ArticlePDF AvailableLiterature Review

Impact of Red Wine Consumption on Cardiovascular Health

Authors:

Abstract and Figures

Background: The devastating effects of heavy alcohol drinking have been long time recognized. In the last decades, potential benefits of modest red wine drinking were suggested. In European countries in which red wide intake is not negligible (such as France), the association between cholesterol and cardiovascular (CV) risk was less evident, suggesting the action of some protective molecules in red wine or other foods and drinks. Methods: This narrative review is based on the material searched for and obtained via PubMed up to May 2016. The search terms we used were: "red wine, cardiovascular, alcohol" in combination with "polyphenols, heart failure, infarction". Results: Epidemiological and mechanistic evidence of a J-shaped relationship between red wine intake and CV risk further supported the "French paradox". Specific components of red wine both in vitro and in animal models were discovered. Polyphenols and especially resveratrol largely contribute to CV prevention mainly through antioxidant properties. They exert beneficial effects on endothelial dysfunction and hypertension, dyslipidemia, metabolic diseases, thus reducing the risk of adverse CV events such as myocardial infarction ischemic stroke and heart failure. Of interest, recent studies pointed out the role of ethanol itself as a potential cardioprotective agent, but a clear epidemiological evidence is still missing. The aim of this narrative review is to update current knowledge on the intracellular mechanism underlying the cardioprotective effects of polyphenols and ethanol. Furthermore, we summarized the results of epidemiological studies, emphasizing their methodological criticisms and the need for randomized clinical trials able to clarify the potential role of red wine consumption in reducing CV risk. Conclusion: Caution in avowing underestimation of the global burden of alcohol-related diseases was particularly used.
Content may be subject to copyright.
Send Orders for Reprints to reprints@benthamscience.ae
Current Medicinal Chemistry, 2017, 24, 1-21 1
REVIEW ARTICLE
0929-8673/17 $58.00+.00 © 2017 Bentham Science Publishers
Impact of Red Wine Consumption on Cardiovascular Health
Luca Liberalea#, Aldo Bonaventuraa#, Fabrizio Montecuccoa,b,c*, Franco Dallegria,b and
Federico Carbonea
aFirst Clinic of Internal Medicine, Department of Internal Medicine, University of Genoa, 6 viale Benedetto XV, 16132
Genoa, Italy; bIRCCS AOU San Martino - IST, Genova, 10 Largo Benzi, 16132 Genoa, Italy; cCentre of Excellence for
Biomedical Research (CEBR), University of Genoa, 9 viale Benedetto XV, 16132 Genoa, Italy
A R T I C L E H I S T O R Y
Received: November 11, 2016
Revised: March 05, 2017
Accepted: March 05, 2017
DOI:
10.2174/0929867324666170518100606
Abstract: The devastating effects of heavy alcohol drinking have been long time recog-
nized. In the last decades, potential benefits of modest red wine drinking were suggested.
In European countries in which red wide intake is not negligible (such as France), the as-
sociation between cholesterol and cardiovascular (CV) risk was less evident, suggesting
the action of some protective molecules in red wine or other foods and drinks. Epidemiol-
ogical and mechanistic evidence of a J-shaped relationship between red wine intake and
CV risk further supported the “French paradox”. Specific components of red wine both in
vitro and in animal models were discovered. Polyphenols and especially resveratrol
largely contribute to CV prevention mainly through antioxidant properties. They exert
beneficial effects on endothelial dysfunction and hypertension, dyslipidemia, metabolic
diseases, thus reducing the risk of adverse CV events such as myocardial infarction
ischemic stroke and heart failure. Of interest, recent studies pointed out th e role of ethanol
itself as a potential cardioprotective agen t, but clear epidemiological evidence is still
missing. The aim of this narrative review is to update current knowledge on the intracellu-
lar mechanism underlying the cardio protective effects of polyphenols and ethanol. Fur-
thermore, we summarized the results of epidemiological studies, emphasizing their meth-
odological criticisms and the need for randomized clinical trials able to clarify the poten-
tial role of red wine consumption in reducing CV risk. Caution in avowing underestima-
tion of the global burden of alcohol-related diseases was particularly used.
