ArticlePDF AvailableLiterature Review

Molecular pathways of arterial aging

Authors:

Abstract and Figures

The incidence of stroke and myocardial infarction increases in aged patients and it is associated with an adverse outcome. Considering the aging population and the increasing incidence of cardiovascular disease, the prediction for population well-being and health economics is daunting. Accordingly, there is an unmet need to focus on fundamental processes underlying vascular aging. A better understanding of the pathways leading to arterial aging may contribute to design mechanism-based therapeutic approaches to prevent or attenuate features of vascular senescence. In the present review, we discuss advances in the pathophysiology of age-related vascular dysfunction including nitric oxide signalling, dysregulation of oxidant/inflammatory genes, epigenetic modifications and mechanisms of vascular calcification as well as insights into vascular repair. Such an overview highlights attractive molecular targets for the prevention of age-driven vascular disease.
Content may be subject to copyright.
Clinical Science (2015) 128, 69–79 (Printed in Great Britain) doi: 10.1042/CS20140302
Molecular pathways of arterial aging
Francesco Paneni, Sarah Costantinoand Francesco Cosentino
Cardiology Unit, Department of Medicine, Karolinska University Hospital, Stockholm, Sweden
Abstract
The incidence of stroke and myocardial infarction increases in aged patients and it is associated with an adverse
outcome. Considering the aging population and the increasing incidence of cardiovascular disease, the prediction
for population well-being and health economics is daunting. Accordingly, there is an unmet need to focus on
fundamental processes underlying vascular aging. A better understanding of the pathways leading to arterial aging
may contribute to design mechanism-based therapeutic approaches to prevent or attenuate features of vascular
senescence. In the present review, we discuss advances in the pathophysiology of age-related vascular dysfunction
including nitric oxide signalling, dysregulation of oxidant/inflammatory genes, epigenetic modifications and
mechanisms of vascular calcification as well as insights into vascular repair. Such an overview highlights attractive
molecular targets for the prevention of age-driven vascular disease.
Key words: arterial aging, inflammation, oxidative stress, pathways, vascular repair
AGING AND CARDIOVASCULAR DISEASE
Aging is the main risk factor and driver of incident cardiovas-
cular disease (CVD) and results in a progressive functional and
structural decline of the vasculature [1]. Arterial aging, through
its impact on physical and mental health, impairs quality of life
and the ability of individuals to carry out the tasks of every-
day life [2]. As the average lifespan of the European population
is increasing, the overall future CVD burden in Europe is pre-
dicted to increase dramatically leading to a pandemic of frailty
syndromes, poor quality of life and high morbidity in the aging
population [3,4]. Since half of all people aged 65 years and older
develop CVD, arterial aging has a profound impact on the health
of populations in Europe [5,6]. Moreover, clustering of chrono-
logical age with other cardiovascular (CV) risk factors clearly
anticipates senescent features of heart and vessels [7]. Intricate
signalling cascades are emerging as determinants of accelerated
arterial aging in the presence of CV risk factors. Accordingly,
there is an unmet need to focus on fundamental processes of
vascular aging enabling early diagnosis and prevention of CVD.
This is important in the context that, although aging is inevitable,
vascular aging is a modifiable risk factor.
Abbreviations: ALDH2, aldehyde dehydrogenase 2; AMPK, AMP-activated protein kinase; AP-1, activated protein 1; ApoE, apolipoprotein E; Arg II, arginase II; BAP, bone alkaline
phosphatase; BH4, tetrahydrobiopterin; CV, cardiovascular; CVD, cardiovascular disease; eNOS, endothelial nitric oxide synthase; EOC, early angiogenic outgrowth cell; EPC,
endothelial progenitor cell; fOXO, Forkhead box O; HIF-1α, hypoxia inducible factor 1α; MCC, myeloid calcifying cell; NF-κB, nuclear factor κB; OC, osteocalcin; PI3K, phosphoinositide
3-kinase; RANKL, receptor activator of NF-κB ligand; ROS, reactive oxygen species; SDF-1, stromal cell-derived factor 1; VSMC, vascular smooth muscle cell; WT, wild-type.
Correspondence: Professor Francesco Cosentino (email francesco.cosentino@ki.se).
STRUCTURAL ABNORMALITIES OF ARTERIAL
AGING
Clinical and pre-clinical data have shown that aging is associ-
ated with structural and functional properties of large arteries
[8]. Aging blood vessels become thicker and stiffer, resulting in
a reduced ability to adjust vessel shape and function to chan-
ging tissue demands. In aged healthy humans, these alterations
are represented by luminal dilation, increased arterial stiffness,
endothelial dysfunction and diffuse intimal thickening. A de-
crease in vascular plasticity may be the result of different factors
including enhanced elastin degradation and collagen deposition
in the vascular media [9]. Importantly, remodelling of the vas-
culature is accompanied by key alterations of endothelial ho-
moeostasis [10]. Indeed, seminal studies have clearly shown that
endothelium-dependent vasorelaxation is impaired in aged ves-
sels and this phenomenon is associated with increased vascular
permeability and inflammation, as well as impaired angiogenesis
[1,11]. Moreover, age-related alterations of endothelial cell func-
tionality may, in turn, aggravate media thickness and vascular
fibrosis.
C
The Authors Journal compilation C
2015 Biochemical Society 69
Clinical Science www.clinsci.org
F. Paneni, S. Costantino and F. Cosentino
NITRIC OXIDE SIGNALLING
Vascular aging, characterized by endothelial dysfunction and in-
creased vascular stiffness, is associated with reduced endothelial
nitric oxide (NO) bioavailability and increased generation of re-
active oxygen species (ROS) [12]. Increased ROS production
in vascular aging derives from enzymatic and non-enzymatic
sources such as mitochondria [13]. Interestingly enough, dysreg-
ulation of the endothelial nitric oxide synthase (eNOS), known as
eNOS uncoupling, results in loss of endothelial NO generation
and increased ROS production [14]. The mechanism of eNOS-
uncoupling seems multiple and includes oxidation of the cofactor
tetrahydrobiopterin (BH4), decreased intracellular availability of
the substrate L-arginine due to either increased arginase activ-
ity or accumulation of endogenous arginine analogues such as
asymmetric dimethyl-L-arginine that competes with L-arginine
for eNOS binding [15].
Tetrahydrobiopterin
The small molecule BH4is a redox cofactor for eNOS and a
pivotal regulator of its catalytic activity [14]. Loss of BH4is
associated with vascular disease states, including aging, and res-
ults in eNOS uncoupling reduced NO release and superoxide
anion (O2) generation, in turn, leading to production of hydro-
gen peroxide and peroxynitrite (ONOO) (Figure 1) [12]. The
role of BH4in vascular aging has been demonstrated by sev-
eral studies over the last 10 years. Older animals have reduced
vascular BH4levels due to its oxidation into dihydrobiopterin
(BH2), which is not acting as a cofactor for eNOS. Such a short-
age of cofactor leads to a conformational change in eNOS from
a dimeric to monomeric state resulting in loss of NO produc-
tion and ROS generation [15]. Pharmacological supplementation
of BH4can improve endothelial function in aged humans com-
pared with young subjects [16], suggesting that BH4treatment
might be a rational approach to improve vascular function in
aging [17].
Arginase II
L-Arginine, a major eNOS substrate, is rapidly metabolized by
arginase enzyme to urea and L-ornithine, leading to a decrease
in its availability and, hence, reduced NO synthesis [18]. An
increasing body of evidence suggests that arginase II (Arg II)
is deregulated in aging and participates to early vascular senes-
cence by altering the phenotype of endothelial cells and vascular
smooth muscle cells (VSMCs). A recent study performed in VS-
MCs isolated from human umbilical veins showed that Arg II
overexpression activates p66Shc/p53 signalling, thus triggering
mitochondrial dysfunction and cell apoptosis, key features of
cellular aging. Consistently, gene silencing of Arg II attenuates
senescence of human endothelial cells by maintaining a detri-
mental vicious cycle involving the mammalian target of rapamy-
cin complex 1 (mTORC1)/S6 kinase (S6K1) pathway.
Activation of Arg II may represent an important link between
CV risk factors and the aging process. Indeed, several common
risk factors such as hyperglycaemia, hyperlipidaemia and obesity
have been shown to increase Arg II expression/activity. Genetic
deletion of Arg II reduces vascular inflammation in mice fed on
a high-cholesterol and high-fat diet, prevents atherosclerosis and
ameliorates insulin sensitivity as well as glucose homoeostasis
[19,20]. These findings shed light on the role of Arg II as a
common determinant across the spectrum of metabolic disease,
atherosclerosis and senescence. The clinical impact of this gene is
supported by the notion that arginase inhibition improves brachial
artery vasodilation in subjects with Type 2 diabetes [21]. Phar-
macological suppression of arginase activity may indeed restore
physiological substrate levels, thus suppressing molecular path-
ways involved in endothelial aging and dysfunction (Figure 1).
Endothelial nitric oxide synthase
Alteration of eNOS transcription may be involved in age-related
endothelial dysfunction but its biological significance remains
unclear. Some studies found increased eNOS expression in aging
and such observations have been interpreted as a compensatory
but futile mechanism to counter-regulate the loss of NO [2,22].
In contrast, other work in humans and rodents did not show
significant changes in eNOS expression but rather an increase
or decrease in its activating Ser1177 phosphorylation [15,23,24].
Age-dependent impairment of NO release also plays an import-
ant role in the coronary circulation [25]. Indeed, eNOS deregu-
lation is critically involved in microvascular dysfunction and im-
paired ventricular contractility in aged patients [26]. Moreover,
advanced age is associated with less collaterals to the infarct-
related artery in patients with acute myocardial infarction [27].
This abnormality may contribute to the poor prognosis of older
patients with acute coronary syndromes. Collectively, these evid-
ence strongly suggest that modulation of eNOS functionality is a
promising target for prevention of age-related vascular disease.
OXIDATIVE STRESS
A major feature of CV aging is the imbalance between NO
bioavailability and accumulation of ROS, leading to endothelial
dysfunction [28]. Indeed, age-dependent generation of O2in-
activates NO to form ONOO, a powerful oxidant that easily
penetrates across phospholipid membranes leading to substrate
nitration [22,29]. Protein nitrosylation blunts activity of antiox-
idant enzymes and eNOS [30]. Generation of ROS into the vessel
wall has been postulated as a major pathophysiological step fa-
vouring arterial aging [1,11]. Among the different aging theories,
the mitochondrial free radical theory has been in the spotlight for
several decades [31]. Accordingly, ROS are considered to be
by-products of aerobic metabolism that induce oxidative cellular
damage [32]. ROS, small highly reactive molecules, are gener-
ated within the cell by means of several metabolic and enzymatic
pathways. The majority of ROS are produced within mitochon-
drial oxidative phosphorylation. During this process, electrons
are extracted from NADH and FADH and transferred to molecu-
lar oxygen through a chain of four enzymatic complexes ensuring
phosphorylation of ADP into ATP and final reduction of molecu-
lar oxygen to water. However, electrons derived from NADH or
FADH can directly react with oxygen or other electron accept-
ors within the mitochondrial electron transport chain upstream
70 C
The Authors Journal compilation C
2015 Biochemical Society
Molecular pathways of arterial aging
Figure 1 Schematic representation of molecular pathways involved in arterial aging
VCAM-1, vascular cell adhesion molecule-1; ICAM-1, intercellular adhesion molecule-1; MCP-1, monocyte chemoattractant
protein-1; MnSOD, manganese superoxide dismutase; ecSOD, extracellular superoxide dismutase.
the last enzymatic complex (complex IV, which is responsible
for the reduction of molecular oxygen to water) and generate
free radicals [33]. Accumulation of ROS causes mitochondrial
disruption leading to cytochrome crelease and subsequent activ-
ation of caspase 3 [1,33]. CV risk factors such as hyperglycaemia
induce a ROS-dependent alteration of the mitochondrial net-
work resembling early signs of vascular aging [28]. Indeed, en-
dothelial cells isolated from middle-age diabetic subjects show
a premature derangement of organelle structure, which correl-
ates with impaired flow-mediated vasodilation of the brachial
artery [34].
