

ing programs resulted in an increased vascular expression


of antioxidative enzymes, such as SOD, catalase, and
glutathione peroxidase133135 and a reduced expression of
ROS-generating enzymes such as nicotinamide adenine
dinucleotide phosphate (NAD[P]H) oxidase and xanthine
oxidase.94,136 Nicotinamide adenine dinucleotide phosphate (NAD[P]H) oxidase expression and activity are
regulated mainly by the dominant influence of angiotensin
II type 1 receptor activation by inducing a rapid rac1
translocation to the cell membrane.137 In addition, a
4-week exercise training program in patients with coronary
artery disease resulted in a significantly lower expression
of angiotensin II type 1 receptor compared with a sedentary control group.94
Exercise and EPCs.
The first description of postnatal vasculogenesis and the
characterization of EPCs derived from adult bone marrow
changed our concept of vascular plasticity in health and
disease.138 In addition to EPCs, mesenchymal stem cells also
possess the potential for vascular regeneration.139 Both mesenchymal stem cells and EPCs are mobilized from the bone
marrow in response to ischemia, exercise, and neurohormonal
factors (eg, vascular endothelial growth factor [VEGF],
placental growth factor) and critically contribute to maintaining the integrity of the endothelial cell layer. EPC number and
function correlate with the number of cardiovascular risk
factors and disease severity and predict cardiovascular events
and death from cardiovascular causes.140
It is a matter of continuing debate whether training-induced
flow changes or ischemia induction is required to stimulate
EPC release. In 2 clinical studies, we demonstrated that an
increase in circulating EPCs was only achieved in response to
exercise-induced ischemia, whereas matrigel assays indicated
improved EPC function in nonischemic training, as
well.141,142 Other authors showed exercise-related EPC increase also in the absence of ischemia.143,144 Prolonged
strenuous exercise (half-marathon, marathon, or spartathlon),
however, showed no change or a significant increase in
circulating progenitor cells.145148
The principal mechanism of EPC mobilization from the
bone marrow seems to depend on the activation of eNOS in
the presence of several mobilizing factors such as VEGF149 or
placental growth factor.150 Gene-targeting studies using either
MMP-2151 or MMP-9152 knockout mice demonstrated that the
presence of MMPs is crucial in ischemia-induced mobilization of EPCs and hence neovascularization. At the molecular/
cellular level, the following scenario for the mobilization is
proposed (Figure 5): Under steady state conditions, progenitor cells reside in a niche of the bone marrow, bound to
stroma cells via adhesion molecules such as vascular cell
adhesion molecule/very late antigen-4.153 Signal-induced upregulation of MMP-9 results in a release of sKitL, conferring
signals that enhance mobility of progenitor cells into a
vascular-enriched niche favoring liberalization of the cells
into the circulation.154 How can we explain the exerciseinduced mobilization of EPCs? In several studies, it was
demonstrated that exercise increases the concentration of
NO,132 which in turn can activate MMP-9 in bone marrow,155

1230

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Figure 5. A model for mobilization of circulating progenitor cells from the bone marrow by increased shear stress induced by exercise.
Increased laminar shear stress results in an activation of eNOS, resulting in an increase of NO. Subsequently, MMP-9 and MMP-2 are
activated, resulting in the release of sKitL. SKitL confers signals enhancing mobility of circulating progenitor cells. Along with stromalderived factor-1 (SDF-1 gradient, the circulating progenitor cells are mobilized into the peripheral circulation. CPC indicates circulating progenitor cells; VCAM-1, vascular cell adhesion molecule-1; and VLA4, very late antigen-4.

leading to enhanced mobilization of progenitor cells, as


depicted above. As soon as the EPCs are circulating, the most
important factors for tissue engraftment of the mobilized cells
are the local concentration of stromal-derived factor-1 and
its cell receptor CXCR4.156 This notion is further supported
by the observation that mice lacking CXCR4 die in utero
because of defects in vascular development.157 Although the
animal data make shear stressinduced NO generation a
likely mechanism for NO-mediated EPC mobilization, animal
experiments with ischemic exercise are still lacking. Therefore, the aforementioned controversy still awaits experimental resolution.
Exercise and Plaque Regression in CAD
Although it is high on the agenda in lipid research, plaque
regression ceased to be a research focus of exercise research
in recent years. Two decades ago, a series of angiographic
long-term follow-up studies documented that regular endurance exercise training can retard the progression of coronary
atherosclerosis158,159 or even reduce stenosis diameter if
combined with lipid-management techniques of smoking
cessation and other lifestyle interventions.160,161 Surprisingly,
not a single study has addressed exercise-mediated changes in
plaque volume by intravascular ultrasound thus far to control
the findings of the early regression studies with a more

accurate technique, which also permits insight into changes of


plaque composition.
Arterial Resistance Vessels
Larger and medium-sized resistance vessels (150 m) are
exposed to the predominant influence of endotheliummediated vasodilatation, whereas smaller arterioles are subject to metabolic and myogenic control. The vasomotor state
of the distal microvasculature is influenced by the dominant
effects of local myocardial metabolic demands.
Metabolic Control
In principle, all concepts for metabolic control of resistance
vessel diameter are based on the presence of an oxygen/metabolic sensor coupled to vascular smooth muscle cell tension
to control vascular tone.162 Four models are currently under
discussion, as follows:
1. An oxygen sensor in the vessel wall leads to production
of the vasodilatory prostaglandin G2 in response to
decreases in PO2.163 There is evidence for prostaglandin-mediated vasodilation via KATP channels.164
2. Adenosine serves as a myocardial oxygen sensor because it is generated at an accelerated rate when oxygen
supply-to-demand relationship falls.165
3. Carbon dioxide production, which in turn leads to tissue
acidosis, is known to increase coronary flow.166 Acido-

Gielen et al

Cardiovascular Effects of Exercise Training

1231

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Figure 6. In the microcirculation, several pathways for metabolic vasodilation have been described. PO2-dependent regulation represents hypoxic coronary flow control; metabolic rate of oxygen (MRO2) dependent regulation represents normoxic coronary flow control. Under hypoxemic conditions, Ca2i is increased in endothelial cells, leading to phospholipase activation with consecutive generation of prostaglandins and cAMP-mediated metabolic vasodilation. Alternatively, ATP release either from erythrocytes or from the
working skeletal muscle activates the adenosine purinergic A2A receptor, resulting in activation of the KATP channel with potassium loss
and hyperpolarization of the vascular smooth muscle cell. As a consequence, voltage-gated Ca2 channels (VGCC) are opened, and
Ca2i is decreased with vascular relexation and vasodilation. Acidosis can directly influence the opening probability of the KATP channel. Exercise increases the adenosine sensitivity, resulting in larger metabolic vasorelaxation. Hb indicates hemoglobin; AC, adenylate
cyclase; and PGI2, prostaglandin I2. Adapted from Deussen et al 162 with kind permission of Springer Science and Business Media.
Copyright 2006, Springer Science and Business Media

sis in turn increases the sensitivity of adenosine A2A


receptors.
4. Erythrocyte ATP release may cause vasodilatation, as
documented in studies confirming the necessity of
erythrocytes to mediate vasodilation under hypoxic
conditions.167
All vasodilatory stimuli converge on the activation of the
KATP channels, the opening of which leads to vascular
smooth muscle cell hyperpolarization and reduced activation
of voltage-gated calcium channels (Figure 6).
Exercise training increases resistance vessel sensitivity and
maximal responsiveness to adenosine. This has been confirmed in dogs and miniature swine in vivo.168,169 In humans,
adenosine-induced microvascular function improved by 29%
after exercise training, as assessed by coronary flow reserve
ratio in response to intracoronary adenosine infusion. With
regard to the different mechanisms described above, no data
are currently available to describe exercise-mediated specific
changes.