Keywords: Alcohol, ethanol, flavonoids, polyphenols, red wine, resveratrol
1. INTRODUCTION
More than twenty years ago, a lower incidence of
cardiovascular (CV) disease was observed in the
French population, despite their dietary intake of satu-
rated fats [1]. This unexpected observation, also known
as “French paradox”, raised the interest on the potential
protective effects of red wine on CV health. The find-
ing of a U-/J-shaped relationship between red wine
*Address correspondence to this author at the First Clinic o f Inter-
nal Medicine, Department of Internal Medicine, IRCCS AOU San
Martino - IST, and Centre of Excellence for Biomedical Research
(CEBR), University of Genoa. 6 viale Benedetto XV, 16132 Genoa,
Italy; Tel: +39 010 353 86 94; Fax: +39 0101 353 86 86;
E-mail: fabrizio.montecucco@.unige.it
#These authors equally contributed as the first author to this work
intake and CV risk was the first epidemiological evi-
dence explaining the “French paradox” [2]. Further
studies on red wine composition focused on polyphe-
nols contained in red wine and their antioxidant proper-
ties [3]. Several in vitro and animal studies investigat-
ing the polyphenol content in red wine (and partially in
the beer) strongly suggested their active protective role
in CV prevention [4]. More recently, also the Mediter-
ranean diet, characterized by moderate intake of red
wine during meal, has been associated with some
benefit in the prevention of CV diseases and cancer [5].
In this context, both the European Society of Cardiol-
ogy and the European Atherosclerosis Society sug-
gested a moderate alcohol consumption in non-
abstainers with normal levels of triglycerides (TAG),
2 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
defined as up to 20-30 g/day for men and 10-20 g/day
for women [6, 7].
In addition, WHO recently defined light, moderate
and heavy drinkers, as reported in Table 1 [8].
Table 1. Alcohol consumption definition based on
drinking quantity and frequ ency.
LIGHT DRINKER
1 drink/day
MODERATE DRINKER
2-3 drinks/day
HEAVY DRINKER
4 drinks/day
standard drink equivalent 12.5 g of absolute ethanol
100 ml wine (13% alc/vol)
240 ml regular beer (4.9% alc/vol)
33 ml spirit (40% al/vol)
Nevertheless, clinical studies have not always con-
firmed experimental evidence. These contrasting re-
sults may be due to different confounding factors such
as age, gender, comorbidities and drinking pattern. In
addition, it should also consider that polyphenolic me-
tabolites reach very low concentrations in the blood-
stream. On this basis, the aim of this narrative review is
to update the knowledge of different classes of poly-
phenols contained in red wine, focusing on experimen-
tal and clinical evidence. In addition, potential benefi-
cial effects of low-moderate ethanol consumption will
be discussed. Finally, epidemiological evidence sup-
porting a clinical association between red wine intake
and low CV risk will be critically discussed.
2. POLYPHENOLS
2.1. Structure, Biosynthesis, and General Properties
Grapes contain a lot of polyphenols, a large group
of compounds having an aromatic ring(s), character-
ized by the presence of one or more hydroxyl groups
with varying structural complexities (Fig. 1A). Poly-
phenols are classified as flavonoids and non-flavonoid
compounds. Flavonoids include proanthocyanidins an-
thocyanins and flavonols. Non-flavonoid polyphenols
include stilbene (also known as resveratrol), but also
saponin, curcumin, and tannins. Quantification of die-
tary intake of polyphenols is quite difficult [9, 10], and
this is due to the large variability of the phenolic source
and bioavailability [11]. In this regard, physical proper-
ties of the solution are a major determinant of polyphe-
nol bioavailability. Thus, as compared with wine, the
high alcohol content of whiskey improves polyphenol
absorption [12], whereas absorption of flavonols was
demonstrated to be higher in onion or black tea [13].
The major polyphenols contained in red wine are an-
thocyanin and catechins [14], but the polyphenolic
composition is largely dependent on soil composition,
geographic position, plant disease and farming tech-
nique. Biological activities of polyphenols are depend-
ent on their pharmacokinetic properties. Lipophilic pro-
perties of procyanidins, quercetin and flavonols pro-
mote a passive transport through enteric cells mem-
branes, whereas absorption of hydrophilic polyphenols
(ingested as esters, glycosides, or polymers) require
enzymatic hydrolysis [15] and conjugation by sul-
fation, methylation and glucuronidation [16]. However,
all classes of polyphenols undergo hepatic metabolism,
so that serum concentrations of original compounds are
very low (usually less than 1µmol/l), and circulating
polyphenols are mostly active metabolites.