NAPDH oxidase
NADPH oxidases of the Nox family are differentially expressed
in the CV system and may critically contribute to oxidative bur-
den and vascular disease [35]. The NOX family consists of seven
members: the classic NOXs, NOX1–NOX5, and the dual oxi-
dases Duox1 and Duox2 [36,37]. A significant expression in the
CV system has been reported for Nox1, Nox2, Nox4 and Nox5
[37]. As a consequence of the interaction between the differ-
ent ROS-generating systems such as mitochondria and eNOS,
NADPH oxidases have been demonstrated to contribute to ROS
formation for almost all risk factor conditions. However, the con-
tribution of NADPH oxidase in aging remains incompletely un-
derstood. This is likely to be due to the fact that the role of
NADPH in cellular senescence has been overlooked, since the
overproduction of ROS associated with aging has mainly been
attributed to leakiness of the mitochondrial respiratory chain.
Indeed, most of the available studies suggest that arterial aging
is rather the result of impaired antioxidant enzyme expression,
eNOS uncoupling, or increased ROS production by mitochon-
dria [1]. However, few studies support a putative involvement of
NOXs in arterial aging. A study has demonstrated that silencing
of NOX4 in human endothelial cells leads to delayed replicat-
ive senescence and preserves cell functionality [38]. In line with
these findings, we have recently shown that Nox2 and Nox4 are
significantly up-regulated in the vasculature of aged mice, as well
as only Nox2 increases in peripheral blood monocytes isolated
from older individuals [39]. Furthermore, NADPH oxidase is a
key mediator of endothelin-1-induced vasoconstriction, as well as
cerebrovascular damage in aged rodents [40,41]. In contrast, a
previous study showed that activation of NADPHsubunits p47phox
and p67phox remain unchanged in aged endothelial cells, suggest-
ing that age-related oxidative stress mechanistically differs from
endothelial dysfunction seen in the context of other CV risk
factors [42]. Taken together, these results suggest that further re-
search is needed to better understand the role of NADPH oxidase
in arterial aging.
C
The Authors Journal compilation C
2015 Biochemical Society 71
F. Paneni, S. Costantino and F. Cosentino
Activated protein-1 transcription factor JunD
Recently, the activated protein-1 (AP-1) transcription factor JunD
has emerging as an important modulator of age-driven mitochon-
drial oxidative stress [43]. AP-1 is a collection of dimeric com-
plexes made by different members of three families of DNA-
binding proteins: Jun, Fos and activating transcription factor
(ATF)/cAMP-response-element-binding protein-binding protein
(CREB) [44]. These members assemble to form AP-1 transcrip-
tion factors with activities that are strongly influenced by their
specific components and their cellular environment. JunD, the
most recent gene of the Jun family, regulates cell growth and sur-
vival and protects against oxidative stress by modulating genes
involved in antioxidant defence and ROS production [45]. Our
recent work has shown that JunD is a longevity gene implicated
in the preservation of vascular homoeostasis during life [39] (Fig-
ure 1). We found that JunD is down-regulated by aging both in
mouse aorta and peripheral blood monocytes of old as compared
with young healthy individuals. Interestingly, age-dependent re-
duction in JunD expression was explained by a specificepigenetic
signature, namely reduced methylation of CpG dinucleotides on
the JunD promoter [39]. This latter finding strengths the no-
tion that epigenetic signatures may critically participate in early
phenotypes of vascular disease during the lifetime [46,47]. In our
study, young JunD/mice showed early endothelial dysfunction
and vascular senescence, which were comparable with the one
observed in aged wild-type (WT) mice. JunD deletion was in-
deed associated with up-regulation of the aging markers p53 and
p16INK4a, reduced telomerase activity and mitochondrial DNA
damage [39]. Interestingly enough, transient overexpression of
JunD was able to rescue endothelial dysfunction in aged mice.
Mechanistically, we found that JunD is required for the expres-
sion of mitochondrial antioxidant enzymes such as manganese
superoxide dismutase (MnSOD) and aldehyde dehydrogenase 2
(ALDH-2). This latter enzyme was almost abolished in young
JunD/mice compared with age-matched littermates. The rel-
evance of this finding is supported by the notion that ALDH-2
protects against ischaemia/reperfusion injury [48] and cardiac
arrhythmias [49]. Moreover, JunD participates to eNOS tran-
scription, thus contributing to preserve NO availability during
aging [50]. The transcription factor also modulates the expression
and activity of NADPH oxidase in the vasculature [39]. Indeed,
JunD/mice display a premature up-regulation of the NADPH
isoforms p47phox, Nox2 and Nox4 leading to increased vascu-
lar oxidative stress already in early stages of life. Moreover, in
our study JunD expression negatively correlated with NADPH
subunits in aged individuals. In line with our findings, lack
of JunD promotes pressure-overload-induced apoptosis, hyper-
trophic growth and angiogenesis in the heart [51]. Collectively,
these findings indicate that JunD down-regulation may represent
a common molecular event linking premature senescence with
CVD development.
Mitochondrial adaptor p66Shc
The identification of molecular pathways modulating the en-
dothelial cell redox state is crucial in understanding the mech-
anisms linking vascular aging to endothelial dysfunction and
atherosclerosis. In view of its role on the cellular redox state, mi-
tochondrial p66Shc adaptor protein has been regarded as part of
a putative transduction pathway relevant to endothelial integrity
[11,52]. Intracellular free radicals are reduced in cells lacking the
p66Shc gene (p66Shc/cells), and both systemic as well as intra-
cellular free radicals are diminished in a p66Shc/mouse model
exposed to high oxidative stress [53,54]. The p66Shc adaptor pro-
tein functions as a redox enzyme implicated in mitochondrial
ROS generation and translation of oxidative signals into apop-
tosis (Figure 1) [53,55–57]. Several chronic stimuli activate the
protein kinase C βII (PKCβII) isoform to induce Ser36 phos-
phorylation of p66Shc, allowing transfer of the protein from the
cytosol to the mitochondrion where it catalyses ROS produc-
tion via cytochrome coxidation [53]. This latter event leads to
mitochondrial disruption and cell death. Indeed, increased ROS
generation alters mitochondrial permeability facilitating the re-
lease of proapoptotic proteins such as cytochrome c[53]. Once
into the cytosol, cytochrome cis responsible for activation of
the apoptosis execution enzyme caspase-3. Accordingly, mice
lacking the p66Shc gene (p66Shc/) display prolonged lifespan,
increased resistance to oxidative stress and apoptosis. We have
previously reported that endothelium-dependent relaxations in
response to acetylcholine were age-dependently impaired in WT
mice but not in p66Shc/mice [55]. Accordingly, p66Shc/mice
were protected against the age-related decline in NO release.
This study implies that p66Shc signalling is required to induce
a ROS-driven vascular senescent phenotype. Importantly, p66Shc
activation is thought to be upstream of the NADPH and mam-
malian target of rapamycin pathway, two important determinants
of vascular damage [58]. Activation of p66Shc also promotes the
heart senescent phenotype and development of heart failure in
diabetic mice [59]. Accordingly, diabetic p66Shc/mice are pro-
tected against myocardial oxidative stress, apoptosis and telomere
shortening. Despite these studies provided interesting insights
into the role of p66Shc in CV aging, it remains unclear whether
protein inhibition may reverse vascular disease phenotype. In this
regard, we have demonstrated for the first time that in vivo RNA
interference blunts vascular p66Shc expression, ROS generation
and endothelial dysfunction in diabetic mice [60]. Moreover, ex-
perimental studies have recently shown that genetic deletion of
p66Shc prevents age-related CV disease such as myocardial in-
farction and stroke [56,61]. Of interest, p66Shc expression is also
affected by CV risk factors such as oxidized low-density lipopro-
tein and blood pressure [62,63]. These latter findings imply that
the adaptor p66Shc may represent a critical molecular intermediate
between CV risk factors and premature aging. The clinical relev-
ance of p66Shc is supported by the notion that p66Shc gene expres-
sion is increased in mononuclear cells obtained from patients with
Type 2 diabetes and coronary artery disease [64,65]. A study has
shown that p66Shc expression is higher in fibroblasts isolated from
centenarians [66]. This finding probably indicates that p66Shc
expression may increase in a time-dependent manner. In con-
trast, early gene up-regulation due to risk factors may anticip-
ate features of CV aging in middle-age individuals [64,67]. On
the whole, the evidence reported so far supports the concept
that modulation of p66Shc expression and activity may be a
novel and effective approach for the treatment of age-related
CVD.
72 C
The Authors Journal compilation C
2015 Biochemical Society
Molecular pathways of arterial aging
INFLAMMATION
Inflammatory processes are known to contribute to age-related
vascular disease, myocardial infarction, stroke and heart failure
[32]. The molecular events underpinning age-dependent inflam-
mation are not completely understood. Inflammatory alterations
in aging include induction of cellular adhesion molecules, an in-
crease in endothelial–leucocyte interactions and alterations in the
secretion of autocrine/paracrine mediators, which are pivotal to
inflammatory responses.
ROLE AND MOLECULAR TARGETS OF
SIRTUINS IN THE AGED VASCULATURE
The family of nicotinamide adenine dinucleotide-dependent
proteins termed sirtuins has recently emerged as an import-
ant regulator of arterial aging. SIRT1 is considered a ma-
jor gatekeeper against oxidative stress and tissue inflam-
mation [68]. Increased SIRT1 activity confers resistance to
many of the CV sequelae associated with aging [69]. Ac-
tivation of SIRT1 in endothelial tissues may be of bene-
fit in preserving endothelial cell function during aging. Mice
with endothelial specific SIRT1 overexpression on an apoli-
poprotein E (ApoE)/genetic background exhibit attenuated
development of atherosclerotic lesions [70]. In contrast, SIRT1
insufficiency results in greater foam cell formation and ather-
osclerosis [71]. In the human endothelium, overexpression of
SIRT1 prevents oxidative stress-induced senescence, whereas its
inhibition leads to a premature senescence-like phenotype. In-
terestingly, immunosuppressant drugs, like sirolimus and ever-
olimus, induce endothelial cellular senescence via SIRT1 down-
regulation [72]. SIRT1 inhibition also impairs eNOS functional-
ity, whereas its activation improves endothelial NO availability
[73]. Hence, the SIRT1/NO axis may represent a relevant tar-
get against vascular senescence [74,75]. Interestingly enough,
hyperglycaemia, obesity and hypertension are able to reproduce
age-related effects on SIRT1 expression, thus strengthening a
common role of SIRT1 in CVD and early aging.
SIRT1/p53/p66Shc axis
SIRT1-mediated deacetylation modulates the function of pro-
teins via transcriptional and post-translational changes. A previ-
ous study reported that vascular p66Shc gene transcription may
be the result of decreased promoter deacetylation due to down-
regulation of SIRT1 [76]. Expression of p66Shc gene transcript
and protein was significantly increased by different kinds of class
III histone deacetylase inhibitors in human endothelial cells. Con-
sistently, SIRT1 overexpression inhibited high glucose-induced
p66Shc up-regulation, whereas SIRT1 knockdown exerted op-
posite effects. Moreover, endothelium-specific SIRT1 transgenic
mice had blunted p66Shc gene and protein expression and im-
proved endothelial function, as well as reduced accumulation of
oxidative stress markers, compared with WT littermates [76].