Myogenic Control
It is well established in porcine animal models of treadmill
training that myogenic vasoconstriction to increases in perfusion pressure (particularly 40 mm Hg) is augmented after
exercise.170 Evidence is accumulating that L-type voltagegated calcium channels and protein kinase C determine
myogenic tone under physiological conditions.171 Korzick et
al172 demonstrated that increased myogenic tone in trained
pigs involves changes in protein kinase C- expression and
phosphorylation, which augment the depolarization-related
opening of voltage-gated calcium channels in smooth muscle
cells and increase of intracellular calcium.
Angiogenesis and Microcirculation
In the cardiac muscle, White et al173 conclusively showed in
a porcine model of exercise training that training increases
the total vascular bed cross-sectional area by up to 37% after
16 weeks. In humans, changes in cardiac vascularization are
difficult to assess.
More data, however, are available for the skeletal muscle.
In addition to an effect on metabolic alterations and the

1232

Circulation

September 21, 2010

Figure 7. A model depicting exerciseinduced angiogenesis in the peripheral


skeletal muscle. A central player for
enhanced capillary density, VEGF can be
activated either by PGC-1 or by hypoxia-inducible transcription factor 1 (HIF1) because of local hypoxia or a reduction in energy equivalents. TF indicates
transcription factor.
Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

catabolic/anabolic balance within the skeletal muscle, exercise also affects the vascularization of the skeletal muscle. It
has been known for a long time that endurance exercise
training induces a significant increase in muscle capillaries in
either animals or humans.174,175 This process is largely
coordinated by soluble factors emanating from tissues in need
of vascularization. One of the best factors to study in this
scenario is VEGF. It is a powerful activator of endothelial
proliferation and is crucial for nearly all forms of neovascularization.176 With respect to exercise training, several studies
showed that VEGF transcription in skeletal muscle is increased by short-term177 as well as endurance training.142 It is
thought that local skeletal muscle hypoxia stabilizes the
transcription factor hypoxia inducible factor-1, leading to
the induction of VEGF. However, this concept must be
questioned because the deletion of hypoxia inducible
factor-1 in the skeletal muscle increases rather than decreases capillary density.178
In recent years, the transcriptional coactivator peroxisome
proliferator-activated receptor- coactivator (PGC)-1 has
been regarded as an important exercise-mediated regulator of
VEGF transcription. Mice lacking PGC-1 in their skeletal
muscle failed to increase capillary density in response to
exercise training.179 The central role of PGC-1 is furthermore supported by the observation that transgenic expression
of PGC-1 in skeletal muscle dramatically increases capillary
density.180 The mechanism by which PGC-1 induces VEGF
transcription appears to involve the coactivation of the orphan
nuclear receptor ERR.180 Which signals are upstream of
PGC-1 mediating the exercise-induced angiogenesis? One
pathway is the exercise-dependent modulation of high-energy
phosphates and the concomitant activation of AMP-activated

protein kinase.181 However, if this pathway is only responsible for basal VEGF expression and capillarization and not for
exercise-induced angiogenesis is still an ongoing discussion.182 In addition to activation of AMPK, the activation of
the p38MAPK (Mitogen-activated protein kinase) pathway
by exercise training also leads to the activation of PGC-1.183
Last but not least, the increased production of ROS by
exercise training can also contribute to the induction of
PGC-1.184 Therefore, one may summarize that endurance
exercise by activating PGC-1 via AMPK or p38MAPK or
ROS leads to an enhanced transcription of VEGF, finally
resulting in angiogenesis (Figure 7).
Venous Circulation
Compared with the arterial bed, the venous circulation has
received less attention in vascular research, particularly with
regard to training effects. The reasons are diverse, ranging
from methodological problems to underestimation of the
physiological relevance of venous function.185
Key parameters of venous function such as venous capacitance, which reflects the blood volume stored in the venous
system, and venous compliance, which represents the change
in volume in response to a change in pressure, change with
aging, posture, and physical activity. Aging alters the composition of the venous wall, resulting in lower venous
compliance in sedentary senior individuals.186 In both young
and elderly endurance-trained men, however, venous compliance was 70% to 120% higher compared with their sedentary
age-matched counterparts. It is currently unclear whether
these changes are primarily the result of factors such as
increased venous NO availability and decreased venous
smooth muscle tone, sympathetic -adrenergic vasomotor
control, or structural effects on elastin-to-collagen ratio.

Gielen et al
Immobilization, as tested in bed rest studies, demonstrated
a significant decline in venous capacitance after 52 days,
which was not affected by resistive vibration exercise.187
Venous compliance remained unaffected.
In patients with heart failure, exercise forearm venous
capacitance and venous outflow, as measured by strain-gauge
plethysmography, are significantly reduced and seem to be
related to maximal walking distance.188 Although the methodology of strain-gauge plethysmography was used in a
number of studies to examine training effects on vascular
function,189,190 changes of venous capacitance induced by
training have not been reported thus far. From animal
experiments, we know that the capillary domain area values
for venular capillaries were significantly decreased after 4
weeks of training.191

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Pulmonary Circulation
Pulmonary hypertension is a frequent accompanying problem
in heart failure or valvular heart disease. Similar to heart
failure, exercise has long been regarded as detrimental and
was discouraged in pulmonary hypertension because of the
fear of fatal cardiovascular compromise.
Human Data
The first reports on training effects on pulmonary pressure
came from aerobic training studies in patients with stable
CHF. Lee et al192 were the first to demonstrate the lack of any
exercise-induced rise in resting pulmonary artery pressure in
patients with CHF, a finding confirmed by Dubach et al193 18
years later. In a large prospective randomized study involving
73 patients with CHF, Hambrecht et al65 were finally able to
demonstrate that pulmonary vascular resistance at rest and at
peak exercise was decreased after 6 months of endurance
training.
Just recently, the first clinical study in patients with stable
precapillary pulmonary hypertension confirmed that 15
weeks of endurance exercise are effective in significantly
improving 6-minute walking distance on top of optimal
medical therapy.194
Animal Experiments
The first rat model of pulmonary hypertension to be used with
exercise training was hypobaric hypoxia. Rats were trained
for 4 and 6 weeks under hypobaric hypoxia (equivalent to
altitudes of 2500 and 5500 m). Kashimura et al196 found that
the increase in resting mean pulmonary arterial pressure with
increasing equivalent altitude was lower in the 2 trained
groups than in the control group and greater after 6 than after
4 weeks of exercise.195 In a follow-up study, they showed that
6 weeks of training can attenuate the hypoxia-induced and
angiotensin IIinduced PH. These findings were extended by
Favret et al,197 who demonstrated that exercise training
improved lung gas exchange and attenuated acute hypoxic
pulmonary hypertension but did not prevent pulmonary hypertension of prolonged hypoxia.
In rabbits, it was soon documented that treadmill exercise
improved acetylcholine-induced endothelium-dependent vasodilatation not only in aortic rings but also in precontracted
pulmonary artery rings,198 a finding that could not be reproduced in Sprague-Dawley rats, however.199 A possible expla-

Cardiovascular Effects of Exercise Training

1233

nation for the discrepant results can be derived from 2 studies


by Johnson et al.200,201 Whereas long-term training enhanced
endothelium-dependent vasorelaxation in pulmonary arteries
by mechanisms of increased reliance on NO and reduced
production of a prostanoid constrictor in miniature swine with
surgically induced chronic coronary artery occlusion, it did
not alter pulmonary vasorelaxation in normal pigs.200 However, short-term 1-week training was effective in improving
bradykinin-induced vasorelaxation and in increasing pulmonary artery eNOS protein expression. SOD remained unchanged.201 In essence, training-induced changes in pulmonary vasomotion are modified by preexistent endothelial
dysfunction and by training duration.
In a rat model, Handoko et al202 induced compensated and
progressive pulmonary hypertension using different doses of
monocrotaline. Only the compensated pulmonary hypertension animals benefited from exercise training with increased
exercise tolerance and right ventricular capillarization,
whereas animals with progressive right ventricular failure had
reduced cardiac output and died prematurely from right
ventricular failure.

Conclusion
To summarize, cardiac effects of exercise include an improved protection against I/R injury primarily as a result of
higher antioxidative protection and improved LV diastolic
and systolic function in chronic heart failure due to favorable
changes in neurohormonal state, activation of PI3K (p110)
attenuating pathological hypertrophy, normalization of cardiomyocyte calcium handling, and inhibition of the catabolic
ubiquitin-proteasome system.
Vascular effects are based on improved endotheliummediated flow-induced vasodilatation in conduit arteries and
larger resistance arteries, higher myogenic control, and increased metabolic vasodilatation in small resistance arteries.
Vascular regeneration by mobilization of endothelial progenitor cells is augmented by exercise. Recently, improved
endothelial vasodilatation has also been confirmed in the
pulmonary artery.
Extending our concept of molecular mechanisms of exercise is essential to further optimize training interventions with
regard to their clinical efficacy and to identify novel targets
for pharmaceutical intervention. However, certain areas (ie,
microcirculation and the venous system) are still incompletely understood and merit further molecular studies. Residual incomplete molecular understanding should not prevent the clinical use of training interventions with established
clinical benefit.

Sources of Funding
Dr Gielen is supported by the Deutsche Forschungsgemeinschaft,
grant GI535/11.

Disclosures
None.

References
1. Booth F, Laye M, Lees S, Rector R, Thyfault J. Reduced physical
activity and risk of chronic disease: the biology behind the consequences. Eur J Appl Physiol. 2008;102:381390.