2.2. Flavonoid Polyphenols
2.2.1. Flavonoid Polyphenol-Mediated Signaling
Polyphenols have beneficial effects on different bio-
logical pathways. In addition to promoting free-radical
scavenging, metal chelation, and enzyme modulation,
they have major effects on cell cycle, xenobiotic me-
tabolism, immune-related factors and transcription
(about 2200 transcripts have been estimated to be di-
rectly influenced by flavonoids) [17]. Among polyphe-
nols, flavonoids mediated the most of the anti-oxidant
properties ascribed to red wine, including inhibition of
lipid oxidation, peroxide generation and lipid peroxida-
tion [18-20]. The flavonoid quercetin was also shown
to enhance neuronal resistance to age-related diseases
in mice via AMPK activation [21]. Furthermore, quer-
cetin may prevent adipogenesis by inducing apoptosis
of 3T3-L1 pre-adipocytes via AMPK-mediated activa-
tion of ERK and JNK pathways [22]. However, a sig-
nificant reduction of fat accumulation in mature adipo-
cytes was observed only with a very high dose of quer-
cetin [23]. Alongside AMPK, also, sirtuins (SIRT)
have been described as flavonoid target. In particular,
polyphenol-dependent SIRT-1 activation has been
shown to increase brain and muscle mitochondrial bio-
genesis [24, 25], as well as to suppresses endothelial
oxidative injuries induced by oxLDL [26]. Finally, die-
tary quercetin in mice has also been reported to reduce
macrophage infiltration and inflammation in adipose
tissue, likely through an AMPK/SIRT1-mediated
mechanism [27].
An anti-estrogenic role of polyphenols has been
proposed but it still need fully elucidation [28]. An es-
trogen receptor (ER)-dependent activation of endothe-
lial nitric oxide synthase (eNOS) was first shown in
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 3
bovine aortic endothelial cells exposed to black tea
polyphenols [29]. Then, ER-mediated benefits on endo-
thelial activity have been demonstrated for several spe-
cific flavonoids such as delphinidin, epicatechin, sy-
ringic acid, apigenin, malvidin and ellagic acid [30,
31]. However, these results were not later confirmed in
cultured endothelial cells of rats [32], and also a recent
study on ovariectomized mice described pro-angio-
genic properties of red wine polyphenols as independ-
ent of ERα activation [33].
2.2.2. Flavonoid Polyphenol in Endothelial Dysfunc-
tion and Hypertension
Regardless of the mechanism, the protective role of
red wine polyphenols on CV health is largely related to
the up-regulation of the eNOS activity. Polyphenols-
induced vasorelaxation has been demonstrated in dif-
ferent experimental models such as rat aortic rings and
mesenteric arteries [34-36]. In later studies on the iso-
lated porcine coronary artery, electron spin resonance
4 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
spectroscopy analysis demonstrated an enhanced endo-
thelial generation of nitric oxide (NO) and endothe-
lium-derived hyperpolarizing factor, as a result of PI3-
kinase/Akt and Src phosphorylation [37, 38]. There-
fore, anti-hypertensive effects of polyphenol-rich com-
pounds were observed in various animal models. Red
wine polyphenols have been shown to reduce arterial
blood pressure in angiotensin II-induced hypertensive
mice, together with preventing vascular superoxide
anion production and NADPH oxidase subunit expres-
sion [39]. Similar beneficial results have been also ob-
served in hypertensive rats treated with specific fla-
vonoid compounds such as provinol, proanthocyanid-
ins, and catechins [40-42].
2.2.3. Anti-atherosclerotic Properties of Flavonoid
Polyphenol
Aside from preventing endothelial dysfunction, red
wine flavonoids were reported to exert anti-atheroscle-
rotic effects in different experimental studies. Treat-
ment with quercetin in diet-induced atherosclerotic
hamsters was associated with significant reduction of
aortic fatty streaks and improved lipid profile, charac-
terized by low total plasma cholesterol and increased
apolipoprotein (apo)A-I levels [43]. Furthermore, the
administration of flavonoids-rich grape juice has been
shown to reduce the detrimental effects of hypercholes-
terolemic diet in rabbit, by reducing serum cholesterol,
blood pressure, and platelet aggregation [44]. Interest-
ingly, this intriguing effect of flavonoids on platelet
activity was confirmed in two clinical studies. Quer-
cetin and catechin showed a synergic effect in sup-
pressing collagen-induced platelet aggregation and
platelet adhesion to collagen [45], whereas platelet in-
cubation with diluted purple grape juice was associated
with increased NO release and decreased superoxide
production [46].
2.2.4. Flavonoid Polyphenols in Metabolic Diseases
and Inflammation
Anti-oxidant properties of polyphenols have also
been shown to be effective in metabolic diseases. By
suppressing NADPH oxidase activity, polyphenol-rich
grape extract prevents the development of insulin resis-
tance, hypertension, cardiac hypertrophy and endothe-
lial dysfunction in high fructose-fed rats [47]. Simi-
larly, through the suppression of NADPH oxidase ac-
tivity, catechin normalized blood pressure and pre-
vented endothelial dysfunction and insulin resistance in
pre-diabetic rat (Otsuka Long-Evans Tokushima Fatty
[OLETF] model) [48], whereas provinol improved car-
diac performance in Zucker fatty rats, as evidenced by
an increase in left ventricular fractional shortening and
cardiac output associated with decreased peripheral
arterial resistances [49]. On the other hand, many bene-
ficial effects of polyphenols on CV health may be as-
cribed to their anti-inflammatory properties. Alhough,
anti-inflammatory properties of polyphenols are in part
associated with antioxidant activity, they also showed
capacity to directly modulate cytokine expression.