The study by Zhou et al. [76] demonstrated that SIRT1 binds to
the p66Shc promoter (508 to 250 bp) where it deacetylates
histone 3, thus suppressing gene transcription of the mitochon-
drial adaptor. Decreased SIRT1-dependent deacetylation is the
main epigenetic mark responsible for p66Shc overexpression in
the vasculature (Figure 1). Therefore, one can certainly pos-
tulate that SIRT1 and p66Shc stand along the same molecular
pathway involved in the modulation of vascular and myocardial
integrity during aging (Figure 1). Moreover, SIRT1 and p66Shc
have molecular circuits with the tumour suppressor p53, crit-
ically involved in age-dependent mitochondrial disruption and
apoptosis (Figure 1) [74]. Indeed, SIRT1 inhibition increases p53
acetylation and transcriptional activity [77]. In this regard, p53
is a master regulator of p66Shc transcription [78]. Accordingly,
down-regulation of p66Shc expression as well as inhibition of p53
function in mice restored impairment of acetylcholine-induced
vascular relaxations and increased NO bioavailability [78]. Taken
together, these observations strongly suggest that p53 is a crit-
ical intermediate between the upstream regulator SIRT1 and its
downstream target p66Shc (Figure 1).
SIRT-1 dependent regulation of energy balance
Age-dependent down-regulation of SIRT1 also favours acetyla-
tion of nuclear factor κB(NF-κB) p65, leading to its nuclear
translocation and transcription of inflammatory genes [71]. The
maintenance of sirtuin homoeostasis (SIRT1 and 3) is funda-
mental to repress detrimental pathways of arterial aging such
as forkhead box O (FOXO) [75]. Inactivation of sirtuins during
aging favours FOXO acetylation and subsequent transcription of
FOXO-dependent genes favouring cellular apoptosis, cell cycle
arrest and accumulation of ROS as well as metabolic derange-
ments [79]. FOXOs are key downstream effectors of the phos-
phoinositide 3-kinase (PI3K)/Akt pathway that blocks FOXO
target gene expression. Following stimulation of PI3K/Akt sig-
nalling by growth factors, Akt phosphorylates FOXOs on three
conserved residues, which leads to their cytoplasmic sequestra-
tion and inactivation [75]. By contrast, Akt is inactive in the
aged vasculature thus promoting FOXO-dependent transcrip-
tional programmes, culminating with endothelial senescent phen-
otypes (Figure 1). Genetic studies in experimental animal mod-
els have clearly shown that many lifespan-extending mutations
affect metabolic regulators and control circuits [80,81]. In this
regard, particular attention deserves the regulation of LKB/AMP-
activated protein kinase (AMPK) pathway by SIRT1 [82]. In the
cytoplasm, SIRT1 deacetylates LKB1 leading to activation of the
final effector enzyme AMPK, a central energy regulator involved
in glucose homoeostasis, maintenance of cellular ATP levels and
endothelial integrity via regulation of eNOS activity [74,75].
Perturbation of SIRT1/LKB/AMPK pathway leads to energy im-
balance, cellular stress and activation of the apoptotic machinery
thus contributing to arterial aging (Figure 1) [75,83].
Nuclear factor κB
Increasing evidence suggests that activationof the redox-sensitive
transcription factor NF-κB plays a key role in endothelial activ-
ation and vascular inflammatory changes in aging [31]. NF-κB
is an important transcription factor expressed in all mammalian
cell types [84]. It regulates expression of genes controlling cell
C
The Authors Journal compilation C
2015 Biochemical Society 73
F. Paneni, S. Costantino and F. Cosentino
Figure 2 Mechanisms underlying age-dependent impairment of vascular healing
spred1, Sprouty-related, EVH1 domain-containing protein 1.
adhesion, proliferation, inflammation and redox state. Activa-
tion of NF-κB mediates vascular and myocardial inflammation
in metabolic and age-related diseases [84]. A previous study
has clearly demonstrated that endothelial suppression of NF-κB
prolongs lifespan in mice and ameliorates obesity-induced en-
dothelial insulin resistance [85]. Impaired insulin signalling is
indeed an important hallmark linking metabolic disease with
premature CV aging [86]. The relevance of these findings is
supported by the notion that NF-κB protein is up-regulated in
vascular endothelial cells isolated from obese and aged adults as
compared with normoweight and young controls [87]. Moreover,
age-dependent NF-κB activation is associated with systemic in-
flammation and impaired endothelial function [88]. Taken to-
gether, these data validate SIRT1 as a key orchestrator of arterial
aging via modulation of oxidative stress, energy balance and vas-
cular inflammation (Figure 1).
ECTOPIC VASCULAR CALCIFICATION
Vascular calcification represents an inevitable hallmark feature
of vascular aging which favours atherothrombosis [89]. Global
measures of coronary artery calcification were shown to inde-
pendently predict both CV events and mortality [90,91]. Current
investigations are now focusing on the possibility that ectopic cal-
cification is mediated by cellular elements acquiring osteogenic
phenotypes. Previous work clearly demonstrated that circulating
osteoblastic cells isolated from human peripheral blood are able
to calcify in vitro and in vivo [92]. These cells, which express
the bone protein osteocalcin (OC) and bone alkaline phosphatase
(BAP), have been considered circulating osteoprogenitor cells
and might participate to vascular calcification and atherosclero-
sis. Indeed, preliminary clinical studies found that coronary ath-
erosclerosis and arterial stiffening are associated with activation
of an osteogenic programme in bone-marrow-derived cells [93].
A study has demonstrated that OC+/BAP+cells originate from
the myeloid lineage and retain monocyte/machrophage markers
[94]. These cells, described as ‘myeloid calcifying cells’ (MCCs),
can be differentiated from pheripheral blood mononuclear cells
and ectopic calcifications in vivo. Interestingly, MCCs are sig-
nificantly increased in patients with Type 2 diabetes and athero-
sclerotic lesions [94], thus providing an important link between
CV risk factors and accelerated arterial aging. Among putative
pathways involved, activation of the transcription factor Runx2
seems to play a major role in osteogenic differentiation (Figure 1).
In conditions of oxidative stress, Akt activates Runx2 which in
turn binds to the promoter of receptor activator of NF-κB ligand
(RANKL), thus favouring changes of VSMC phenotype as well
as differentiation of machrophages into osteoclast-like cells [95].
The Runx2/RANKL pathway is particularly active in senescent
VSMCs, suggesting the importance of this signalling in age-
related calcification and stiffening [96]. In summary, circulating
(MCCs) and resident (VSMCs and machropahges) cells may sig-
nificantly contribute to accelerate ectopic vascular calcification
in aging (Figure 1).
AGE-DEPENDENT DECLINE OF VASCULAR
HEALING PROCESS: CELLULAR AND
MOLECULAR TARGETS
Recovery after ischaemia or infarction in any organrequires blood
vessel growth [97]. The incidence of stroke, claudication and
myocardial infarction all increase in older patients, and they have
worse outcomes when ischaemia and infarction occurs [6]. Un-
derstanding the mechanisms involved in vascular repair is an
important challenge to reduce CVD in aged individuals.
Endothelial progenitor and angiogenic outgrowth
cells
Endothelial progenitor cells (EPCs) and early angiogenic out-
growth cells (EOCs) significantly contribute to endothelial re-
pair, a phenomenon which is less efficient in aging [98]. EPCs
are thought to directly mediate endothelial regeneration, whereas
EOCs represent a heterogeneous pool of cell precursors, mostly
of myeloid origin, which favour endothelial healing via a parac-
rine mechanism [99]. Although we are still far from a clear
understanding of these processes, available knowledge support
74 C
The Authors Journal compilation C
2015 Biochemical Society
Molecular pathways of arterial aging
Figure 3 Unmet scientific needs and future applications
BM, bone marrow; MRI, magnetic resonance imaging.
the notion that aging impairs the function of ex vivo-expanded
EPCs [100]. Age-related EPC dysfunction is mediated by the im-
balance between factors promoting growth, migration/survival
and those enhancing oxidative stress/senescence. Hypoxia indu-
cible factor 1α(HIF-1α) induces the expression of stromal cell-
derived factor 1 (SDF-1) that enhances the recruitment of EPCs
in injured or ischaemic tissues in mice. HIF-1α/SDF-1 signalling
is impaired in aging and contributes to altered vascular repair
(Figure 2) [101]. Interestingly, CV risk factors mirror the aging
process by impairing EPCs functionality and, hence, vascular re-
pair [102,103]. Indeed, the aging gene p66Shc is up-regulated in
EPCs isolated from diabetic subjects and contributes to impaired
migration and tube formation [104]. By contrast, vascular repair
capacities are preserved in EPCs isolated from p66Shc/diabetic
mice [104]. Hence, targeting p66Shc may contribute to rejuvenate
EPCs thus improving cell functionality and angiogenic properties
(Figure 2).
microRNAs
Recent work suggests that microRNAs (miRs) may be involved
in the pathogenesis of age-related EPCs dysfunction (Figure 2)
[105]. These small non-coding RNAs orchestrate EPCs func-
tionality by regulating gene expression at the post-transcriptional
level. In order to map the microRNA/gene expression signa-
tures of EPCs senescence, Zhu et al. [106] performed a mi-
croRNA profiling and microarray analysis in lineage-negative
bone marrow cells from young and aged as well as ApoE/
mice. The analysis found miR-10Aand miR-21, and their com-
mon target gene high-mobility group AT-hook 2 (Hmga2)as
critical regulators of EPCs senescence. Indeed, overexpression of
miR-10Aand miR-21 in young EPCs suppressed Hmga2, leading
to increased p16INK4a/p19ARF expression and impaired EPCs
angiogenesis in vitro and in vivo. In contrast, suppression of
miR-10Aand miR-21 in aged EPCs increased Hmga2 expres-
sion and rejuvenated EPCs, thus improving angiogenesis [106].
Furthermore, miR-34a was found to inhibit EPC-mediated an-
giogenesis by suppressing SIRT1 [107]. Recent studies showed
that reprogramming angiomiR-126 in EOCs and circulating EPCs
restores endothelial homeostasis and favours vascular healing
[99,108]. These novel data suggest the possibility that in vitro
reprogramming of human EPCs may improve vascular repair.
CONCLUSIONS
In the present paper, we have addressed key molecular mech-
anisms implicated in arterial aging and CVD. Premature ac-
tivation of aging genes such as p66Shc and NF-κBaswellas
down-regulation of lifespan determinants JunD and SIRT1 may
trigger senescence features leading to early CVD. Dysregula-
tion of protective and detrimental genes caused by aging and
other CV risk factors suggests the possibility to reprogramme
such modifications in circulating and resident vascular wall cells.
In this perspective, ex vivo reprogramming of autologous EPCs
may rescue age-related cell dysfunction and restore angiogenic
and reparative capacities after myocardial infarction or stroke.
Future experimental and clinical research should address unmet
scientific needs and provide insights for their clinical application
in the context of CV aging (Figure 3).
FUNDING
Our own work was supported by the Swiss Heart Foundation and
the Italian Ministry of Education, University and Research [grant
number PRIN 2010–2011 (to F.C.).] F.P is the recipient of a PhD
programme fellowship in Experimental Medicine at the University
of Rome “Sapienza”.
REFERENCES
1 Kovacic, J. C., Moreno, P., Hachinski, V., Nabel, E. G. and Fuster,
V. (2011) Cellular senescence, vascular disease, and aging:
part 1 of a 2-part review. Circulation. 123, 1650–1660
CrossRef PubMed
2 Katusic, Z. S. and Austin, S. A. (2014) Endothelial nitric oxide:
protector of a healthy mind. Eur. Heart J. 35, 888–894
CrossRef PubMed
3 Afilalo, J., Karunananthan, S., Eisenberg, M. J., Alexander, K. P.
and Bergman, H. (2009) Role of frailty in patients with
cardiovascular disease. Am. J. Cardiol. 103, 1616–1621
CrossRef PubMed
4 Phan, H. M., Alpert, J. S. and Fain, M. (2008) Frailty,
inflammation, and cardiovascular disease: evidence of a
connection. Am. J. Geriatr. Cardiol. 17, 101–107 PubMed
5 Afilalo, J., Alexander, K. P., Mack, M. J., Maurer, M. S., Green, P.,
Allen, L. A., Popma, J. J., Ferrucci, L. and Forman, D. E. (2014)
Frailty assessment in the cardiovascular care of older adults. J.