1234

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

2. Mokdad AH, Giles WH, Bowman BA, Mensah GA, Ford ES, Smith
SM, Marks JS. Changes in health behaviors among older Americans,
1990 to 2000. Public Health Rep. 2004;119:356 361.
3. Perhonen MA, Franco F, Lane LD, Buckey JC, Blomqvist CG, Zerwekh
JE, Peshock RM, Weatherall PT, Levine BD. Cardiac atrophy after bed
rest and spaceflight. J Appl Physiol. 2001;91:645 653.
4. Dorfman TA, Levine BD, Tillery T, Peshock RM, Hastings JL,
Schneider SM, Macias BR, Biolo G, Hargens AR. Cardiac atrophy in
women following bed rest. J Appl Physiol. 2007;103:8 16.
5. Sipola P, Heikkinen J, Laaksonen DE, Kettunen R. Influence of 12
weeks of jogging on magnetic resonance-determined left ventricular
characteristics in previously sedentary subjects free of cardiovascular
disease. Am J Cardiol. 2009;103:567571.
6. Navarro-Arevalo A, Canavate C, Sanchez-del-Pino M. Myocardial and
skeletal muscle aging and changes in oxidative stress in relationship to
rigorous exercise training. Mech Ageing Dev. 1999;108:207217.
7. Rinaldi B, Corbi G, Boccuti S, Filippelli W, Rengo G, Leosco D, Rossi
F, Filippelli A, Ferrara N. Exercise training affects age-induced changes
in SOD and heat shock protein expression in rat heart. Exp Gerontol.
2006;41:764 770.
8. Devi SA, Prathima S, Subramanyam MVV. Dietary vitamin E and
physical exercise, II: antioxidant status and lipofuscin-like substances in
aging rat heart. Exp Gerontol. 2003;38:291297.
9. Martin JL, Mestril R, Hilal-Dandan R, Brunton LL, Dillmann WH.
Small heat shock proteins and protection against ischemic injury in
cardiac myocytes. Circulation. 1997;96:4343 4348.
10. Starnes JW, Choilawala AM, Taylor RP, Nelson MJ, Delp MD. Myocardial heat shock protein 70 expression in young and old rats after
identical exercise programs. J Gerontol A Biol Sci Med Sci. 2005;60:
963969.
11. Olivetti G, Melissari M, Capasso JM, Anversa P. Cardiomyopathy of the
aging human heart: myocyte loss and reactive cellular hypertrophy. Circ
Res. 1991;68:1560 1568.
12. Phaneuf S, Leeuwenburgh C. Cytochrome c release from mitochondria
in the aging heart: a possible mechanism for apoptosis with age. Am J
Physiol. 2002;282:R423R430.
13. Sussman MA, Anversa P. Myocardial aging and senescence: where have
the stem cells gone? Annu Rev Physiol. 2004;66:29 48.
14. Kwak HB, Song W, Lawler JM. Exercise training attenuates ageinduced elevation in Bax/Bcl-2 ratio, apoptosis, and remodeling in the
rat heart. FASEB J. 2006;20:791793.
15. Pugh KG, Wei JY. Clinical implications of physiological changes in the
aging heart. Drugs Aging. 2001;18:263276.
16. Arbab-Zadeh A, Dijk E, Prasad A, Fu Q, Torres P, Zhang R, Thomas JD,
Palmer D, Levine BD. Effect of aging and physical activity on left
ventricular compliance. Circulation. 2004;110:1799 1805.
17. Pluim BM, Zwinderman AH, van der Laarse A, van der Wall EE. The
athletes heart: a meta-analysis of cardiac structure and function.
Circulation. 2000;101:336 344.
18. Pelliccia A, Culasso F, Di Paolo FM, Maron BJ. Physiologic left ventricular cavity dilatation in elite athletes. Ann Intern Med. 1999;130:
2331.
19. Fagard R. Athletes heart. Heart. 2003;89:14551461.
20. Maron BJ. Distinguishing hypertrophic cardiomyopathy from athletes
heart physiological remodelling: clinical significance, diagnostic
strategies and implications for preparticipation screening. Br J Sports
Med. 2009;43:649 656.
21. Neri Serneri GG, Boddi M, Modesti PA, Cecioni I, Coppo M, Padeletti
L, Michelucci A, Colella A, Galanti G. Increased cardiac sympathetic
activity and insulin-like growth factor-I formation are associated with
physiological hypertrophy in athletes. Circ Res. 2001;89:977982.
22. McMullen JR, Amirahmadi F, Woodcock EA, Schinke-Braun M,
Bouwman RD, Hewitt KA, Mollica JP, Zhang L, Zhang Y, Shioi T,
Buerger A, Izumo S, Jay PY, Jennings GL. Protective effects of exercise
and phosphoinositide 3-kinase(p110) signaling in dilated and hypertrophic cardiomyopathy. Proc Natl Acad Sci U S A. 2007;104:612 617.
23. Luo J, McMullen JR, Sobkiw CL, Zhang L, Dorfman AL, Sherwood
MC, Logsdon MN, Horner JW, DePinho RA, Izumo S, Cantley LC.
Class IA phosphoinositide 3-kinase regulates heart size and physiological cardiac hypertrophy. Mol Cell Biol. 2005;25:94919502.
24. Care` A, Catalucci D, Felicetti F, Bonci D, Addario A, Gallo P, Bang
ML, Segnalini P, Gu Y, Dalton ND, Elia L, Latronico MVG, Hoydal M,
Autore C, Russo MA, Dorn GW, Ellingsen O, Ruiz-Lozano P, Peterson
KL, Croce CM, Peschle C, Condorelli G. MicroRNA-133 controls
cardiac hypertrophy. Nat Med. 2007;13:613 618.

25. Elia L, Contu R, Quintavalle M, Varrone F, Chimenti C, Russo MA,


Cimino V, De Marinis L, Frustaci A, Catalucci D, Condorelli G.
Reciprocal regulation of microRNA-1 and insulin-like growth factor-1
signal transduction cascade in cardiac and skeletal muscle in physiological and pathological conditions. Circulation. 2009;120:23772385.
26. Catalucci D, Latronico MVG, Ceci M, Rusconi F, Young HS, Gallo P,
Santonastasi M, Bellacosa A, Brown JH, Condorelli G. Akt increases
sarcoplasmic reticulum Ca2 cycling by direct phosphorylation of
phospholamban at Thr17. J Biol Chem. 2009;284:28180 28187.
27. Kemi OJ, Haram PM, Wisloff U, Ellingsen O. Aerobic fitness is associated with cardiomyocyte contractile capacity and endothelial function
in exercise training and detraining. Circulation. 2004;109:28972904.
28. Kemi OJ, Haram PM, Loennechen JP, Osnes JB, Skomedal T, Wisloff
U, Ellingsen O. Moderate vs. high exercise intensity: differential effects
on aerobic fitness, cardiomyocyte contractility, and endothelial function.
Cardiovasc Res. 2005;67:161172.
29. Whyte GP, George K, Sharma S, Lumley S, Gates P, Prasad K,
McKenna WJ. Cardiac fatigue following prolonged endurance exercise
of differing distances. Med Sci Sports Exerc. 2000;32:10671072.
30. Scott J, Warburton DER. Mechanisms underpinning exercise-induced
changes in left ventricular function. Med Sci Sports Exerc. 2008;40:
1400 1407.
31. Dawson EA, Shave R, Whyte G, Ball D, Selmer C, Jans O, Secher NH,
George KP. Preload maintenance and the left ventricular response to
prolonged exercise in men. Exp Physiol. 2007;92:383390.
32. Welsh RC, Warburton DER, Humen DP, Taylor DA, McGavock J,
Haykowsky MJ. Prolonged strenuous exercise alters the cardiovascular
response to dobutamine stimulation in male athletes. J Physiol. 2005;
569:325330.
33. Vatner SF, Hittinger L. Coronary vascular mechanisms involved in
decompensation from hypertrophy to heart failure. J Am Coll Cardiol.
1993;22:34A 40A.
34. Marrero MB, Horwath SM, Wilkerson JE. Acute blood biochemical
alterations in response to marathon running. Eur J Appl Physiol Occup
Physiol. 1975;34:173181.
35. Friedman DB, Ordway GA, Williams RS. Exercise-induced functional
desensitization of canine cardiac beta-adrenergic receptors. J Appl
Physiol. 1987;62:17211723.
36. Reeve JL, Duffy AM, OBrien T, Samali A. Dont lose heart: therapeutic value of apoptosis prevention in the treatment of cardiovascular
disease. J Cell Mol Med. 2005;9:609 622.
37. Powers SK, Quindry JC, Kavazis AN. Exercise-induced cardioprotection against myocardial ischemia-reperfusion injury. Free Radic Biol
Med. 2008;44:193201.
38. Laughlin MH, Oltman CL, Bowles DK. Exercise training-induced adaptations in the coronary circulation. Med Sci Sports Exerc. 1998;30:
352360.
39. Quindry JC, Hamilton KL, French JP, Lee Y, Murlasits Z, Tumer N,
Powers SK. Exercise-induced HSP-72 elevation and cardioprotection
against infarct and apoptosis. J Appl Physiol. 2007;103:1056 1062.
40. Hamilton KL, Powers SK, Sugiura T, Kim S, Lennon S, Tumer N,
Mehta JL. Short-term exercise training can improve myocardial tolerance to I/R without elevation in heat shock proteins. Am J Physiol.
2001;281:H1346 H1352.
41. Quindry J, French J, Hamilton K, Lee Y, Selsby J, Powers S. Exercise
does not increase cyclooxygenase-2 myocardial levels in young or
senescent hearts. J Physiol Sci. 2010;60:181186.
42. Murlatsis Z, Lee Y, Powers SK. Short-term exercise does not increase
ER stress protein expression in cardiac muscle. Med Sci Sports Exerc.
2007;39:15221528.
43. Zweier JL, Talukder MAH. The role of oxidants and free radicals in
reperfusion injury. Cardiovasc Res. 2006;70:181190.
44. Li C, Jackson RM. Reactive species mechanisms of cellular hypoxiareoxygenation injury. Am J Physiol. 2002;282:C227C241.
45. Adlam VJ, Harrison JC, Porteous CM, James AM, Smith RAJ, Murphy
MP, Sammut IA. Targeting an antioxidant to mitochondria decreases
cardiac ischemia-reperfusion injury. FASEB J. 2005;19:1088 1095.
46. Powers SK, Demirel HA, Vincent HK, Coombes JS, Naito H, Hamilton
KL, Shanely RA, Jessup J. Exercise training improves myocardial tolerance to in vivo ischemia-reperfusion in the rat. Am J Physiol. 1998;
275:R1468 R1477.
47. Ascensao A, Ferreira R, Magalhnes J. Exercise-induced cardioprotection: biochemical, morphological and functional evidence in whole
tissue and isolated mitochondria. Int J Cardiol. 2007;117:16 30.