Treatment with procyanidins has been shown to sup-
press the release of interleukin (IL)-6 and monocyte
chemo-attractant protein-1 from adipocyte and macro-
phage-like cells, in addition, to reduce the plasmatic
levels of C-reactive protein (CRP) and IL-6 in high fat-
fed rats [50].
2.2.5. Flavonoid Polyphenols and Cardiovascular
Risk: Clinical Evidence
Alongside experimental evidence, many clinical
studies investigated the potential protective role of
polyphenols on CV health [51]. In healthy subjects, the
endothelial function assessed by flow-mediated dilata-
tion (FMD) increased after the consumption of 3mL/kg
or two glasses of red wine with or without alcohol [52-
54]. This effect was moreover amplified in a synergis-
tic manner, by combining red wine with other compo-
nents of Mediterranean diet, such as olive oil [55].
Beneficial effects on FMD were also observed after the
administration of non-alcoholic beverages or dark cho-
colate, thus emphasizing the role of polyphenols [56-
59]. Interestingly, these results have been confirmed in
several cohorts with high CV risk such as post-
menopausal women [60], smokers [61-63], and meta-
bolic disorders (hypercholesterolemia, diabetes, over-
weight) [61-67]. Finally, protective effects of polyphe-
nols on vascular function (defined as increased FMD
and reduced intima-media thickness) have been ob-
served after administration of black tea, epigallocate-
chin gallate-rich green tea, and red grape juice pome-
granate juice in patients with hypertension [68-71] and
established coronary heart disease [72-77].
2.3. Resveratrol and Non-flavonoid Polyphenols
Resveratrol (C14H12O3; 3,5,4’-trihydroxystilbene)
is a natural polyphenolic compound characterized by two
phenol rings linked by a styrene double bond (Fig. 2).
It is produced by a lot of plant species and fruits,
such as purple grapes, blueberries, mulberries, cranber-
ries, rhubarb, peanuts, groundnuts, and pines [78-80].
Grapes are the most important source of resveratrol for
humans, since the compound is present in the wine, in
particular, the red wine (concentration between 1 and
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 5
14 mg/L) [81]; their highest concentration is found in
the skin and seeds of grapes (50-100 µg/g equal to 5-
10% of their biomass) [82]. The main enzyme respon-
sible for resveratrol biosynthesis is stilbene synthase
(STS), which is rapidly activated in response to exoge-
nous or environmental stress factors, such as injury,
ultra-violet irradiation, and fungal attack [83]. STS
catalyzes 3 condensation reactions between coumaroyl-
coenzyme A (CoA) and 3 molecules of malonyl-CoA
via cleavage of 3 carbon dioxide molecules and the loss
of the terminal carboxyl group, leading to the produc-
tion of the C14 molecule resveratrol [83, 84]. Resis-
tance against stress factors is warranted if the phytoa-
lexin accumulation occurs rapidly on the exposed site.
Typically, resveratrol levels peak 24 hours after stress
exposure and decline after 42-72 hours following stil-
bene oxidase activation [82, 85]. Resveratrol presents
with pleiotropic beneficial effects on health including
antioxidant, anti-inflammatory, cardioprotective and
anti-tumor properties [86].
2.3.1. Resveratrol-mediated Signaling and Endothe-
lial Dysfunction
In animal models of hypertension, diabetes, and hy-
percholesterolemia, oral treatment with resveratrol re-
sulted in the improvement of the endothelium-depen-
dent relaxation [86-88]. Similar results were found in
overweight/obese men and post-menopausal women
with untreated, borderline hypertension, who showed
an increase in the FMD of the brachial artery [89]. The
enhancement of the endothelial function is mostly due
to the eNOS-mediated NO generation through at least
seven mechanisms [90]: (i) the stimulation of eNOS
gene transcription and the increase in the eNOS mes-
senger ribonucleic acid stability, partly due to the up-
regulation of eNOS expression by the SIRT-1 [91, 92];
(ii) the stimulation of eNOS phosphorylation mediated
by the estrogen receptor α and the signaling pathway
concerning the α-subunit of G-protein, caveolin (Cav)-
1, the tyrosine kinase Src, and the mitogen-activated
protein kinase (MAPK)/extracellular signal-regulated
kinase 1/2 [93, 94]; (iii) the induction of SIRT-1-
mediated deacetylation of lysines 496 and 506 in the
calmodulin-binding domain of eNOS [95]; (iv) the de-
crease of intracellular levels of the endogenous eNOS
inhibitor asymmetric dimethylarginine [96]; (v) the
reduction of Cav-1 levels and Cav-1/eNOS interactions
[94, 97]; (vi) the blockage of eNOS uncoupling by
augmenting the eNOS cofactor tetrahydrobiopterin
(BH4) intracellular levels and the prevention of the
BH4 oxidation by decreasing the oxidative stress [98];
and (vii) the amelioration of NO bioactivity via the in-
hibition of superoxide-mediated NO inactivation [99].