Am. Coll. Cardiol. 63, 747–762 CrossRef PubMed
C
The Authors Journal compilation C
2015 Biochemical Society 75
F. Paneni, S. Costantino and F. Cosentino
6 Kovacic, J. C., Moreno, P., Nabel, E. G., Hachinski, V. and Fuster,
V. (2011) Cellular senescence, vascular disease, and aging:
part 2 of a 2-part review: clinical vascular disease in the elderly.
Circulation 123, 1900–1910 CrossRef PubMed
7 Laurent, S. (2012) Defining vascular aging and cardiovascular
risk. J. Hypertens. 30 (Suppl.), S3–S8 CrossRef PubMed
8 Laurent, S., Alivon, M., Beaussier, H. and Boutouyrie, P. (2012)
Aortic stiffness as a tissue biomarker for predicting future
cardiovascular events in asymptomatic hypertensive subjects.
Ann. Med. 44 (Suppl. 1), S93–S97 CrossRef PubMed
9 Najjar, S. S., Scuteri, A. and Lakatta, E. G. (2005) Ar terial
aging: is it an immutable cardiovascular risk factor?
Hypertension 46, 454–462 CrossRef PubMed
10 Virdis, A., Ghiadoni, L., Giannarelli, C. and Taddei, S. (2010)
Endothelial dysfunction and vascular disease in later life.
Maturitas 67, 20–24 CrossRef PubMed
11 Cosentino, F., Francia, P., Camici, G. G., Pelicci, P. G., Luscher,
T. F. and Volpe, M. (2008) Final common molecular pathways of
aging and cardiovascular disease: role of the p66Shc protein.
Arterioscler. Thromb. Vasc. Biol. 28, 622–628
CrossRef PubMed
12 Cosentino, F. and Luscher, T. F. (1999) Tetrahydrobiopterin and
endothelial nitric oxide synthase activity. Cardiovasc. Res. 43,
274–278 CrossRef PubMed
13 Brunner, H., Cockcroft, J. R., Deanfield, J., Donald, A.,
Ferrannini, E., Halcox, J., Kiowski, W., Luscher, T. F., Mancia, G.,
Natali, A. et al. (2005) Endothelial function and dysfunction.
Part II: Association with cardiovascular risk factors and
diseases. A statement by the Working Group on Endothelins
and Endothelial Factors of the European Society of
Hypertension. J. Hypertens. 23, 233–246 CrossRef PubMed
14 Alp, N. J. and Channon, K. M. (2004) Regulation of endothelial
nitric oxide synthase by tetrahydrobiopterin in vascular disease.
Arterioscler. Thromb. Vasc. Biol. 24, 413–420
CrossRef PubMed
15 Forstermann, U. and Sessa, W. C. (2012) Nitric oxide
synthases: regulation and function. Eur. Heart J. 33, 829–837
CrossRef PubMed
16 Higashi, Y., Sasaki, S., Nakagawa, K., Kimura, M., Noma, K.,
Hara, K., Jitsuiki, D., Goto, C., Oshima, T., Chayama, K. and
Yoshizumi, M. (2006) Tetrahydrobiopterin improves
aging-related impairment of endothelium-dependent
vasodilation through increase in nitric oxide production.
Atherosclerosis 186, 390–395 CrossRef PubMed
17 Pierce, G. L. and Larocca, T. J. (2008) Reduced vascular
tetrahydrobiopterin (BH4) and endothelial function with ageing:
is it time for a chronic BH4 supplementation trial in
middle-aged and older adults? J. Physiol. 586, 2673–2674
CrossRef PubMed
18 Pernow, J. and Jung, C. (2013) Arginase as a potential target in
the treatment of cardiovascular disease: reversal of arginine
steal? Cardiovasc. Res. 98, 334–343 CrossRef PubMed
19 Ming, X. F., Rajapakse, A. G., Yepuri, G., Xiong, Y., Carvas, J.
M., Ruffieux, J., Scerri, I., Wu, Z., Popp, K., Li, J., Sartori, C.
et al. (2012) Arginase II promotes macrophage inflammatory
responses through mitochondrial reactive oxygen species,
contributing to insulin resistance and atherogenesis. J. Am.
Heart Assoc. 1, e000992 CrossRef PubMed
20 Yepuri, G., Velagapudi, S., Xiong, Y., Rajapakse, A. G., Montani,
J. P., Ming , X. F. and Yang, Z. (2012) Positive crosstalk between
arginase-II and S6K1 in vascular endothelial inflammation and
aging. Aging Cell. 11, 1005–1016 CrossRef PubMed
21 Shemyakin, A., Kovamees, O., Rafnsson, A., Bohm, F.,
Svenarud, P., Settergren, M., Jung, C. and Pernow, J. (2012)
Arginase inhibition improves endothelial function in patients
with coronary artery disease and type 2 diabetes mellitus.
Circulation 126, 2943–2950 CrossRef PubMed
22 van der Loo, B., Labugger, R., Skepper, J. N., Bachschmid, M.,
Kilo, J., Powell, J. M., Palacios-Callender, M., Erusalimsky, J. D.,
Quaschning, T., Malinski, T. et al. (2000) Enhanced peroxynitrite
formation is associated with vascular aging. J. Exp. Med. 192,
1731–1744 CrossRef PubMed
23 Donato, A. J., Gano, L. B., Eskurza, I., Silver, A. E., Gates, P.E.,
Jablonski, K. and Seals, D. R. (2009) Vascular endothelial
dysfunction with aging: endothelin-1 and endothelial nitric oxide
synthase. Am. J. Physiol. Heart Circ. Physiol. 297, H425–H432
PubMed
24 Donato, A. J., Magerko, K. A., Lawson, B. R., Durrant, J. R.,
Lesniewski, L. A. and Seals, D. R. (2011) SIRT-1 and vascular
endothelial dysfunction with ageing in mice and humans. J.
Physiol. 589, 4545–4554 PubMed
25 Camici, P. G. and Crea, F. (2007) Coronary microvascular
dysfunction. N. Engl. J. Med. 356, 830–840 CrossRef PubMed
26 O’Rourke, M. F., Safar, M. E. and Dzau, V. (2010) The
Cardiovascular Continuum extended: aging effects on the aorta
and microvasculature. Vasc. Med. 15, 461–468
CrossRef PubMed
27 Kurotobi, T., Sato, H., Kinjo, K., Nakatani, D., Mizuno, H.,
Shimizu, M., Imai, K., Hirayama, A., Kodama, K., Hori, M. and
Group, O. (2004) Reduced collateral circulation to the
infarct-related artery in elderly patients with acute myocardial
infarction. J. Am. Coll. Cardiol. 44, 28–34 CrossRef PubMed
28 Kluge, M. A., Fetterman, J. L. and Vita, J. A. (2013)
Mitochondria and endothelial function. Circ. Res. 112,
1171–1188 CrossRef PubMed
29 Paneni, F., Beckman, J. A., Creager, M. A. and Cosentino, F.
(2013) Diabetes and vascular disease: pathophysiology, clinical
consequences, and medical therapy: part I. Eur. Heart J. 34,
2436–2443 CrossRef PubMed
30 Cosentino, F., Eto, M., De Paolis, P., van der Loo, B.,
Bachschmid, M., Ullrich, V., Kouroedov, A., Delli Gatti, C., Joch,
H., Volpe, M. and Luscher, T. F. (2003) High glucose causes
upregulation of cyclooxygenase-2 and alters prostanoid profile
in human endothelial cells: role of protein kinase C and reactive
oxygen species. Circulation 107, 1017–1023
CrossRef PubMed
31 Ungvari, Z., Parrado-Fernandez, C., Csiszar, A. and de Cabo, R.
(2008) Mechanisms underlying caloric restriction and lifespan
regulation: implications for vascular aging. Circ. Res. 102,
519–528 CrossRef PubMed
32 El Assar, M., Angulo, J. and Rodriguez-Manas, L. (2013)
Oxidative stress and vascular inflammation in aging. Free
Radic. Biol. Med. 65, 380–401 CrossRef PubMed
33 Dai, D. F., Rabinovitch, P. S. and Ungvari, Z. (2012)
Mitochondria and cardiovascular aging. Circ. Res. 110,
1109–1124 CrossRef PubMed
34 Shenouda, S. M., Widlansky, M. E., Chen, K., Xu, G., Holbrook,
M., Tabit, C. E., Hamburg, N. M., Frame, A. A., Caiano, T. L.,
Kluge, M. A. et al. (2011) Altered mitochondrial dynamics
contributes to endothelial dysfunction in diabetes mellitus.
Circulation 124, 444–453 CrossRef PubMed
35 Czypiorski, P., Rabanter, L. L., Altschmied, J. and Haendeler, J.
(2013) Redox balance in the aged endothelium. Z. Gerontol.
Geriatr. 46, 635–638 CrossRef PubMed
36 Montezano, A. C. and Touyz, R. M. (2014) Reactive oxygen
species, vascular Noxs, and hypertension: focus on
translational and clinical research. Antioxid. Redox Signal. 20,
164–182 CrossRef PubMed
37 Brandes, R. P., Weissmann, N. and Schroder, K. (2010) NADPH
oxidases in cardiovascular disease. Free Radic. Biol. Med. 49,
687–706 CrossRef PubMed
76 C
The Authors Journal compilation C
2015 Biochemical Society
Molecular pathways of arterial aging
38 Lener, B., Koziel, R., Pircher, H., Hutter, E., Greussing, R.,
Herndler-Brandstetter, D., Hermann, M., Unterluggauer, H. and
Jansen-Durr, P. (2009) The NADPH oxidase Nox4 restricts the
replicative lifespan of human endothelial cells. Biochem. J.
423, 363–374 CrossRef PubMed
39 Paneni, F., Osto, E., Costantino, S., Mateescu, B., Briand, S.,
Coppolino, G., Perna, E., Mocharla, P., Akhmedov, A., Kubant, R.
et al. (2013) Deletion of the activated protein-1 transcription
factor JunD induces oxidative stress and accelerates
age-related endothelial dysfunction. Circulation 127,
1229–1240 CrossRef PubMed
40 Meyer, M. R., Fredette, N. C., Barton, M. and Prossnitz, E. R.
(2014) Endothelin-1 but not angiotensin II contributes to
functional aging in murine carotid arteries. Life Sci.,
doi:10.1016/j.lfs.2014.02.027 CrossRef PubMed
41 Park, L., Anrather, J., Girouard, H., Zhou, P. and Iadecola, C.
(2007) Nox2-derived reactive oxygen species mediate
neurovascular dysregulation in the aging mouse brain. J. Cereb.
Blood Flow Metab. 27, 1908–1918 CrossRef PubMed
42 Bachschmid, M., van der Loo, B., Schuler, K., Labugger, R.,
Thurau, S., Eto, M., Kilo, J., Holz, R., Luscher, T. F. and Ullrich,
V. (2004) Oxidative stress-associated vascular aging is
independent of the protein kinase C/NAD(P)H oxidase
pathway. Arch. Gerontol. Geriatr. 38, 181–190
CrossRef PubMed
43 Gerald, D., Berra, E., Frapart, Y. M., Chan, D. A., Giaccia, A. J.,
Mansuy, D., Pouyssegur, J., Yaniv, M. and Mechta-Grigoriou, F.