Gielen et al

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

48. Hamilton KL, Quindry JC, French JP, Staib J, Hughes J, Mehta JL,
Powers SK. MnSOD antisense treatment and exercise-induced protection against arrhythmias. Free Radic Biol Med. 2004;37:1360 1368.
49. Powers SK, Criswell D, Lawler J, Martin D, Lieu FK, Ji LL, Herb RA.
Rigorous exercise training increases superoxide dismutase activity in
ventricular myocardium. Am J Physiol. 1993;34:H2094 H2098.
50. Kanter MM, Hamlin RL, Unverferth DV, Davis HW, Merola AJ. Effect
of exercise training on antioxidant enzymes and cardiotoxicity of doxorubicin. J Appl Physiol. 1985;59:1298 1303.
51. Lennon SL, Quindry JC, Hamilton KL, French JP, Hughes J, Mehta JL,
Powers SK. Elevated MnSOD is not required for exercise-induced
cardioprotection against myocardial stunning. Am J Physiol. 2004;287:
H975H980.
52. Marcil M, Bourduas K, Ascah A, Burelle Y. Exercise training induces
respiratory substrate-specific decrease in Ca2-induced permeability
transition pore opening in heart mitochondria. Am J Physiol. 2006;290:
H1549 H1557.
53. Starnes JW, Barnes BD, Olsen ME. Exercise training decreases rat heart
mitochondria free radical generation but does not prevent Ca2-induced
dysfunction. J Appl Physiol. 2007;102:17931798.
54. Kavazis AN. Exercise preconditioning of the myocardium. Sports Med.
2009;39:923935.
55. Pchejetski D, Kunduzova O, Dayon A, Calise D, Seguelas MH, Leducq
N, Seif I, Parini A, Cuvillier O. Oxidative stress dependent sphingosine
kinase-1 inhibition mediates monoamine oxidase Aassociated cardiac
cell apoptosis. Circ Res. 2007;100:41 49.
56. Gross GJ, Peart JN. KATP channels and myocardial preconditioning: an
update. Am J Physiol. 2003;285:H921H930.
57. Goll DE, Thompson VF, Li H, Wei W, Cong J. The calpain system.
Physiol Rev. 2003;83:731 801.
58. French JP, Quindry JC, Falk DJ, Staib JL, Lee Y, Wang KKW, Powers
SK. Ischemia-reperfusion-induced calpain activation and SERCA2a
degradation are attenuated by exercise training and calpain inhibition.
Am J Physiol. 2006;290:H128 H136.
59. Starnes JW, Taylor RP, Ciccolo JT. Habitual low-intensity exercise does
not protect against myocardial dysfunction after ischemia in rats. Eur
J Cardiovasc Prev Rehabil. 2005;12:169 174.
60. Lennon SL, Quindry JC, French JP, Kim S, Mehta JL, Powers SK.
Exercise and myocardial tolerance to ischaemia-reperfusion. Acta
Physiol Scand. 2004;182:161169.
61. Demirel HA, Powers SK, Zergeroglu MA, Shanely RA, Hamilton K,
Coombes J, Naito H. Short-term exercise improves myocardial tolerance
to in vivo ischemia-reperfusion in the rat. J Appl Physiol. 2001;91:
22052212.
62. Lennon SL, Quindry J, Hamilton KL, French J, Staib J, Mehta JL,
Powers SK. Loss of exercise-induced cardioprotection after cessation of
exercise. J Appl Physiol. 2004;96:1299 1305.
63. Volterrani M, Clark AL, Ludman PF, Swan JW, Adamopoulos S,
Piepoli M, Coats AJS. Predictors of exercise capacity in chronic heart
failure. Eur Heart J. 1994;15:801 809.
64. Sullivan MJ, Higginbotham MB, Cobb FR. Exercise training in patients
with severe left ventricular dysfunction: hemodynamic and metabolic
effects. Circulation. 1988;78:506 515.
65. Hambrecht R, Gielen S, Linke A, Fiehn E, Yu J, Walther C, Schoene N,
Schuler G. Effects of exercise training on left ventricular function and
peripheral resistance in patients with chronic heart failure: a randomised
trial. JAMA. 2000;283:30953101.
66. Giannuzzi P, Temporelli PL, Corra U, Tavazzi L. Antiremodeling effect
of long-term exercise training in patients with stable chronic heart
failure: results of the Exercise in Left Ventricular Dysfunction and
Chronic Heart Failure (ELVD-CHF) Trial. Circulation. 2003;108:
554 559.
67. Hambrecht R, Fiehn E, Weigl C, Gielen S, Hamann C, Kaiser R, Yu J,
Adams V, Niebauer J, Schuler G. Regular physical exercise corrects
endothelial dysfunction and improves exercise capacity in patients with
chronic heart failure. Circulation. 1998;98:2709 2715.
68. Adams V, Link A, Gielen S, Erbs S, Hambrecht R, Schuler G. Modulation of Murf-1 and MAFbx expression in the myocardium by physical
exercise training. Eur J Cardiovasc Prev Rehabil. 2008;15:293299.
69. Lenk K, Schur R, Linke A, Erbs S, Matsumoto Y, Adams V, Schuler G.
Impact of exercise training on myostatin expression in the myocardium
and skeletal muscle in a chronic heart failure model. Eur J Heart Fail.
2009;11:342348.
70. Rolim NPL, Medeiros A, Rosa KT, Mattos KC, Irigoyen MC, Krieger
EM, Krieger JE, Negrao CE, Brum PC. Exercise training improves the

Cardiovascular Effects of Exercise Training

71.

72.

73.

74.

75.

76.

77.

78.

79.

80.

81.

82.

83.

84.

85.

86.
87.

88.

1235

net balance of cardiac Ca2 handling protein expression in heart failure.