On the basis of controversial results in some animal
studies, resveratrol cannot be considered a direct
SIRT1 activator [100-102], although an enzymatic ac-
tivation of SIRT1 by resveratrol in vivo cannot be kept
out [103, 104]. In this regard, resveratrol effects might
also be mediated by an up-regulation of SIRT1 expres-
sion [92, 105]. Moreover, the SIRT1 activation by res-
veratrol in vivo is thought to be an indirect effect medi-
ated by AMPK as an upstream target [106]. Overall,
6 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
the increased release of endothelial NO induced by res-
veratrol is protective for blood vessels because of its
antihypertensive (in terms of vasodilation), anti-throm-
botic (by the inhibition of platelet aggregation), and anti-
atherosclerotic (avoiding leukocyte adhesion to vascu-
lar endothelium, reducing the low-density lipoprotein
(LDL) oxidation, and inhibiting vascular smooth mus-
cle cell proliferation) characteristics.
2.3.2. Resveratrol, Vascular Inflammation, and Oxi-
dative Stress
Resveratrol treatment was found to block the pro-
duction of reactive oxygen species (ROS) induced by
tumor necrosis factor (TNF)-α, activation of NF-κB,
intercellular adhesion molecule-1, and IL-6 [107, 108].
Given that NF-κB is believed the most important target
of resveratrol, the putative mechanisms of inhibition
are the following: (i) reduction of H2O2 levels; (ii) in-
hibition of the phosphorylation of IκB, which is re-
sponsible for the NF-κB activation; (iii) the inhibition
of p65 phosphorylation, that is necessary for the tran-
scriptional activation of NF-κB; (iv) the possible
deacetylation of the p65 subunit of NF-κB by SIRT1
[109-111]. Resveratrol was shown to be an ROS scav-
enger, even if its direct antioxidant activity is relatively
poor and can be referred to the up-regulation of the en-
dogenous cellular antioxidant systems [112, 113]. Res-
veratrol may induce antioxidant enzymes: superoxide
dismutase (SOD)1, SOD2 and SOD3 in endothelial
cells, and SOD2 in cardiac myoblasts [114-116]. Res-
veratrol also increases the expression of thioredoxin
(Trx)-1 and -2, glutaredoxin (Grx)-1 and -2, heme oxy-
genase (HO)-1, NADPH, quinone oxidoreductases, and
γ-glutamylcysteine synthetase, the rate-limiting en-
zyme for glutathione synthesis [117-120]. Although
SIRT1 and nuclear factor-E2-related factor-2 (Nrf2)
play a critical role in this process, molecular mecha-
nisms involved in the induction of antioxidant enzymes
by resveratrol are not yet fully known. Moreover, res-
veratrol was demonstrated to decrease oxidative stress
by inhibiting ROS production. In the heart of hyper-
cholesterolemic mice and in the trauma-related hemor-
rhagic aorta of rats, the expression of NADPH oxidase
(NOX)1, NOX2, and NOX4 was reduced by resvera-
trol treatment [98, 121]. Resveratrol also reduces the
activity of the NOX enzyme complex by hitting the
regulatory subunits [122]. Finally, in platelets, protein
kinase C (PKC)-mediated phosphorylation of p47phox
is blocked by resveratrol [123].
2.3.3. Resveratrol and Hypertension
Anti-hypertensive effects of resveratrol have been
studied in animal models of hypertension and hyper-
tension combined with insulin resistance [124, 125].
High doses of resveratrol lowered high pressure and
prevented cardiac hypertrophy in two different hyper-
tensive animal models [126, 127], whereas some stud-
ies underscored the ability of resveratrol to reverse car-
diac hypertrophy and contractile dysfunction [128,
129]. With regard to clinical studies, a meta-analysis
with 247 subjects recently showed that only high doses
of resveratrol (150 mg per day) were able to signifi-
cantly reduce blood pressure [130]. The mechanisms
underlying the anti-hypertensive properties of resvera-
trol are mainly endothelium-dependent, characterized
by increased NO availability due to an up-regulation of
eNOS strictly linked to AMPK, SIRT1, and Nrf2 acti-
vation [124]. On the other hand, the inhibition of vas-
cular smooth muscle cells contractility, via AMPK ac-
tivation, represents an endothelium-independent mech-
anism [131].