(2004) JunD reduces tumor angiogenesis by protecting cells
from oxidative stress. Cell 118, 781–794 CrossRef PubMed
44 Mechta-Grigoriou, F., Gerald, D. and Yaniv, M. (2001) The
mammalian Jun proteins: redundancy and specificity. Oncogene
20, 2378–2389 CrossRef PubMed
45 Hernandez, J. M., Floyd, D. H., Weilbaecher, K. N., Green, P.L.
and Boris-Lawrie, K. (2008) Multiple facets of junD gene
expression are atypical among AP-1 family members. Oncogene
27, 4757–4767 CrossRef PubMed
46 Paneni, F., Costantino, S., Volpe, M., Luscher, T. F. and
Cosentino, F. (2013) Epigenetic signatures and vascular risk in
type 2 diabetes: A clinical perspective. Atherosclerosis 230,
191–197 CrossRef PubMed
47 Handy, D. E., Castro, R. and Loscalzo, J. (2011) Epigenetic
modifications: basic mechanisms and role in cardiovascular
disease. Circulation 123, 2145–2156 CrossRef PubMed
48 Chen, C. H., Budas, G. R., Churchill, E. N., Disatnik, M. H.,
Hurley, T. D. and Mochly-Rosen, D. (2008) Activation of
aldehyde dehydrogenase-2 reduces ischemic damage to the
heart. Science 321, 1493–1495 CrossRef PubMed
49 Koda, K., Salazar-Rodriguez, M., Corti, F., Chan, N. Y.,
Estephan, R., Silver, R. B., Mochly-Rosen, D. and Levi, R.
(2010) Aldehyde dehydrogenase activation prevents reperfusion
arrhythmias by inhibiting local renin release from cardiac mast
cells. Circulation 122, 771–781 CrossRef PubMed
50 Srinivasan, S., Hatley, M. E., Bolick, D. T., Palmer, L. A.,
Edelstein, D., Brownlee, M. and Hedrick, C. C. (2004)
Hyperglycaemia-induced superoxide production decreases
eNOS expression via AP-1 activation in aortic endothelial cells.
Diabetologia 47, 1727–1734 CrossRef PubMed
51 Ricci, R., Eriksson, U., Oudit, G. Y., Eferl, R., Akhmedov, A.,
Sumara, I., Sumara, G., Kassiri, Z., David, J. P., Bakiri, L. et al.
(2005) Distinct functions of junD in cardiac hypertrophy and
heart failure. Genes Dev. 19, 208–213 CrossRef PubMed
52 Paneni, F. and Cosentino, F. (2012) p66 Shc as the engine of
vascular aging. Curr. Vasc. Pharmacol. 10, 697–699
CrossRef PubMed
53 Giorgio, M., Migliaccio, E., Orsini, F., Paolucci, D., Moroni, M.,
Contursi, C., Pelliccia, G., Luzi, L., Minucci, S., Marcaccio, M.
et al. (2005) Electron transfer between cytochrome c and
p66Shc generates reactive oxygen species that trigger
mitochondrial apoptosis. Cell 122, 221–233
CrossRef PubMed
54 Migliaccio, E., Giorgio, M., Mele, S., Pelicci, G., Reboldi, P.,
Pandolfi, P.P., Lanfrancone, L. and Pelicci, P. G. (1999) The
p66shc adaptor protein controls oxidative stress response and
life span in mammals. Nature 402, 309–313 CrossRef PubMed
55 Francia, P., delli Gatti, C., Bachschmid, M., Mar tin-Padura, I.,
Savoia, C., Migliaccio, E., Pelicci, P. G., Schiavoni, M., Luscher,
T. F., Volpe, M. and Cosentino, F. (2004) Deletion
of p66shc gene protects against age-related endothelial
dysfunction. Circulation 110, 2889–2895
CrossRef PubMed
56 Spescha, R. D., Shi, Y., Wegener, S., Keller, S., Weber, B.,
Wyss, M. M., Lauinger, N., Tabatabai, G., Paneni, F., Cosentino,
F. et al. (2013) Deletion of the ageing gene p66(Shc) reduces
early stroke size following ischaemia/reperfusion brain injury.
Eur. Heart J. 34, 96–103 CrossRef PubMed
57 Pinton, P., Rimessi, A., Marchi, S., Or sini, F., Migliaccio, E.,
Giorgio, M., Contursi, C., Minucci, S., Mantovani, F.,
Wieckowski, M. R. et al. (2007) Protein kinase C beta and prolyl
isomerase 1 regulate mitochondrial effects of the life-span
determinant p66Shc. Science 315, 659–663 CrossRef PubMed
58 Tomilov, A. A., Bicocca, V., Schoenfeld, R. A., Giorgio, M.,
Migliaccio, E., Ramsey, J. J., Hagopian, K., Pelicci, P.G.and
Cortopassi, G. A. (2010) Decreased superoxide
production in macrophages of long-lived p66Shc knock-out
mice. J. Biol. Chem. 285, 1153–1165
CrossRef PubMed
59 Rota, M., LeCapitaine, N., Hosoda, T., Boni, A., De Angelis, A.,
Padin-Iruegas, M. E., Esposito, G., Vitale, S., Urbanek, K. et al.
(2006) Diabetes promotes cardiac stem cell aging and heart
failure, which are prevented by deletion of the p66shc gene.
Circ. Res. 99, 42–52 CrossRef PubMed
60 Paneni, F., Mocharla, P., Akhmedov, A., Costantino, S., Osto, E.,
Volpe, M., Luscher, T. F. and Cosentino, F. (2012) Gene
silencing of the mitochondrial adaptor p66(Shc) suppresses
vascular hyperglycemic memory in diabetes. Circ. Res. 111,
278–289 CrossRef PubMed
61 Carpi, A., Menabo, R., Kaludercic, N., Pelicci, P., Di Lisa, F. and
Giorgio, M. (2009) The cardioprotective effects elicited by
p66(Shc) ablation demonstrate the crucial role of
mitochondrial ROS formation in ischemia/reperfusion injury.
Biochim. Biophys. Acta. 1787, 774–780
CrossRef PubMed
62 Shi, Y., Cosentino, F., Camici, G. G., Akhmedov, A., Vanhoutte,
P. M., Tanner, F. C. and Luscher, T. F. (2011) Oxidized low-density
lipoprotein activates p66Shc via lectin-like oxidized low-density
lipoprotein receptor-1, protein kinase C-beta, and c-Jun
N-terminal kinase kinase in human endothelial cells.
Arterioscler. Thromb. Vasc. Biol. 31, 2090–2097
CrossRef PubMed
63 Spescha, R. D., Glanzmann, M., Simic, B., Witassek, F., Keller,
S., Akhmedov, A., Tanner, F. C., Luscher, T. F. and Camici, G. G.
(2014) Adaptor protein p66Shc mediates hypertension-
associated, cyclic stretch-dependent, endothelial damage.
Hypertension 64, 347–353 CrossRef PubMed
64 Pagnin, E., Fadini, G., de Toni, R., Tiengo, A., Calo, L. and
Avogaro, A. (2005) Diabetes induces p66shc gene expression
in human peripheral blood mononuclear cells: relationship to
oxidative stress. J. Clin. Endocrinol. Metab. 90, 1130–1136
CrossRef PubMed
65 Franzeck, F. C., Hof, D., Spescha, R. D., Hasun, M., Akhmedov,
A., Steffel, J., Shi, Y., Cosentino, F., Tanner, F. C., von
Eckardstein, A. et al. (2012) Expression of the aging gene
p66Shc is increased in peripheral blood monocytes of patients
with acute coronary syndrome but not with stable coronary
artery disease. Atherosclerosis 220, 282–286
CrossRef PubMed
66 Pandolfi, S., Bonafe, M., Di Tella, L., Tiberi, L., Salvioli, S.,
Monti, D., Sorbi, S. and Franceschi, C. (2005) p66(shc) is
highly expressed in fibroblasts from centenarians. Mech.
Ageing Dev. 126, 839–844 CrossRef PubMed
C
The Authors Journal compilation C
2015 Biochemical Society 77
F. Paneni, S. Costantino and F. Cosentino
67 Miao, Q., Wang, Q., Dong, L., Wang, Y., Tan, Y. and Zhang , X.
(2014) The expression of p66shc in peripheral blood
monocytes is increased in patients with coronary heart disease
and correlated with endothelium-dependent vasodilatation.
Heart Vessels. doi:10.1007/s00380-014-0497-4
68 Corbi, G., Conti, V., Scapagnini, G., F ilippelli, A. and Ferrara, N.
(2012) Role of sirtuins, calorie restriction and physical activity
in aging. Front. Biosci. 4, 768–778 CrossRef
69 Hall, J. A., Dominy, J. E., Lee, Y. and Puigser ver, P. (2013) The
sirtuin family’s role in aging and age-associated pathologies. J.
Clin. Invest. 123, 973–979 CrossRef PubMed
70 Zhang, Q. J., Wang, Z., Chen, H. Z., Zhou, S., Zheng, W., Liu,
G., Wei, Y. S., Cai, H., Liu, D. P. and Liang, C. C. (2008)
Endothelium-specific overexpression of class III deacetylase
SIRT1 decreases atherosclerosis in apolipoprotein E-deficient
mice. Cardiovasc. Res. 80, 191–199 CrossRef PubMed
71 Stein, S., Lohmann, C., Schafer, N., Hofmann, J., Rohrer, L.,
Besler, C., Rothgiesser, K. M., Becher, B., Hottiger, M. O.,
Boren, J. et al. (2010) SIRT1 decreases Lox-1-mediated foam
cell formation in atherogenesis. Eur. Heart J. 31, 2301–2309
CrossRef PubMed
72 Ota, H., Eto, M., Ako, J., Ogawa, S., Iijima, K., Akishita, M. and
Ouchi, Y. (2009) Sirolimus and everolimus induce endothelial
cellular senescence via sirtuin 1 down-regulation: therapeutic
implication of cilostazol after drug-eluting stent implantation. J.
Am. Coll. Cardiol. 53, 2298–2305 CrossRef PubMed
73 Mattagajasingh, I., Kim, C. S., Naqvi, A., Yamamori, T.,
Hoffman, T. A., Jung, S. B., DeRicco, J., Kasuno, K. and Irani, K.
(2007) SIRT1 promotes endothelium-dependent vascular
relaxation by activating endothelial nitric oxide synthase. Proc.
Natl. Acad. Sci. U.S.A. 104, 14855–14860 CrossRef PubMed
74 Paneni, F., Volpe, M., Luscher, T. F. and Cosentino, F. (2013)
SIRT1, p66Shc, and Set7/9 in vascular hyperglycemic memory:
bringing all the strands together. Diabetes 62, 1800–1807
CrossRef PubMed
75 Oellerich, M. F. and Potente, M. (2012) FOXOs and sirtuins in
vascular growth, maintenance, and aging. Circ. Res. 110,
1238–1251 CrossRef PubMed
76 Zhou, S., Chen, H. Z., Wan, Y. Z., Zhang, Q. J., Wei, Y. S.,
Huang, S., Liu, J. J., Lu, Y. B., Zhang , Z. Q., Yang, R. F. et al.
(2011) Repression of P66Shc expression by SIRT1 contributes
to the prevention of hyperglycemia-induced endothelial
dysfunction. Circ. Res. 109, 639–648 CrossRef PubMed
77 Orimo, M., Minamino, T., Miyauchi, H., Tateno, K., Okada, S.,
Moriya, J. and Komuro, I. (2009) Protective role of SIRT1 in
diabetic vascular dysfunction. Arterioscler. Thromb. Vasc. Biol.
29, 889–894 CrossRef PubMed
78 Kim, C. S., Jung, S. B., Naqvi, A., Hoffman, T. A., DeRicco, J.,
Yamamori, T., Cole, M. P., Jeon, B. H. and Irani, K. (2008) p53
impairs endothelium-dependent vasomotor function through
transcriptional upregulation of p66shc. Circ. Res. 103,
1441–1450 CrossRef PubMed
79 Calnan, D. R. and Brunet, A. (2008) The FoxO code. Oncogene
27, 2276–2288 CrossRef PubMed
80 Bishop, N. A. and Guarente, L. (2007) Genetic links between
diet and lifespan: shared mechanisms from yeast to humans.