Physiol Genomics. 2007;29:246 252.
Lu L, Mei DF, Gu AG, Wang S, Lentzner B, Gutstein DE, Zwas D,
Homma S, Yi GH, Wang J. Exercise training normalizes altered
calcium-handling proteins during development of heart failure. J Appl
Physiol. 2002;92:1524 1530.
Kemi OJ, Ellingsen O, Ceci M, Grimaldi S, Smith GL, Condorelli G,
Wisloff U. Aerobic interval training enhances cardiomyocyte contractility and Ca2 cycling by phosphorylation of CaMKII and Thr-17 of
phospholamban. J Mol Cell Cardiol. 2007;43:354 361.
Wisloff U, Loennechen JP, Falck G, Beisvag V, Currie S, Smith G,
Ellingsen O. Increased contractility and calcium sensitivity in cardiac
myocytes isolated from endurance trained rats. Cardiovasc Res. 2001;
50:495508.
Kemi OJ, Wisloff U. Mechanisms of exercise-induced improvements in
the contractile apparatus of the mammalian myocardium. Acta Physiol.
2010;199:425 439.
Coats AJS, Adamopoulos S, Radaelli A, McCance A, Meyer TE,
Bernardi L, Solda PL, Davey P, Ormerod O, Forfar C, Conway J,
Sleight P. Controlled trial of physical training in chronic heart
failure: exercise performance, hemodynamics, ventilation, and
autonomic function. Circulation. 1992;85:2119 2131.
Braith R, Welsch M, Feigenbaum M, Kluess HA, Pepine C. Neuroendocrine activation in heart failure is modified by endurance training.
J Am Coll Cardiol. 1999;34:1170 1175.
Xu X, Wan W, Powers AS, Li J, Ji LL, Lao S, Wilson B, Erikson JM,
Zhang JQ. Effects of exercise training on cardiac function and myocardial remodeling in post myocardial infarction rats. J Mol Cell
Cardiol. 2008;44:114 122.
Paulus WJ, Tschope C, Sanderson JE, Rusconi C, Flachskampf FA,
Rademakers FE, Marino P, Smiseth OA, De Keulenaer G, Leite-Moreira
AF, Borbely A, Edes I, Handoko ML, Heymans S, Pezzali N, Pieske B,
Dickstein K, Fraser AG, Brutsaert DL. How to diagnose diastolic heart
failure: a consensus statement on the diagnosis of heart failure with
normal left ventricular ejection fraction by the Heart Failure and Echocardiography Associations of the European Society of Cardiology. Eur
Heart J. 2007;28:2539 2550.
Brandao MU, Wajngarten M, Rondon E, Giorgi MC, Hironaka F,
Negrao CE. Left ventricular function during dynamic exercise in
untrained and moderately trained subjects. J Appl Physiol. 1993;75:
1989 1995.
Levy WC, Cerqueira MD, Abrass IB, Schwartz RS, Stratton JR.
Endurance exercise training augments diastolic filling at rest and during
exercise in healthy young and older men. Circulation. 1993;88:
116 126.
Yu CM, Li LS-W, Lam MF, Siu DC-W, Miu RK-M, Lau CP. Effect of
a cardiac rehabilitation program on left ventricular diastolic function and
its relationship to exercise capacity in patients with coronary heart
disease: experience from a randomized, controlled study. Am Heart J.
2004;147:874.
Smart N, Haluska B, Jeffriess L, Marwick TH. Exercise training in
systolic and diastolic dysfunction: effects on cardiac function, functional
capacity, and quality of life. Am Heart J. 2007;153:530 536.
Brassard P, Legault S, Garneau C, Bogaty P, Dumesnil JG, Poirier P.
Normalization of diastolic dysfunction in type 2 diabetics after exercise
training. Med Sci Sports Exerc. 2007;39:1896 1901.
Gary RA, Sueta CA, Dougherty M, Rosenberg B, Cheek D, Preisser J,
Neelon V, McMurray R. Home-based exercise improves functional
performance and quality of life in women with diastolic heart failure.
Heart Lung. 2004;33:210 218.
Pieske B, Maier LS, Bers DM, Hasenfuss G. Ca2 handling and
sarcoplasmic reticulum Ca2 content in isolated failing and nonfailing
human myocardium. Circ Res. 1999;85:38 46.
Marks AR. Ryanodine receptors/calcium release channels in heart
failure and sudden cardiac death. J Mol Cell Cardiol. 2001;33:615 624.
Marx SO, Reiken S, Hisamatsu Y, Jayaraman T, Burkhoff D, Rosemblit
N, Marks AR. PKA phosphorylation dissociates FKBP12.6 from the
calcium release channel (ryanodine receptor): defective regulation in
failing hearts. Cell. 2000;101:365376.
Medeiros A, Rolim NPL, Oliveira RSF, Rosa KT, Mattos KC, Casarini
DE, Irigoyen MC, Krieger EM, Krieger JE, Negrao CE, Brum PC.
Exercise training delays cardiac dysfunction and prevents calcium
handling abnormalities in sympathetic hyperactivity-induced heart
failure mice. J Appl Physiol. 2008;104:103109.

1236

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

89. Choi SY, Chang HJ, Choi SI, Kim KI, Cho YS, Youn TJ, Chung WY,
Chae IH, Choi DJ, Kim HS, Kim CH, Oh BH, Kim MH. Long-term
exercise training attenuates age-related diastolic dysfunction: association of myocardial collagen cross-linking. J Korean Med Sci. 2009;
24:3239.
90. Goldbarg SH, Elmariah S, Miller MA, Fuster V. Insight into degenerative aortic valve disease. J Am Coll Cardiol. 2007;50:12051213.
91. Agmon Y, Khandheria BK, Meissner I, Sicks JD, OFallon WM,
Wiebers DO, Whisnant JP, Seward JB, Tajik AJ. Aortic valve sclerosis
and aortic atherosclerosis: different manifestations of the same disease?
Insights from a population-based study. J Am Coll Cardiol. 2001;38:
827 834.
92. Aikawa E, Aikawa M, Libby P, Figueiredo JL, Rusanescu G, Iwamoto
Y, Fukuda D, Kohler RH, Shi GP, Jaffer FA, Weissleder R. Arterial and
aortic valve calcification abolished by elastolytic cathepsin S deficiency
in chronic renal disease. Circulation. 2009;119:17851794.
93. Laufs U, Wassmann S, Czech T, Munzel T, Eisenhauer M, Bohm M,
Nickenig G. Physical inactivity increases oxidative stress, endothelial
dysfunction, and atherosclerosis. Arterioscler Thromb Vasc Biol. 2005;
25:809 814.
94. Adams V, Linke A, Krankel N, Erbs S, Gielen S, Mobius-Winkler S,
Gummert JF, Mohr FW, Schuler G, Hambrecht R. Impact of regular
physical activity on the NAD(P)H oxidase and angiotensin receptor
system in patients with coronary artery disease. Circulation. 2005;111:
555562.
95. Matsumoto Y, Adams V, Jacob S, Mangner N, Schuler G, Linke A.
Regular exercise training prevents aortic valve disease in LDLR
deficient mice. Circulation. 2010;121:759 767.
96. Mohler ER III. Mechanisms of aortic valve calcification. Am J Cardiol.
2004;94:1396 1402.
97. Helske S, Syvaranta S, Lindstedt KA, Lappalainen J, Oorni K, Mayranpaa MI, Lommi J, Turto H, Werkkala K, Kupari M, Kovanen PT.
Increased expression of elastolytic cathepsins S, K, and V and their
inhibitor cystatin C in stenotic aortic valves. Arterioscler Thromb Vasc
Biol. 2006;26:17911798.
98. Miller JD, Chu Y, Brooks RM, Richenbacher WE, Peta-Silva R, Heistad
DD. Dysregulation of antioxidant mechanisms contributes to increased
oxidative stress in calcific aortic valvular stenosis in humans. J Am Coll
Cardiol. 2008;52:843 850.
99. Weiss RM, Ohashi M, Miller JD, Young SG, Heistad DD. Calcific aortic
valve stenosis in old hypercholesterolemic mice. Circulation. 2006;114:
20652069.
100. Hruska KA, Mathew S, Saab G. Bone morphogenetic proteins in
vascular calcification. Circ Res. 2005;97:105114.
101. Altschul R. Endothelium, its Development, Morphology, Function, and
Pathology. New York, NY: Macmillan; 1954.
102. Ross J. The pathogenesis of atherosclerosis: a perspective for the 1990s.
Nature. 1993;362:801 809.
103. Schachinger V, Britten MB, Zeiher A. Prognostic impact of coronary
vasodilator dysfunction on adverse long-term outcome of coronary heart
disease. Circulation. 2000;101:1899 906.
104. Perticone F, Ceravolo R, Pujia A, Ventura G, Iacopino S, Scozzafava A,
Ferraro A, Chello M, Mastroroberto P, Verdecchia P, Schillaci G.
Prognostic significance of endothelial dysfunction in hypertensive
patients. Circulation. 2001;104:191196.
105. Landmesser U, Drexler H. The clinical significance of endothelial dysfunction. Curr Opin Cardiol. 2005;20:547551.
106. Ghiadoni L, Versari D, Giannarelli C, Faita F, Taddei S. Non-invasive
diagnostic tools for investigating endothelial dysfunction. Curr Pharm
Des. 2008;14:37153722.
107. Jones CJ, Kuo L, Davis MJ, Chilian WM. Regulation of coronary blood
flow: coordination of heterogeneous control mechanisms in vascular
microdomains. Cardiovasc Res. 1995;29:585596.
108. Kuo L, Davis MJ, Chilian WM. Longitudinal gradients for endothelium-dependent and -independent vascular responses in the coronary
microcirculation. Circulation. 1995;92:518 525.
109. Wagenseil JE, Mecham RP. Vascular extracellular matrix and arterial
mechanics. Physiol Rev. 2009;89:957989.
110. Dart AM, Lacombe F, Yeoh JK, Cameron JD, Jennings GL, Laufer E,
Esmore DS. Aortic distensibility in patients with isolated hypercholesterolaemia, coronary artery disease, or cardiac transplant. Lancet. 1991;
338:270 273.
111. Avolio A, Jones D, Tafazzoli-Shadpour M. Quantification of alterations
in structure and function of elastin in the arterial media. Hypertension.
1998;32:170 175.