2.3.4. Resveratrol and Atherosclerosis
In the hypercholesterolemic rabbit model, resvera-
trol reduced the size, density, and mean area of athero-
sclerotic plaque [132]. It was also shown to decrease
plasma TAG and LDL-cholesterol (LDL-c) levels, and
increase HDL-cholesterol (HDL-c) [133], in addition to
enhancing the expression of the LDL receptors in hepa-
tocytes in vitro [134]. However, clinical studies have so
far provided controversial results [135]. However, res-
veratrol was shown to lower LDL-c in diabetic patients
[136] as well as to reduce plasma TAG in obese men
[137] and smokers [137]. Notably, in high CV risk pa-
tients under statin treatment for primary prevention,
resveratrol was shown to reduce oxidized LDL and
LDL-c [138]. Different mechanisms may contribute to
the anti-atherosclerotic properties of resveratrol: stimu-
lation of NO production [139]; inhibition of LDL oxi-
dation [140]; inhibition of smooth muscle cell prolif-
eration [141, 142]. A possible contribution of SIRT1 in
blocking atherogenesis is also to be considered, but
further studies are required to clarify this issue [91].
2.3.5. Resveratrol and Diabetes
Resveratrol as also showed to reduce blood glucose
and to protect pancreatic β-cells from oxidative damage
in animal studies [143], to attenuate diabetic vascular
complications [144], and to decrease diabetic cardio-
myopathy [145]. In obese men treated with resveratrol
(150 mg per day) for 30 days, modest metabolic
changes, including insulin resistance, have been in-
duced [137]. The anti-hyperglycemic effects of res-
veratrol are likely to be dependent on increasing glu-
cose uptake by peripheral tissues, as a result of en-
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 7
hanced glucose transporter-4 translocation of to the
plasma membrane [146] and the increase in glycogen
storage via glycogen synthase and glycogen phos-
phorylase activities [147]. Moreover, resveratrol was
demonstrated to potentiate the glucose-stimulated insu-
lin secretion in β-cells through SIRT1-dependent
mechanisms [148], in addition to potentiating the oxi-
dation of fatty acids and decrease their synthesis via
AMPK activation [149]. Therefore, SIRT1 and AMPK
are both crucial also in mediating the effects of resvera-
trol on insulin resistance [150, 151].
2.3.6. Resveratrol and Ischemic Heart Disease
Resveratrol is protective against ischemic cardiac
disease through different mechanisms. First of all, it
prevents myocardial infarction due to the inhibition of
platelet aggregation, in both healthy subjects and in
aspirin-resistant patients [152, 153]. The antiplatelet
effect may be mediated by the NO diffusion into plate-
lets thus activating the guanylyl cyclase and producing
cyclic guanosine monophosphate, by blocking PKC-
mediated phosphorylation, as previously mentioned,
and by inhibiting the pathway involving p38 MAP
kinase leading to a reduced thromboxane synthesis
[123]. Secondly, resveratrol protects the myocardial
tissue from ischemia/reperfusion injury, characterized
by great inflammatory burst [154, 155], precondition-
ing-like effect [117] and regeneration of infarcted
myocardial tissue [156]. In this context, by up-
regulating eNOS expression, resveratrol may determine
an NO-dependent increase of HO-1, both identified as
of preconditioning effect in the ischemic heart [157,
158]. Additional pathways activated by resveratrol may
include those triggered by adenosine receptors A1 and
A3, and cyclic adenosine monophosphate (cAMP) re-
sponse element-binding protein (CREB) [159]. Also
the SIRT1-FOXO1 pathway was supposed to be in-
volved in the anti-apoptotic effect of resveratrol on
heart ischemia [160], as well as in the induction of
autophagy [161]. The regeneration of infarcted myo-
cardium was demonstrated in a resveratrol-pretreated
rat left anterior descending coronary artery occlusion
model by injecting adult cardiac stem cells in the bor-
der zone of the myocardium due to a significant reduc-
tion of oxidative stress, increase in cell survival and
following proliferation of cardiomyocytes [156]. Fur-
thermore, resveratrol was shown to enhance myocar-
dial angiogenesis in the peri-infarct myocardium both
in vivo and in vitro by inducing vascular endothelial
growth factor (VEGF). This effect is regulated by Trx-
1 and HO-1 signaling and has been associated with
early improvement of left ventricular function after
myocardial infarction [117].