Nat. Rev. Genet. 8, 835–844 CrossRef PubMed
81 Houtkooper, R. H., Williams, R. W. and Auwerx, J. (2010)
Metabolic networks of longevity. Cell 142, 9–14
CrossRef PubMed
82 Li, C. and Keaney, Jr, J. F. (2010) AMP-activated protein kinase:
a stress-responsive kinase with implications for cardiovascular
disease. Curr. Opin. Pharmacol. 10, 111–115
CrossRef PubMed
83 Nagata, D. and Hirata, Y. (2010) The role of AMP-activated
protein kinase in the cardiovascular system. Hypertens. Res.
33, 22–28 CrossRef PubMed
84 Baker, R. G., Hayden, M. S. and Ghosh, S. (2011) NF-kappaB,
inflammation, and metabolic disease. Cell Metab. 13, 11–22
CrossRef PubMed
85 Hasegawa, Y., Saito, T., Ogihara, T., Ishigaki, Y., Yamada, T.,
Imai, J., Uno, K., Gao, J., Kaneko, K., Shimosawa, T. et al.
(2012) Blockade of the nuclear factor-kappaB pathway in the
endothelium prevents insulin resistance and prolongs life
spans. Circulation 125, 1122–1133
CrossRef PubMed
86 Avogaro, A., de Kreutzenberg, S. V., Federici, M. and Fadini, G. P.
(2013) The endothelium abridges insulin resistance to
premature aging. J. Am. Heart Assoc. 2, e000262
CrossRef PubMed
87 Donato, A. J., Pierce, G. L., Lesniewski, L. A. and Seals, D. R.
(2009) Role of NFkappaB in age-related vascular endothelial
dysfunction in humans. Aging 1, 678–680 PubMed
88 Tabit, C. E., Shenouda, S. M., Holbrook, M., Fetterman, J. L.,
Kiani, S., Frame, A. A., Kluge, M. A., Held, A., Dohadwala,
M. M., Gokce, N. et al. (2013) Protein kinase C-beta contributes
to impaired endothelial insulin signaling in humans with
diabetes mellitus. Circulation 127, 86–95 CrossRef PubMed
89 Kovacic, J. C. and Randolph, G. J. (2011) Vascular calcification:
harder than it looks. Arterioscler. Thromb. Vasc. Biol. 31,
1249–1250 CrossRef PubMed
90 Budoff, M. J., Shaw, L. J., Liu, S. T., Weinstein, S. R., Mosler,
T. P., Tseng, P. H., Flores, F. R., Callister, T. Q., Raggi, P.and
Berman, D. S. (2007) Long-term prognosis associated with
coronary calcification: observations from a registry of 25,253
patients. J. Am. Coll. Cardiol. 49, 1860–1870
CrossRef PubMed
91 Arad, Y., Goodman, K. J., Roth, M., Newstein, D. and Guerci,
A. D. (2005) Coronary calcification, coronary disease risk
factors, C-reactive protein, and atherosclerotic cardiovascular
disease events: the St. Francis Heart Study. J. Am. Coll.
Cardiol. 46, 158–165 CrossRef PubMed
92 Eghbali-Fatourechi, G. Z., Lamsam, J., Fraser, D., Nagel, D.,
Riggs, B. L. and Khosla, S. (2005) Circulating
osteoblast-lineage cells in humans. N. Engl. J. Med. 352,
1959–1966 CrossRef PubMed
93 Flammer, A. J., Gossl, M., Widmer, R. J., Reriani, M., Lennon,
R., Loeffler, D., Shonyo, S., Simari, R. D., Lerman, L. O., Khosla,
S. and Lerman, A. (2012) Osteocalcin positive
CD133+/CD34/KDR+progenitor cells as an independent
marker for unstable atherosclerosis. Eur. Heart J. 33,
2963–2969 CrossRef PubMed
94 Fadini, G. P., Albiero, M., Menegazzo, L., Boscaro, E., Vigili de
Kreutzenberg, S., Agostini, C., Cabrelle, A., Binotto, G.,
Rattazzi, M., Bertacco, E. et al. (2011) Widespread increase in
myeloid calcifying cells contributes to ectopic vascular
calcification in type 2 diabetes. Circ. Res. 108, 1112–1121
CrossRef PubMed
95 Byon, C. H., Sun, Y., Chen, J., Yuan, K., Mao, X., Heath, J. M.,
Anderson, P. G., Tintut, Y., Demer, L. L., Wang, D. and Chen, Y.
(2011) Runx2-upregulated receptor activator of nuclear factor
kappaB ligand in calcifying smooth muscle cells promotes
migration and osteoclastic differentiation of macrophages.
Arterioscler. Thromb. Vasc. Biol. 31, 1387–1396
CrossRef PubMed
96 Nakano-Kurimoto, R., Ikeda, K., Uraoka, M., Nakagawa, Y.,
Yutaka, K., Koide, M., Takahashi, T., Matoba, S., Yamada, H.,
Okigaki, M. and Matsubara, H. (2009) Replicative senescence
of vascular smooth muscle cells enhances the calcification
through initiating the osteoblastic transition. Am. J. Physiol.
Heart Circ. Physiol. 297, H1673–H1684 PubMed
97 Lahteenvuo, J. and Rosenzweig, A. (2012) Effects of aging on
angiogenesis. Circ. Res. 110, 1252–1264 CrossRef PubMed
98 Williamson, K., Stringer, S. E. and Alexander, M. Y. (2012)
Endothelial progenitor cells enter the aging arena. Front.
Physiol. 3,30CrossRef PubMed
78 C
The Authors Journal compilation C
2015 Biochemical Society
Molecular pathways of arterial aging
99 Jakob, P., Doerries, C., Briand, S., Mochar la, P., Krankel, N.,
Besler, C., Mueller, M., Manes, C., Templin, C., Baltes, C. et al.
(2012) Loss of angiomiR-126 and 130a in angiogenic early
outgrowth cells from patients with chronic heart failure: role for
impaired in vivo neovascularization and cardiac repair capacity.
Circulation 126, 2962–2975 CrossRef PubMed
100 Dzau, V. J., Gnecchi, M., Pachori, A. S., Morello, F. and Melo,
L. G. (2005) Therapeutic potential of endothelial progenitor
cells in cardiovascular diseases. Hypertension 46, 7–18
CrossRef PubMed
101 Hoenig, M. R., Bianchi, C. and Sellke, F. W. (2008) Hypoxia
inducible factor-1 alpha, endothelial progenitor cells,
monocytes, cardiovascular risk, wound healing, cobalt and
hydralazine: a unifying hypothesis. Curr. Drug Targets. 9,
422–435 CrossRef PubMed
102 Templin, C., Krankel, N., Luscher, T. F. and Landmesser, U.
(2011) Stem cells in cardiovascular regeneration: from
preservation of endogenous repair to future cardiovascular
therapies. Curr. Pharm. Des. 17, 3280–3294 CrossRef PubMed
103 Giannotti, G., Doerries, C., Mocharla, P. S., Mueller, M. F.,
Bahlmann, F. H., Horvath, T., Jiang, H., Sorrentino, S. A.,
Steenken, N., Manes, C. et al. (2010) Impaired endothelial
repair capacity of early endothelial progenitor cells in
prehypertension: relation to endothelial dysfunction.
Hypertension 55, 1389–1397 CrossRef PubMed
104 Di Stefano, V., Cencioni, C., Zaccagnini, G., Magenta, A.,
Capogrossi, M. C. and Martelli, F. (2009) p66ShcA modulates
oxidative stress and survival of endothelial progenitor cells in
response to high glucose. Cardiovasc. Res. 82, 421–429
CrossRef PubMed
105 Liang, R., Bates, D. J. and Wang, E. (2009) Epigenetic control
of microRNA expression and aging. Curr. Genomics 10,
184–193 CrossRef PubMed
106 Zhu, S., Deng, S., Ma, Q., Zhang, T., Jia, C., Zhuo, D., Yang, F.,
Wei, J., Wang, L., Dykxhoorn, D. M. et al. (2013)
MicroRNA-10Aand MicroRNA-21 modulate endothelial
progenitor cell senescence via suppressing
high-mobility group A2. Circ. Res. 112, 152–164
CrossRef PubMed
107 Zhao, T., Li, J. and Chen, A. F. (2010) MicroRNA-34a induces
endothelial progenitor cell senescence and impedes its
angiogenesis via suppressing silent information regulator 1.
Am. J. Physiol. Endocrinol. Metab. 299, E110–E116
CrossRef PubMed
108 Mocharla, P., Briand, S., Giannotti, G., Dorries, C., Jakob, P.,
Paneni, F., Luscher, T. and Landmesser, U. (2013)
AngiomiR-126 expression and secretion from circulating
CD34(+) and CD14(+) PBMCs: role for proangiogenic effects
and alterations in type 2 diabetics. Blood 121, 226–236
CrossRef PubMed
Received 13 May 2014/27 June 2014; accepted 22 July 2014
Published on the Internet 5 September 2014, doi: 10.1042/CS20140302
C
The Authors Journal compilation C
2015 Biochemical Society 79
... Exercise stimulates NO release through increased blood flow and shear stress (Lin and Lee, 2018). Additionally, a balance between NO and reactive oxidative species (ROS) is crucial for vascular health (Paneni et al., 2015), and PA reduces ROS levels while increasing NO bioavailability (Dorner et al., 2003). Lipid-lowering effects are another putative route for PA-induced advantages to the retinal microvascular system. ...
Article
Introduction: Examining the retina represents a non-invasive method to evaluate abnormalities pertaining to the nervous and cardiovascular systems. Evidence indicates that physical activity is a non-pharmacological inter- vention to enhance the nervous and cardiovascular systems. However, little is unknown about its effects on ocular characteristics in children and adolescents. The purpose of this study was to examine the effects of physical activity interventions on ocular characteristics in children and adolescents. Method: The electronic bases Web of Science, Embase, Cochrane Library, PubMed, SPORTDiscus, CINAHL, and ERIC were searched from inception to May 2023. Incorporated were randomized controlled trials or quasi- experimental designs that had implemented acute or chronic physical activity interventions among children and adolescents to evaluate various eye-related attributes via clinical examinations or surveys. Two authors independently performed the data extraction and risk of bias assessment, utilizing the Physiotherapy Evidence Database checklist. Results: A total of 474 articles were identified, of which eight articles underwent a systematic review, and six were chosen for meta-analysis. Chronic physical activity interventions positively impacted central retinal artery equivalent (CRAE) with a small to moderate effect (SMD = 0.21; 95% CI 0.04 to 0.39, p = 0.034, I2 = 0%) and central retinal venular equivalent (CRVE) with a small effect (SMD = 0.098; 95% CI 0.08 to 0.11; p = 0.008, I2 = 0%). Intraocular pressure, kinetic visual acuity, and eye strain also improved significantly after physical activity interventions. Discussion: Participating in chronic physical activity programs appear to impact children and adolescents’ eye- related attributes positively.
... Age-related structuralfunctional changes of the vessel walls and heart cause tissue and organ dysfunction in elderly, which makes aging main risk factor of cardiovascular diseases. Aging of arteries is characterized by rigidity, decrease of elasticity, disruption of extensibility and endothelium dysfunction [128]. The main ECM proteins of the vessel wall are elastin, collagens, and soluble PGs [129]. ...
Article
Extracellular matrix (ECM) is an extracellular tissue structure that, in addition to mechanical support to the cell, is involved in regulation of many cellular processes, including chemical transport, growth, migration, differentiation, and cell senescence. Age-related changes in the structure and composition of the matrix and increase of ECM stiffness with age affect functioning of many tissues and contribute to the development of various pathological conditions. This review considers age-related changes of ECM in various tissues and organs, in particular, effect of ECM changes on aging is discussed.