112. Petersen SE, Wiesmann F, Hudsmith LE, Robson MD, Francis JM,
Selvanayagam JB, Neubauer S, Channon KM. Functional and structural
vascular remodeling in elite rowers assessed by cardiovascular magnetic
resonance. J Am Coll Cardiol. 2006;48:790 797.
113. Otsuki T, Maeda S, Iemitsu M, Saito Y, Tanimura Y, Ajisaka R,
Miyauchi T. Vascular endothelium-derived factors and arterial stiffness
in strength- and endurance-trained men. Am J Physiol. 2007;292:
H786 H791.
114. Collier SR, Kanaley JA, Carhart R Jr, Frechette V, Tobin MM, Hall AK,
Luckenbaugh AN, Fernhall B. Effect of 4 weeks of aerobic or resistance
exercise training on arterial stiffness, blood flow and blood pressure in
pre- and stage-1 hypertensives. J Hum Hypertens. 2008;22:678 686.
115. Edwards DG, Schofield RS, Magyari PM, Nichols WW, Braith RW.
Effect of exercise training on central aortic pressure wave reflection in
coronary artery disease. Am J Hypertens. 2004;17:540 553.
116. Ludmer PL, Selwyn AP, Shook TL, Wayne RR, Mudge GH, Alexander
RW, Ganz P. Paradoxical vasoconstriction induced by acetylcholine in
atherosclerotic coronary arteries. N Engl J Med. 1986;315:1046 1051.
117. Hambrecht R, Wolff A, Gielen S, Linke A, Hofer J, Erbs S, Schoene N,
Schuler G. Effect of exercise on coronary endothelial function in
patients with coronary artery disease. N Engl J Med. 2000;342:
454 460.
118. Gielen S, Erbs S, Linke A, Mobius-Winkler S, Schuler G, Hambrecht R.
Home-based versus hospital-based exercise programs in patients with
coronary artery disease: effects on coronary vasomotion. Am Heart J.
2003;145:E3.
119. Gokce N, Vita JA, Bader DS, Sherman DL, Hunter LM, Holbrook M,
OMalley JF, Balady GJ. Effect of exercise on upper and lower
extremity endothelial function in patients with coronary artery disease.
Am J Cardiol. 2002;90:124 127.
120. Linke A, Schoene N, Gielen S, Hofer J, Erbs S, Schuler G, Hambrecht
R. Endothelial dysfunction in patients with chronic heart failure:
systemic effects of lower-limb exercise training. J Am Coll Cardiol.
2001;37:392397.
121. Davis PF, Helmke BP. Endothelial mechanotransduction. In: Mofrad
MRK, Kamm RD, eds. Cellular Mechanotransduction: Diverse Perspectives From Molecules to Tissue. New York, NY: Cambridge University Press; 2010:20 60.
122. Li YS, Haga JM, Chien S. Molecular basis of the effects of shear stress
on vascular endothelial cells. J Biomech. 2005;38:1949 1971.
123. Tzima E, Irani-Tehrani M, Kiosses WB, Dejana E, Schultz DA,
Engelhardt B, Cao G, DeLisser H, Schwartz MA. A mechanosensory
complex that mediates the endothelial cell response to fluid shear stress.
Nature. 2005;437:426 431.
124. Boger RH, Bode-Boger SM, Szuba A, Tsao PS, Chan JR, Tangphao O,
Blaschke TF, Cooke JP. Asymmetric dimethylarginine (ADMA): a
novel risk factor for endothelial dysfunction: its role in hypercholesterolemia. Circulation. 1998;98:18421847.
125. Wassmann S, Wassmann K, Nickenig G. Modulation of oxidant and antioxidant enzyme expression and function in vascular cells. Hypertension.
2004;44:381386.
126. Crosby JR, Kaminski WE, Schatteman G, Martin PJ, Raines EW, Seifert
RA, Bowen-Pope DF. Endothelial cells of hematopoietic origin make a
significant contribution to adult blood vessel formation. Circ Res. 2000;
87:728 730.
127. Forstermann U, Munzel T. Endothelial nitric oxide synthase in vascular
disease. Circulation. 2006;113:1708 1714.
128. Widder JD, Chen W, Li L, Dikalov S, Thony B, Hatakeyama K,
Harrison DG. Regulation of tetrahydrobiopterin biosynthesis by shear
stress. Circ Res. 2007;101:830 838.
129. Lam CF, Peterson TE, Richardson DM, Croatt AJ, dUscio LV, Nath
KA, Katusic ZS. Increased blood flow causes coordinated upregulation
of arterial eNOS and biosynthesis of tetrahydrobiopterin. Am J Physiol.
2006;290:H786 H793.
130. Sindler AL, Delp MD, Reyes R, Wu G, Muller-Delp JM. Effects of
ageing and exercise training on eNOS uncoupling in skeletal muscle
resistance arterioles. J Physiol. 2009;587:38853897.
131. Richter B, Niessner A, Penka M, Grdic M, Steiner S, Strasser B, Ziegler
S, Zorn G, Maurer G, Simeon-Rudolf V, Wojta J, Huber K. Endurance
training reduces circulating asymetric dimethylarginine and myeloperoxidase levels in persons at risk of coronary events. Thromb Haemost.
2005;94:1306 1311.
132. Hambrecht R, Adams V, Erbs S, Linke A, Krankel N, Shu Y, Baither
Y, Gielen S, Gummert JF, Mohr FW, Schuler G. Regular physical
activity improves endothelial function in patients with coronary

Gielen et al

133.

134.

135.

136.

137.

138.

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

139.
140.

141.

142.

143.

144.

145.

146.

147.

148.

149.

150.

artery disease by increasing phosphorylation of endothelial nitric


oxide synthase. Circulation. 2003;107:31523158.
Durrant JR, Seals DR, Connell ML, Russell MJ, Lawson BR, Folian BJ,
Donato AJ, Lesniewski LA. Voluntary wheel running restores endothelial function in conduit arteries of old mice: direct evidence for reduced
oxidative stress, increased superoxide dismutase activity and downregulation of NADPH oxidase. J Physiol. 2009;587:32713285.
Fukai T, Siegfried MR, Ushio-Fukai M, Cheng Y, Kojda G, Harrison
DG. Regulation of the vascular extracellular superoxide dismutase by
nitric oxide and exercise training. J Clin Invest. 2000;105:16311639.
Ennezat PV, Malendowicz LS, Testa M, Colombo PC, Cohen-Solal A,
Evans T, LeJemetel TH. Physical training in patients with chronic heart
failure enhances the expression of genes antioxidative enzymes. J Am
Coll Cardiol. 2001;38:194 198.
Zanchi NE, Grassmann-Bechera LR, Tanaka LY, Debbas V, Bartholomeu T, Ramirs PR. Moderate exercise training decreases aortic
superoxide production in myocardial infarcted rats. Eur J Appl Physiol.
2008;104:10451052.
Seshiah PN, Weber DS, Rocic P, Valppu L, Taniyama Y, Griendling
KK. Angiotensin II stimulation of NAD(P)H oxidase activity: upstream
mediators. Circ Res. 2002;91:406 413.
Asahara T, Murohara T, Sullivan A, Silver M, van der Zee R, Li T,
Witzenbichler B, Schatteman G, Isner JM. Isolation of putative progenitor endothelial cells for angioneogenesis. Science. 1997;275:
964 967.
Huang NF, Li S. Mesenchymal stem cells for vascular regeneration.
Regen Med. 2008;3:877 892.
Werner N, Kosiol S, Schiegl T, Ahlers P, Walenta K, Link A, Bohm M,
Nickenig G. Circulating endothelial progenitor cells and cardiovascular
outcomes. N Engl J Med. 2005;353:999 1007.
Adams V, Lenk K, Linke A, Lenz D, Erbs S, Sandri M, Tarnok A,
Gielen S, Emmrich F, Schuler G, Hambrecht R. Increase of circulating
endothelial progenitor cells in patients with coronary artery disease after
exercise-induced ischemia. Arterioscler Thromb Vasc Biol. 2004;24:
684 690.
Sandri M, Adams V, Gielen S, Linke A, Lenk K, Krankel N, Lenz D,
Erbs S, Scheinert D, Mohr FW, Schuler G, Hambrecht R. Effects of
exercise and ischemia on mobilization and functional activation of
blood-derived progenitor cells in patients with ischemic syndromes:
results of 3 randomized studies. Circulation. 2005;111:33913399.
Shaffer RG, Greene S, Arshi A, Supple G, Bantly A, Moores JS,
Parmacek MS, Mohler ER III. Effect of acute exercise on endothelial
progenitor cells in patients with peripheral arterial disease. Vasc Med.
2006;11:219 226.
Van Craenenbroeck EMF, Vrints CJ, Haine SE, Vermeulen K, Goovaerts I, Van Tendeloo VFI, Hoymans VY, Conraads VMA. A maximal
exercise bout increases the number of circulating CD34/KDR endothelial progenitor cells in healthy subjects: relation with lipid profile.
J Appl Physiol. 2008;104:1006 1013.
Bonsignore MR, Morici G, Santoro A, Pagano M, Cascio L, Bonanno A,
Abate P, Mirabella F, Profita M, Insalaco G, Giola M, Vignola AM,
Majolino I, Testa U, Hogg JC. Circulating hematopoietic progenitor
cells in runners. J Appl Physiol. 2002;93:16911697.
Adams V, Linke A, Breuckmann F, Leineweber K, Erbs S, Krankel N,
Brocker-Preuss M, Woitek F, Erbel R, Heusch G, Hambrecht R, Schuler
G, Mohlenkamp S. Circulating progenitor cells decrease immediately
after marathon race in advanced-age marathon runners. Eur J Cardiovasc Prev Rehabil. 2008;15:602 607.
Mobius-Winkler S, Hilberg T, Menzel K, Golla E, Burman A, Schuler
G, Adams V. Time-dependent mobilization of circulating progenitor
cells during strenuous exercise in healthy individuals. J Appl Physiol.
2009;107:19431950.
Goussetis E, Spiropoulos A, Tsironi M, Skenderi K, Margeli A, Graphakos S, Baltopoulos P, Papassotiriou I. Spartathlon, a 246 kilometer
foot race: effects of acute inflammation induced by prolonged exercise
on circulating progenitor reparative cells. Blood Cells Mol Dis. 2009;
42:294 299.
Asahara T, Takahashi T, Masuda H. VEGF contributes to postnatal
neovascularization by mobilizing bone-marrow endothelial progenitor
cells. EMBO J. 1999;18:3964 3972.
Li B, Sharpe EE, Maupin AB, Teleron AA, Pyle AL, Carmeliet P,
Young PP. VEGF and PlGF promote adult vasculogenesis by enhancing
EPC recruitment and vessel formation at the site of tumor neovascularization. FASEB J. 2006;20:14951497.