2.3.7. Resveratrol and Heart Failure
In experimental studies, pre-treatment with resvera-
trol showed prevention of cardiac hypertrophy and im-
provement of cardiac function [162, 163]. Some clini-
cal benefit of resveratrol administration was also dem-
onstrated in secondary prevention of heart failure [164,
165]. Nevertheless, no clinical evidence supporting a
beneficial role of resveratrol in patients with heart fail-
ure is currently available. Conversely, patients with
stable coronary artery disease treated with resveratrol
demonstrated an improvement of diastolic function af-
ter myocardial infarction [166]. Potential mechanisms
explaining the effects of resveratrol on the pathophysi-
ology of cardiac hypertrophy and heart failure include a
decrease oxidative stress due to increased expression of
SOD2 [116], augmented levels of glutathione, eNOS
un-regulation, AMPK/Akt-mediated inhibition of pro-
tein synthesis [167], improvement of calcium cycling
via SIRT1 [145], and inhibition of hypertrophic gene
expression.
2.3.8. Resveratrol and Stroke
Resveratrol was demonstrated to protect against
ischemic stroke in adult rats, in part by its anti-oxidant
effect on the cerebrovascular endothelium [168]. This
effect might depend on SIRT1 because its inhibition
with sirtinol partially blocked the protection of endo-
thelial cells. [169]. Resveratrol decreased the infarct
size in a rat model of focal cerebral ischemia as well as
neuroprotective effects have been described [170, 171].
The mechanisms of neuroprotection were recently pro-
posed; resveratrol inhibited phosphodiesterases and
regulated the cAMP/AMPK/SIRT1 pathway, which is
responsible for reduced adenosine triphosphate (ATP)
energy consumption during ischemia and so it in-
creased ATP, phospho-AMPK, SIRT1, and cAMP lev-
els [172]. Considering that resveratrol is able to inhibit
cyclooxygenases (COX), such as COX1 and COX2
[173], the in vivo neuroprotection induced by resvera-
trol might be associated with the abrogation of COX2
in the substantia nigra that limits 6-hydroxydopamine-
induced toxicity [174]. Currently, no investigation has
been conducted in stroke patients, even if resveratrol
showed to increase cerebral blood flow in healthy adult
subjects and to enhance cerebrovascular perfusion in
postmenopausal women [175-177]. Nevertheless, res-
veratrol was found as a potential adjuvant of recombi-
nant tissue-type plasminogen activator (r-tPA) in the
treatment of ischemic stroke [178]. It was shown to
inhibit the expression or the activity of metalloprotein-
ase (MMP)-2 and MMP-9 both in cells lines and in
animal models of stroke [179-181]. These properties
8 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
may favour the extension of the time window for r-tPA
treatment and the clinical improvement in stroke pa-
tients with late treatment. The inhibitory effect of res-
veratrol is especially found when administered together
with r-tPA, this explaining the reduction in brain
ischemia injuries and the positive correlation between
changes in MMP levels and National Institutes of
Health Stroke Scale scores.
3. ETHANOL
Alongside protective effects of polyphenols, nearly
all epidemiological studies reported a J-shaped curve,
whereby light to moderate ethanol consumption was
associated with lower risk of adverse CV events as
compared with abstainers, while heavy drinkers dem-
onstrated an increased risk [182]. More specifically, a
regular light to moderate consumption of ethanol was
shown to decrease by 30-35% the risk of heart disease
in both men and women, regardless of age [183, 184].
This effect was further increased by healthy lifestyle
behaviors, such as Mediterranean diet, smoking absten-
tion, regular physical activity and low body mass index
(<25 Kg/m2) [185, 186]. Similarly, low to moderate
ethanol consumption has been associated with reduced
risk of ischemic stroke [187] and congestive heart fail-
ure [188]. Cardioprotective effect of ethanol has been
ascribed to a synergic effect involving different targets.
3.1. Ethanol and Endothelial Dysfunction
A J-shaped curve was used also to describe the rela-
tionship between endothelial function and ethanol con-
sumption. Specifically, low-moderate alcohol intake
was shown to increase nitric oxide bioavailability, due
to the up-regulation of eNOS as observed in rats [189,
190]. In addition, ethanol may induce the activation of
transient receptor potential vanilloid 1 channels on
perivascular sensory nerve terminal, which promotes
the release of potent vasodilator agent calcitonin gene-
related peptide [191]. Furthermore, the TNF-α-induced
expression of adhesion molecule [192] was shown to
be down-regulated by red wine consumption in vitro
[193]. As result, in both cross-sectional and random-
ized clinical trials moderate alcohol consumption was
shown to improve endothelial function in healthy sub-
jects, either assessed as FMD or coronary flow velocity
reserve [194-196]. Partially related to endothelial func-
tion, chronic low ethanol consumption was also shown
to reduce systolic blood pressure in rats [197], whereas
the results from clinical studies remain controversial. In
this regard, a recent meta-analysis only identified a
trend toward decreased risk of hypertension [198].