... T2D epigenetics helps in determining the destructive effects of hyperglycemia, i.e., hyperglycemic or metabolic memory, despite optimum control. Epigenetic signatures in inflammatory and pro-oxidant gene promoters are involved in atherosclerotic features, endothelial dysfunction, retinopathy, and diabetic nephropathy, despite normoglycemia being restored [69]. In peripheral blood mononuclear cells (PBMCs) from T2D patients after glycemic control, there was reduced methylation of adaptor p66 Shc promoter (mitochondrial oxidative stress), which was found to be related to oxidative stress levels and endothelial dysfunction [70]. ...
Article
Full-text available
Epigenetics is defined as the study of inheritable changes in the gene expressions and phenotypes that occurs without altering the normal DNA sequence. These changes are mainly due to an alteration in chromatin or its packaging, which changes the DNA accessibility. DNA methylation, histone modification, and noncoding or microRNAs can best explain the mechanism of epigenetics. There are various DNA methylated enzymes, histone-modifying enzymes, and microRNAs involved in the cause of various CVDs (cardiovascular diseases) such as cardiac hypertrophy, heart failure, and hypertension. Moreover, various CVD risk factors such as diabetes mellitus, hypoxia, aging, dyslipidemia, and their epigenetics are also discussed together with CVDs such as CHD (coronary heart disease) and PAH (pulmonary arterial hypertension). Furthermore, different techniques involved in epigenetic chromatin mapping are explained. Among these techniques, the ChIP-on-chip guide is explained with regard to its role in cardiac hypertrophy, a final form of heart failure. This review focuses on different epigenetic factors that are involved in causing cardiovascular diseases.
... The structure and function of the arteries are significantly altered during aging, which makes aging a major risk factor for cardiovascular diseases (CVDs) (Heidenreich et al., 2011;North and Sinclair, 2012;Paneni et al., 2017). Arterial aging is characterized by increased stiffness, reduced elasticity, impaired distensibility, endothelial dysfunction, and deregulated vascular tone (Paneni et al., 2015;Sun, 2015). To develop more efficient treatment to slow down arterial aging and prevent age-associated pathologies, it is necessary to comprehensively understand the structural and functional alterations in the aging arteries. ...
Article
Full-text available
The aging population is booming all over the world and arterial aging causes various age-associated pathologies such as cardiovascular diseases (CVDs). The aorta is the largest elastic artery, and transforms pulsatile flow generated by the left ventricle into steady flow to maintain circulation in distal tissues and organs. Age-associated structural and functional changes in the aortic wall such as dilation, tortuousness, stiffening and losing elasticity hamper stable peripheral circulation, lead to tissue and organ dysfunctions in aged people. The extracellular matrix (ECM) is a three-dimensional network of macromolecules produced by resident cells. The composition and organization of key ECM components determine the structure-function relationships of the aorta and therefore maintaining their homeostasis is critical for a healthy performance. Age-associated remodeling of the ECM structural components, including fragmentation of elastic fibers and excessive deposition and crosslinking of collagens, is a hallmark of aging and leads to functional stiffening of the aorta. In this mini review, we discuss age-associated alterations of the ECM in the aortic wall and shed light on how understanding the mechanisms of aortic aging can lead to the development of efficient strategy for aortic pathologies and CVDs.
... In general, Obi et al. [23] suggested that physiological and metabolic responses contributing to the aetiology of venous thrombi may be different in young and old animals. Ageing-related enhanced oxidative stress can result in the inhibition of antioxidant defence mechanisms as well as the inactivation of enzymes and proteins vital for normal vascular function, which translates into deleterious effects on the vascular system [23,53]. Therefore, it appears that oxidative stress related to ageing may also contribute to the metabolic mechanism of venous thrombosis [23]. ...
Article
Full-text available
Venous thromboembolism (VTE) is a condition comprising deep venous thrombosis (DVT) and pulmonary embolism (PE). The prevalence of this disease is constantly increasing and it is also a chief reason for morbidity. Therefore, the primary prevention of VTE remains a highly important public health issue. At present, its diagnosis generally relies on subjective clinical examination and ultrasound imaging. D-dimer is also used as a biomarker, but it is considered to be poorly specific and only moderately sensitive. There are also no reliable methods that could accurately guide the type of treatment and potentially identify patients who may benefit from more aggressive therapies without the risk of bleeding. The application of metabolomics profiling in the area of vascular diseases may become a turning point in early diagnosis and patient management. Among the most described metabolites possibly related to VTE are carnitine species, glucose, phenylalanine, 3-hydroxybutarate, lactic acid, tryptophan and some monounsaturated and polyunsaturated fatty acids. The cell response to acute PE was suggested to involve the uncoupling between glycolysis and oxidative phosphorylation. Despite technological advancement in the identification of metabolites and their alteration in thrombosis, we still do not understand the mechanisms and pathways responsible for the occurrence of observed alterations.
... 81 It also triggers vascular senescence in VSMCs and endothelial cells. 82 Arginase inhibition restores endothelial-dependent vasorelaxation in small arteries of obese patients, although this effect is attenuated in ageing due to a high prooxidant environment sustained by hyperactivation of the NADPH oxidase (Nox). 41,83 The family of Nox enzymes is another major source of microvascular ROS. ...
Article
Full-text available
The development of novel, non-invasive techniques and standardization of protocols to assess microvascular dysfunction have elucidated the key role of microvascular changes in the evolution of cardiovascular (CV) damage, and their capacity to predict an increased risk of adverse events. These technical advances parallel with the development of novel biological assays that enabled the ex vivo identification of pathways promoting microvascular dysfunction, providing novel potential treatment targets for preventing cerebral-CV disease. In this article, we provide an update of diagnostic testing strategies to detect and characterize microvascular dysfunction and suggestions on how to standardize and maximize the information obtained from each microvascular assay. We examine emerging data highlighting the significance of microvascular dysfunction in the development CV disease manifestations. Finally, we summarize the pathophysiology of microvascular dysfunction emphasizing the role of oxidative stress and its regulation by epigenetic mechanisms, which might represent potential targets for novel interventions beyond conventional approaches, representing a new frontier in CV disease reduction.
... Moreover, Dorner et al. demonstrated a dramatic reduction of FID after experimental inhibition of NO synthase, supporting the assumption that NO is a key driver of retinal endothelial function [70]. An imbalance between NO bioavailability and reactive oxidative species (ROS) has previously been discussed as the key driver for vascular dysfunction [71]. Lower ROS levels have been associated with reduced production of hyperoxide-anion (O2 -) leading to a higher NO bioavailability [72]. ...
Article
Full-text available
Background and aims Physical activity (PA) and fitness are important modulators of vascular ageing and may therefore help expand individual health span. We aimed to systematically review the association of PA and fitness as well as the effects of exercise interventions on the new microvascular biomarkers retinal arteriolar (CRAE) and venular (CRVE) diameters and the retinal flicker light-induced dilatation (FID) in children and adults. Methods PubMed, Ovid, The Cochrane, EMBASE and Web of Science were searched. 805 studies were found, and 25 full-text articles analysed. Twenty-one articles were included in this systematic review. Results Higher PA levels were associated with narrower CRVE in children and adults. Physical inactivity was associated with wider CRVE in both age groups. Combined aerobic and motor skill training in school settings lead to wider CRAE in children. Aerobic exercise interventions in adults with or without CV risk factors induced wider CRAE and narrower CRVE. Studies on the effect of exercise on FID are scarce. In a twelve-week randomized controlled trial, high-intensity interval training significantly improved FID in older patients with CV risk factors. Conclusions Higher PA and fitness levels were associated with improved retinal microvascular health in children and adults. Short-term exercise interventions in healthy children and adults, as well as CV risk patients, improved retinal microvascular structure and function. Exercise has the potential to counteract microvascular remodelling and development of small vessel disease during lifespan. Retinal vessel analysis can differentiate the beneficial effects of exercise on target microvascular organ damage.
Article
Several epigenome studies reported the ability of genes to modulate the lipogenic, and glucogenic pathways during insulin signaling as well as the other pathways involved in the cardiometabolic diseases. Epigenetic plasticity and oxidative stress are interrelated in the pathophysiology of the insulin resistance (IR) and cardiometabolic disease conditions. The aim of this review is to ascertain the previous research evidences pertinent to the role of epigenome and the variations of histone and non-histone proteins during cardiometabolic disease conditions, and insulin signaling finally to develop effective disease-based epigenetic biomarkers and epigenetics-based chromatic therapy. Several public databases including ‘PubMed, National Library of Medicine, and Medline, and google scholar’ were searched for the peer-reviewed & published reports. This study delineated the consistent body of evidences regarding the epigenetic alterations of DNA/histone complexes pertinent to oxidative stress, insulin signaling, metabolic cardiomyopathy, and endothelial dysfunction in the patients with cardiometabolic diseases, T2DM. It has been described that both DNA methylation as well as the post-translational histone alterations across visceral and subcutaneous adipose tissue could facilitate gene transcription to modulate the inflammation, lipogenesis and adipogenesis as complex network of chromatinmodifying enzymatic proteins involved in the defensive insulin signaling across vasculature in the patients with cardiometabolic diseases. Resveratrol, vornistat, trichostatin, and apabetalone are reported to have significant implications as epigenetic modulators. Based on the epigenetic alterations, a wide-range of protein/gene markers such as interleukin-4 (IL-4), interferon-γ (IFNγ) genes may be considered as biomarkers in these patients due to their ability for polarization of immune cells involved in tissue inflammation and atherosclerosis. Hence, it is crucial to unravel the cell specific epigenetic information in advance in order to develop individual risk assessment strategies for chromatin modifying therapies in the patients with cardiometabolic diseases.
Article
Significance: The prevalence of obesity and cardiometabolic phenotypes is alarmingly increasing across the globe and associates with atherosclerotic vascular complications and high mortality. Although multifactorial interventions, vascular residual risk remains high in this patient population, suggesting the need for breakthrough therapies. The mechanisms underpinning obesity-related vascular disease remain elusive and represent an intense area of investigation. Recent advances: Epigenetic modifications - defined as environmentally-induced chemical changes of DNA and histones which do not affect DNA sequence - are emerging as a potent modulator of gene transcription in the vasculature and might significantly contribute to the development of obesity-induced endothelial dysfunction. DNA methylation and histone posttranslational modifications cooperate to build complex epigenetic signals altering transcriptional networks implicated in redox homeostasis, mitochondrial function, vascular inflammation and perivascular fat homeostasis in patients with cardiometabolic disturbances. Critical issues: Deciphering the epigenetic landscape in the vasculature is extremely challenging due to the complexity of epigenetic signals and their function in regulating transcription. An overview of the most important epigenetic pathways is required to identify potential molecular targets to treat or prevent obesity-related endothelial dysfunction and atherosclerotic disease. This would enable the employment of precision medicine approaches in this setting. Future directions: Current and future research efforts in this field entail a better definition of the vascular epigenome in obese patients as well as the unveiling of novel, cell-specific chromatin-modifying drugs able to erase specific epigenetic signals responsible for maladaptive transcriptional alterations and vascular dysfunction in obese patients.
Article
Full-text available
TABLE OF CONTENTS 1. Abstract 2. Introduction 3. Classes and functions of sirtuins 4. Role of sirtuins in the cellular stress response 5. Role of sirtuins in caloric restriction 6. Role of sirtuins during aging and their relation with physical activity 7. Conclusions 8. Acknowledgments 9. References 1. ABSTRACT Recently it has been discovered that Sirtuins represent pivotal regulators of lifespan. Caloric restriction (CR) enhances longevity from yeast to mammals. Whereas the relationship between Sirt-1 and CR is clear, the molecular mechanisms by which Sir2 increases longevity are still unknown. In mammals, CR induces physiological and behavioral changes, and many studies have shown that CR decreases production of reactive oxygen species production thus minimizing oxidative damage, leading to the hypothesis that CR by reducing oxidative stress extends the lifespan by counteraction of aging. In fact, the pathophysiology of aging and age-related diseases involves oxidative stress as an early stage in its development. Recently we found that in aged rats the SIRT1 activity was decreased in heart and adipose tissue, showing as aging is characterized in vivo by a reduced efficiency of this key-regulator of longevity. Whereas several studies have reported that increased physical activity can improve mean life span presumably by reducing mortality risk from many age-related diseases, exercise and longevity studies have failed to document an exercise effect on maximum life span. However, in aged rats a moderate prolonged exercise training is able to induce increase in SIRT1 activity, suggesting that this tool could counteract age-related dysfunctions.