Cardiovascular Effects of Exercise Training

1237

151. Cheng XW, Kuzuya M, Nakamura K, Maeda K, Tsuzuki M, Kim W,


Sasaki T, Liu Z, Inoue N, Kondo T, Jin H, Numaguchi Y, Okumura K,
Yokota M, Iguchi A, Murohara T. Mechanisms underlying the
impairment of ischemia-induced neovascularization in matrix metalloproteinase 2-deficient mice. Circ Res. 2007;100:904 913.
152. Huang PH, Chen YH, Wang CH, Chen JS, Tsai HY, Lin FY, Lo WY,
Wu TC, Sata M, Chen JW, Lin SJ. Matrix metalloproteinase-9 is
essential for ischemia-induced neovascularization by modulating bone
marrow derived endothelial progenitor cells. Arterioscler Thromb Vasc
Biol. 2009;29:1179 1184.
153. Lapidot T, Petit I. Current understanding of stem cell mobilization: the
roles of chemokines, proteolytic enzymes, adhesion molecules, cytokines, and stromal cells. Exp Hematol. 2002;30:973981.
154. Heissig B, Hattori K, Dias S, Friedrich M, Ferris B, Hackett NR, Crystal
RG, Besmer P, Lyden D, Moore MA, Werb Z, Rafii S. Recruitment of
stem and progenitor cells from the bone marrow nighe requires MMP-9
mediated release of kit-ligand. Cell. 2002;109:625 637.
155. Iwakura A, Shastry S, Luedemann C, Hamada H, Kawamoto A, Kishore
R, Zhu Y, Qin G, Silver M, Thorne T, Eaton L, Masuda H, Asahara T,
Losordo DW. Estradiol enhances recovery after myocardial infarction
by augmenting incorporation of bone marrow derived endothelial progenitor cells into sites of ischemia-induced neovascularization via endothelial nitric oxide synthasemediated activation of matrix
metalloproteinase-9. Circulation. 2006;113:16051614.
156. Askari AT, Unzek S, Popovic ZB, Goldman CK, Forudi F, Kiedrowski
M, Rovner A, Ellis SG, Thomas JD, DiCorleto PE, Topol EJ, Penn MS.
Effect of stromal-cell-derived factor 1 on stem-cell homing and tissue
regeneration in ischaemic cardiomyopathy. Lancet. 2003;362:697703.
157. Tachibana K, Hirota S, Iizasa H, Yoshida H, Kawabata K, Kataoka Y,
Kitamura Y, Matsushima K, Yoshida N, Nishikawa Si, Kishimoto T,
Nagasawa T. The chemokine receptor CXCR4 is essential for vascularization of the gastrointestinal tract. Nature. 1998;393:591594.
158. Schuler G, Hambrecht R, Schlierf G, Niebauer J, Hauer K, Neumann J,
Hoberg E, Drinkmann A, Bacher F, Grunze M, Kubler W. Regular
physical exercise and low-fat diet: effects on progression of coronary
artery disease. Circulation. 1992;86:111.
159. Niebauer J, Hambrecht R, Velich T, Hauer K, Marburger C, Kalberer
B, Weiss C, von Hodenberg E, Schlierf G, Schuler G, Zimmermann
R, Kubler W. Attenuated progression of coronary artery disease after
6 years of multifactorial risk intervention: role of physical exercise.
Circulation. 1997;96:2534 2541.
160. Ornish D, Brown SE, Scherwitz LW, Billings JH, Armstrong WT, Ports
TA, McLanahan SM, Kirkeeide RL, Brand RJ, Gould KL. Can lifestyle
changes reverse coronary heart disease? The Lifestyle Heart Trial.
Lancet. 1990;336:129 133.
161. Haskell WL, Alderman EL, Fair JM, Maron DJ, Mackey SF, Superko R,
Williams PT, Johnstone IM, Champagne MA, Krauss RM, Farquhar JW.
Effects of intensive multiple risk factor reduction on coronary atherosclerosis and clinical cardiac events in men and women with coronary
artery disease: the Stanford Coronary Risk Intervention Project (SCRIP).
Circulation. 1994;89:975990.
162. Deussen A, Brand M, Pexa A, Weichsel J. Metabolic coronary flow
regulation: current concepts. Basic Res Cardiol. 2006;101:453 464.
163. Busse R, Forstermann U, Matsuda H, Pohl U. The role of prostaglandins
in the endothelium-mediated vasodilatory response to hypoxia. Pfugers
Arch. 1984;401:77 83.
164. Bouchard JF, Dumont E, Lamontagne D. Evidence that prostaglandins
I2, E2, and D2 may activate ATP sensitive potassium channels in the
isolated rat heart. Cardiovasc Res. 1994;28:901905.
165. Kroll K, Kinzie DJ, Gustafson LA. Open-system kinetics of myocardial
phosphoenergetics during coronary underperfusion. Am J Physiol. 1997;
272:H2563H2576.
166. Feigl EO. Coronary physiology. Physiol Rev. 1983;63:1205.
167. Dietrich HH, Ellsworth ML, Sprague RS, Dacey RG Jr. Red blood cell
regulation of microvascular tone through adenosine triphosphate. Am J
Physiol. 2000;278:H1294 H1298.
168. DiCarlo SE, Blair RW, Bishop VS, Stone HL. Daily exercise enhances
coronary resistance vessel sensitivity to pharmacological activation.
J Appl Physiol. 1989;66:421 428.
169. Laughlin MH, Overholser KA, Bhatte MJ. Exercise training increases
coronary transport reserve in miniature swine. J Appl Physiol. 1989;67:
1140 1149.
170. Muller JM, Myers PR, Laughlin MH. Exercise training alters myogenic
responses in porcine coronary resistance arteries. J Appl Physiol. 1993;
75:26772682.

1238

Circulation

September 21, 2010

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

171. Karibe A, Watanabe J, Horiguchi S, Takeuchi M, Suzuki S, Funakoshi


M, Katoh H, Keitoku M, Satoh S, Shirato K. Role of cytosolic Ca2 and
protein kinase C in developing myogenic contraction in isolated rat
small arteries. Am J Physiol. 1997;272:H1165H1172.
172. Korzick DH, Laughlin MH, Bowles DK. Alterations in PKC signaling
underlie enhanced myogenic tone in exercise-trained porcine coronary
resistance arteries. J Appl Physiol. 2004;96:14251432.
173. White FC, Bloor CM, McKirnan MD, Carroll SM. Exercise training in
swine promotes growth of arteriolar bed and capillary angiogenesis in
heart. J Appl Physiol. 1998;85:1160 1168.
174. Carrow RE, Brown RE, Van Huss WD. Fiber sizes and capillary to fiber
ratios in skeletal muscle of exercised rats. Anat Rec. 1967;159:3339.
175. Ingjer F. Effects of endurance training on muscle fiber ATP-ase activity,
capillary supply and mitochondrial content in man. J Physiol. 1979;294:
419 432.
176. Ferrara N, Gerber HP, LeCouter J. The biology of VEGF and its
receptors. Nat Med. 2003;9:669 676.
177. Tang K, Xia FC, Wagner PD, Breen EC. Exercise-induced VEGF
transcriptional activation in brain, lung and skeletal muscle. Respir
Physiol Neurobiol. 2010;170:16 22.
178. Mason SD, Rundqvist H, Papandreou I, Duh R, McNulty WJ, Howlett
RA, Olfert IM, Sundberg CJ, Denko NC, Poellinger L, Johnson RS.
HIF-1{alpha} in endurance training: suppression of oxidative metabolism. Am J Physiol. 2007;293:R2059 R2069.
179. Chinsomboon J, Ruas J, Gupta RK, Thom R, Shoag J, Rowe GC,
Sawada N, Raghuram S, Arany Z. The transcriptional coactivator
PGC-1 mediates exercise-induced angiogenesis in skeletal muscle.
Proc Natl Acad Sci U S A. 2009;106:2140121406.
180. Arany Z, Foo SY, Ma Y, Ruas JL, Bommi-Reddy A, Girnun G, Cooper
M, Laznik D, Chinsomboon J, Rangwala SM, Baek KH, Rosenzweig A,
Spiegelman BM. HIF-independent regulation of VEGF and angiogenesis by the transcriptional coactivator PGC-1. Nature. 2008;451:
1008 1012.
181. Richter EA, Ruderman NB. AMPK and the biochemistry of exercise:
implications for human health and disease. Biochem J. 2009;418:
261275.
182. Zwetsloot KA, Westerkamp LM, Holmes BF, Gavin TP. AMPK regulates basal skeletal muscle capillarization and VEGF expression, but is
not necessary for the angiogenic response to exercise. J Physiol. 2008;
586:6021 6035.
183. Akimoto T, Pohnert SC, Li P, Zhang M, Gumbs C, Rosenberg PB,
Williams RS, Yan Z. Exercise stimulates PGC-1alpha transcription in
skeletal muscle through activation of the p39MAPK pathway. J Biol
Chem. 2005;280:1958719593.
184. Kang C, OMoore KM, Dickman JR, Ji LL. Exercise activation of
muscle peroxisome proliferator-activated receptor-[gamma] coactivator1[alpha] signaling is redox sensitive. Free Radic Biol Med. 2009;47:
1394 1400.
185. Gielen S, Hambrecht R. Venous function in chronic heart failure: the
neglected part of circulatory regulation. J Cardiopulm Rehabil. 2002;
22:327328.
186. Monahan KD, Dinenno FA, Seals DR, Halliwill JR. Smaller ageassociated reductions in leg venous compliance in endurance exercisetrained men. Am J Physiol. 2001;281:H1267H1273.
187. van Duijnhoven N, Bleeker M, de Groot P, Thijssen D, Felsenberg D,
Rittweger J, Hopman M. The effect of bed rest and an exercise countermeasure on leg venous function. Eur J Appl Physiol. 2008;104:
991998.

188. Welsch M, Alomari M, Parish TR, Wood RH, Kalb D. Influence of


venous function on exercise tolerance in chronic heart failure. J Cardiopulm Rehabil. 2002;22:321326.
189. Maiorana A, ODriscoll G, Dembo L, Cheetham C, Goodman C, Taylor
R, Green D. Effect of aerobic and resistance exercise training on
vascular function in heart failure. Am J Physiol. 2000;279:
H1999 H2005.
190. Roveda F, Middlekauff HR, Rondon MU, Reis SF, Souza M, Nastari L,
Barretto AC, Krieger EM, Negrao CE. The effects of exercise training
on sympathetic neural activation in advanced heart failure: a randomized
controlled trial. J Am Coll Cardiol. 2003;42:854 860.
191. Suzuki J, Kobayashi T, Uruma T, Koyama T. Time-course changes in
arteriolar and venular portions of capillary in young treadmill-trained
rats. Acta Physiol Scand. 2001;171:77 86.
192. Lee AP, Ice R, Blessey R, Sanmarco ME. Long-term effect of physical training
on coronary patients with impaired ventricular function. Circulation. 1979;
60:15191526.
193. Dubach P, Myers J, Dzienkan G, Goebbels U, Reinhart W, Muller P,
Buser P, Stulz P, Vogt P, Ratti R. Effect of high intensity exercise
training on central hemodynamic response to exercise in men with
reduced left ventricular function. J Am Coll Cardiol. 1997;29:
15911598.
194. Mereles D, Ehlken N, Kreuscher S, Ghofrani S, Hoeper MM, Halank M,
Meyer FJ, Karger G, Buss J, Juenger J, Holzapfel N, Opitz C, Winkler
J, Herth FFJ, Wilkens H, Katus HA, Olschewski H, Grunig E. Exercise
and respiratory training improve exercise capacity and quality of life in
patients with severe chronic pulmonary hypertension. Circulation. 2006;
114:14821489.
195. Kashimura O, Sakai A. Effects of physical training on pulmonary
arterial pressure during exercise under hypobaric hypoxia in rats. Int
J Biometerol. 1991;35:214 221.
196. Kashimura O, Sakai A, Yanagidaira Y. Effects of exercise-training on
hypoxia and angiotensin II-induced pulmonary vasoconstrictions. Acta
Physiol Scand. 1995;155:291295.
197. Favret F, Henderson KK, Allen J, Richalet JP, Gonzalez NC. Exercise
training improves lung gas exchange and attenuates acute hypoxic pulmonary hypertension but does not prevent pulmonary hypertension of
prolonged hypoxia. J Appl Physiol. 2006;100:20 25.
198. Chen HI, Li HT. Physical conditioning can modulate endotheliumdependent vasorelaxation in rabbits. Arterioscler Thromb Vasc Biol.
1993;13:852 856.
199. Mitani Y, Maruyama J, Maruyama K, Sakurai M. Exercise training does
not alter acetylcholine-induced responses in isolated pulmonary artery
from rat. Eur Respir J. 1999;13:622 625.
200. Johnson LR, Parker JL, Laughlin MH. Chronic exercise training
improves ACh-induced vasorelaxation in pulmonary arteries of pigs.
J Appl Physiol. 2000;88:443 451.
201. Johnson LR, Rush JWE, Turk JR, Price EM, Laughlin MH. Short-term
exercise training increases ACh-induced relaxation and eNOS protein in
porcine pulmonary arteries. J Appl Physiol 2001;90:11021110.
202. Handoko ML, de Man FS, Happe CM, Schalij I, Musters RJP,
Westerhof N, Postmus PE, Paulus WJ, van der Laarse WJ, VonkNoordegraaf A. Opposite effects of training in rats with stable and
progressive pulmonary hypertension. Circulation. 2009;120:42 49.
KEY WORDS: coronary disease

endothelium

exercise

myocardium

Cardiovascular Effects of Exercise Training: Molecular Mechanisms


Stephan Gielen, Gerhard Schuler and Volker Adams

Downloaded from http://circ.ahajournals.org/ by guest on January 16, 2017

Circulation. 2010;122:1221-1238
doi: 10.1161/CIRCULATIONAHA.110.939959
Circulation is published by the American Heart Association, 7272 Greenville Avenue, Dallas, TX 75231
Copyright 2010 American Heart Association, Inc. All rights reserved.
Print ISSN: 0009-7322. Online ISSN: 1524-4539

The online version of this article, along with updated information and services, is located on the
World Wide Web at:
http://circ.ahajournals.org/content/122/12/1221

Permissions: Requests for permissions to reproduce figures, tables, or portions of articles originally published
in Circulation can be obtained via RightsLink, a service of the Copyright Clearance Center, not the Editorial
Office. Once the online version of the published article for which permission is being requested is located,
click Request Permissions in the middle column of the Web page under Services. Further information about
this process is available in the Permissions and Rights Question and Answer document.
Reprints: Information about reprints can be found online at:
http://www.lww.com/reprints
Subscriptions: Information about subscribing to Circulation is online at:
http://circ.ahajournals.org//subscriptions/

You might also like