Article
Full-text available
Hyperglycemia and insulin resistance are key players in the development of atherosclerosis and its complications. A large body of evidence suggest that metabolic abnormalities cause overproduction of reactive oxygen species (ROS). In turn, ROS, via endothelial dysfunction and inflammation, play a major role in precipitating diabetic vascular disease. A better understanding of ROS-generating pathways may provide the basis to develop novel therapeutic strategies against vascular complications in this setting. Part I of this review will focus on the most current advances in the pathophysiological mechanisms of vascular disease: (i) emerging role of endothelium in obesity-induced insulin resistance; (ii) hyperglycemia-dependent microRNAs deregulation and impairment of vascular repair capacities; (iii) alterations of coagulation, platelet reactivity, and microparticle release; (iv) epigenetic-driven transcription of ROS-generating and proinflammatory genes. Taken together these novel insights point to the development of mechanism-based therapeutic strategies as a promising option to prevent cardiovascular complications in diabetes.
Article
Full-text available
Aims Aging is a major risk factor for carotid artery disease and stroke. Endothelin-1 (ET-1) and angiotensin II (Ang II) are important modifiers of vascular disease, partly through increased activity of NADPH oxidase and vasoconstrictor prostanoids. Since the renin-angiotensin and endothelin systems become activated with age, we hypothesized that aging affects NADPH oxidase- and prostanoid-dependent contractions to ET-1 and Ang II. Methods Carotid artery rings of young (4 month-old) and old (24 month-old) C57BL6 mice were pretreated with the NO synthase inhibitor L-NAME to exclude differential effects of NO. Contractions to ET-1 and Ang II were determined in the presence and absence of NADPH oxidase (gp91ds-tat) or thromboxane-prostanoid receptor (SQ 29,548) inhibition. Gene expression of endothelin and angiotensin receptors was measured by qPCR. Key findings Aging reduced ET-1-induced contractions and diminished ETA but increased ETB receptor gene expression levels. Gp91ds-tat inhibited contractions to ET-1 in young and to a greater extent old animals, whereas SQ 29,548 had no effect. Ang II-induced contractions were weak compared to ET-1 and unaffected by aging, gp91ds-tat, and SQ 29,548. Aging had also no effect on AT1A and AT1B receptor gene expression levels. Significance Aging in carotid arteries decreases ETA receptor gene expression and responsiveness to ET-1, which nevertheless becomes increasingly dependent upon NAPDH oxidase activity with age; responses to Ang II and gene expression of its receptors are however unaffected. These findings suggest that physiological aging differentially regulates functional responses to G protein-coupled receptor agonists and the signaling pathways associated with their activation.
Article
Whereas vascular aging has been identified as an emerging cardiovascular risk factor, definitions of ‘normal’ and ‘early’ vascular aging (EVA) and their precise relationship with cardiovascular risk are currently equivocal. The present review discusses the concept of vascular aging; that structural and functional changes occur in the large arteries with aging; and EVA; that such age-associated changes are accelerated in individuals at increased cardiovascular risk; and their metrics; indeed, in order to provide a definition of when EVA occurs in clinical practice, reference values of normal and accelerated vascular aging are needed. Due to the complex nature of age-associated changes in the large arteries described above, there are different parameters relating to vascular aging which can be measured. These broadly include aortic and carotid stiffening; aortic and carotid lumen dilation; endothelial dysfunction (usually measured via brachial flow-mediated dilatation); and carotid intima-media thickness.
Article
Increased cyclic stretch to the vessel wall, as observed in hypertension, leads to endothelial dysfunction through increased free radical production and reduced nitric oxide bioavailability. Genetic deletion of the adaptor protein p66(Shc) protects mice against age-related and hyperglycemia-induced endothelial dysfunction, as well as atherosclerosis and stroke. Furthermore, p66(Shc) mediates vascular dysfunction in hypertensive mice. However, the direct role of p66(Shc) in mediating mechanical force-induced free radical production is unknown; thus, we studied the effect of cyclic stretch on p66(Shc) activation in primary human aortic endothelial cells and aortic endothelial cells isolated from normotensive and hypertensive rats. Exposure of human aortic endothelial cells to cyclic stretch led to a stretch- and time-dependent p66(Shc) phosphorylation at Ser36 downstream of integrin α5β1 and c-Jun N-terminal kinase. In parallel, nicotinamide adenine dinucleotide phosphate oxidase activation, as well as production of reactive oxygen species, increased, whereas nitric oxide bioavailability decreased. Silencing of p66(Shc) blunted stretch-increased superoxide anion production and nicotinamide adenine dinucleotide phosphate oxidase activation and restored nitric oxide bioavailability. In line with the above, activation of p66(Shc) increased in isolated aortic endothelial cells of spontaneously hypertensive rats compared with normotensive ones. Pathological stretch by activating integrin α5β1 and c-Jun N-terminal kinase phosphorylates p66(Shc) at Ser36, augments reactive oxygen species production via nicotinamide adenine dinucleotide phosphate oxidase, and in turn reduces nitric oxide bioavailability. This novel molecular pathway may be relevant for endothelial dysfunction and vascular disease in hypertension.
Article
The objective of this study is to detect the p66shc mRNA and protein expression of the peripheral blood monocytes (PBMs) in coronary heart disease patients (CHD) and controls, to evaluate the correlation between the expression of p66shc mRNA in the PBMs and endothelium-dependent vasodilatation. This study included 78 coronary angiography-documented CHD patients (CHD group) and 38 non-CHD controls (control group). The p66shc mRNA and protein levels were determined by quantitative real-time PCR and western blotting. The flow-mediated dilatation (FMD, endothelium-dependent), nitroglycerine-induced dilatation (NID, endothelium-independent) and carotid intimal medial thickness (CIMT) were detected using high-resolution ultrasound. The p66shc mRNA and the protein expression levels in the PBMs were significantly higher in the CHD group compared with the control group (p = 0.007 and 0.001). The FMD (p < 0.001) and NID (p = 0.013) were significantly lower and the CIMT (p = 0.007) was significantly thicker in the CHD patients than in the controls. In the univariate analysis, the expression of the p66shc mRNA in the PBMs was significantly positively correlated with the serum LDL-C and homocysteine levels and the CIMT and was inversely correlated with the FMD and the NID (all p < 0.001). In the multiple linear regression analysis, the FMD (p < 0.001), LDL-C (p = 0.002) and homocysteine levels (p = 0.002) remained independently correlated with the p66shc mRNA expression. These findings highlight a pivotal role for the expression of p66shc in CHD and endothelial dysfunction, which might represent a molecular target to prevent endothelial dysfunction-related disease.
Article
Endothelial nitric oxide (NO) is generated by constitutively active endothelial nitric oxide synthase (eNOS), an essential enzyme responsible for cardiovascular homeostasis. Historically, endothelial NO was first recognized as a major vasodilator involved in control of vasomotor function and local blood flow. In this review, our attention is focused on the emerging role of endothelial NO in linking cerebrovascular function with cognition. We will discuss the recognized ability of endothelial NO to modulate processing of amyloid precursor protein (APP), influence functional status of microglia, and affect cognitive function. Existing evidence suggests that the loss of NO in cultured human cerebrovascular endothelium causes increased expression of APP and β-site APP-cleaving enzyme 1 (BACE1) thereby resulting in increased secretion of amyloid β peptides (Aβ1-40 and Aβ1-42). Furthermore, increased expression of APP and BACE1 as well as increased production of Aβ peptides was detected in the cerebral microvasculature and brain tissue of eNOS-deficient mice. Since Aβ peptides are considered major cytotoxic molecules responsible for the pathogenesis of Alzheimer's disease, these observations support the concept that a loss of endothelial NO might significantly contribute to the initiation and progression of cognitive decline. In addition, genetic inactivation of eNOS causes activation of microglia and promotes a pro-inflammatory phenotype in the brain. Behavioural analysis revealed that eNOS-deficient mice exhibit impaired cognitive performance thereby indicating that selective loss of endothelial NO has a detrimental effect on the function of neuronal cells. Together with findings from prior studies demonstrating the ability of endothelial NO to affect synaptic plasticity, mitochondrial biogenesis, and function of neuronal progenitor cells, it is becoming apparent that the role of endothelial NO in the control of central nervous system function is very complex. We propose that endothelial NO represents the key molecule linking cerebrovascular and neuronal function.
Article
Due to the aging and increasingly complex nature of our patients, frailty has become a high-priority theme in cardiovascular medicine. Despite the recognition of frailty as a pivotal element in the evaluation of older adults with cardiovascular disease (CVD), there has yet to be a roadmap to facilitate its adoption in routine clinical practice. Thus, we sought to synthesize the existing body of evidence and offer a perspective on how to integrate frailty in clinical practice. Frailty is a biological syndrome that reflects a state of decreased physiologic reserve and vulnerability to stressors. Upwards of 20 frailty assessment tools have been developed with most tools revolving around the core phenotypic domains of frailty – slow walking speed, weakness, inactivity, exhaustion, and shrinking – as measured by physical performance tests and questionnaires. The prevalence of frailty ranges from 20-60% depending on the cardiovascular pathology as well as the tool and cutoff chosen to define frailty. Epidemiologic studies have consistently demonstrated that frailty carries a relative risk of >2 for mortality and morbidity across a spectrum of stable CVD, acute coronary syndromes, heart failure, surgical and transcatheter interventions. Frailty contributes valuable prognostic insights incremental to existing risk models, and assists clinicians in defining optimal care pathways for their patients. Interventions designed to improve outcomes in frail elders with CVD such as multi-disciplinary cardiac rehabilitation are being actively tested. Ultimately, frailty should not be viewed as a reason to withhold care, but rather as a means of delivering it in a more patient-centered fashion.
Article
Risk of diabetic complications continues to escalate overtime despite a multifactorial intervention with glucose-lowering drugs, anti-hypertensive agents and statins. In this perspective, a mechanisms-based therapeutic approach to vascular disease in diabetes represents a major challenge. Epigenetic signatures are emerging as important determinants of vascular disease in this setting. Methylation and acetylation of DNA and histones is a reversible process leading to dysregulation of oxidant and inflammatory genes such as mitochondrial adaptor p66(Shc) and transcription factor NF-kB p65. Epigenetic modifications associated with diabetes may contribute to the early identification of high risk individuals. Ongoing epigenomic analyses will be instrumental in identifying the epigenetic variations that are specifically associated with cardiovascular disease in patients with diabetes. Here, we describe a complex scenario of epigenetic changes and their putative link with diabetic vascular disease. Pharmacological reprogramming of diabetes-induced epigenetic signatures may be a promising option to dampen oxidative stress and inflammation, and thus prevent cardiovascular complications in this setting.
Article
The endothelium is located in a strategic anatomical position within the blood vessel wall and thus constitutes a barrier between the blood and all tissues. The integrity of the endothelial cells, which line the entire circulatory system like wallpaper, is essential to prevent the onset of cardiovascular disorders. Aging is one of the major risk factors for the development of heart and vascular diseases. However, over the past years it has become clear that the functional capacity of endothelial cells declines with age and that physiological aging occurs independently of pathological changes. One important mechanism contributing to the onset of the aging process is the disturbance of the cellular redox homeostasis. Two key molecules involved in maintaining the delicate balance between oxidative and antioxidative systems are NADPH oxidase 4, an enzyme whose sole function is to produce reactive oxygen species and the oxidoreductase thioredoxin-1, which reduces oxidized proteins. Therefore, this review will focus on the role of these two proteins in cardiovascular aging.