3.2. Ethanol and Dyslipidemia
Low-to-moderate ethanol consumption has been re-
ported to modify composition and function of high-
density lipoprotein cholesterol (HDL-c). A rise of
smaller HDL-c associated with increased concentration
of HDL-c apoA-I and apoA-II has been reported [199,
200], alongside an increase of phospholipids and poly-
unsaturated fatty acid content of HDL-c [201]. Func-
tionally, ethanol ingestion may enhance the activity of
HDL-c as reverse cholesterol transport [202-204], as
well as their anti-atherosclerotic and anti-inflammatory
properties. By promoting the conversion of phosphati-
dylcholine to phosphatidyl ethanol (PEth), ethanol in-
creases HDL binding to endothelial cells and modu-
lates the expression of different growth factor and cy-
tokines [205, 206]. Conversely, the effect of ethanol on
LDL-c is controversial. A significant decrease of circu-
lating LDL-c has been reported in two different cohorts
[207, 208] but not confirmed later by The Italian Lon-
gitudinal Study of Aging [209]. Genetic polymor-
phisms of Apo genes [210, 211] and changes in hepatic
cholesterol metabolism due to ethanol (likely involving
hydroxymethylglutaryl-CoA reductase [HMG-CoA]
reductase and proprotein convertase subtilisin/kexin
type 9 [PCSK9]) [212] may potentially explain these
discrepancies but further studies are required. Contrast-
ing results have been reported also for the relationship
between ethanol and TAG. Although a “U-shaped” re-
lationship has been suggested, this hypothesis has not
been yet validated [208]. As for LDL-c, genetic poly-
morphisms of Apo genes may explain inter-individual
variability in the response of TAG to ethanol [210,
213].
3.3. Ethanol and Metabolic Dysfunction
The association between ethanol consumption and
development of type 2 diabetes has been again de-
scribed as a U or J-formed curve in many observational
studies [214]. More recently, several meta-analysis
confirmed this evidence, thus emphasizing the potential
protective effect of ethanol [215, 216]. Low to moder-
ate ethanol consumption has been found to signifi-
cantly improve insulin sensitivity by different mecha-
nisms. As first, ethanol metabolism inhibits the glu-
coneogenesis, shifts the NADH/NAD-ratio, inhibits the
β-oxidation of fatty acids and inhibits glycogenolysis
[217]. Secondly, ethanol may modulate the endocrine
function of adipose tissue, likely by reducing local in-
flammation. Epidemiological studies reported an asso-
ciation between moderate alcohol consumption and
increased serum levels of adiponectin, which is thought
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 9
to improve insulin sensitivity by suppression of hepatic
glucose production, increased glucose uptake and fatty
acid oxidation in muscle tissue [218-220]. As potential
additional mechanism, a reduction of circulating CRP
was observed in subjects with light to moderate alcohol
intake, further supporting a potential anti-inflammatory
activity of ethanol [221-223]. Moreover, ethanol was
shown to dose-dependently reduce IL-6, IL-8, TNF-
alpha, and MCP-1 expression in human adipose tissue
in vitro [224].
3.4. Ethanol and Atherothrombosis
Of interest, moderate ethanol consumption was also
shown to have anti-thrombotic properties. A direct in-
hibitory effect on platelet activity has been demon-
strated ex vivo by comparing life-long abstainers and
low moderate-wine drinkers [225]. In addition, moder-
ate ethanol intake has been associated with low plasma
fibrinogen [226] and reduced plasma viscosity [227],
whereas the effect on fibrinolytic activity remains con-
troversial [228]. In line with experimental evidence,
epidemiological studies reported a reduced risk of
ischemic stroke and venous thrombosis in moderate
drinkers as compared with abstainers [229, 230].
3.5. Ethanol and Ischemic/Reperfusion Injury
Finally, the protective effect of ethanol has been
demonstrated in ischemic/reperfusion (I/R) injury [182].
On the one hand, ethanol intake increases the tolerance
of cardiomyocytes by activating cell survival pathways.
Different experimental model of ischemic pre-condi-
tioning emphasized the role protein kinase C, mito-
chondrial KATP channel, and Akt [231-233], resulting
in the de novo synthesis of protective proteins such as
COX-2, heat shock proteins, inducible nitric oxide syn-
thase, and aldehyde dehydrogenase [234, 235]. Fur-
thermore, the up-regulation of aldehyde dehydrogen-
ase-2 induced by ethanol reduced the generation of cy-
totoxic aldehydes such as 4-hydroxynonenal in heart
undergoing I/R injury [236]. On the other hand, ethanol
may prevent microvascular dysfunction after I/R injury
by abrogating P-selectin expression [237], leukocyte-
platelet/endothelial cell interactions [238, 239], ROS
generation [240], capillary no-reflow, mitochondrial
dysfunction [233, 241], release of pro-inflammatory
mediators, and microvascular barrier disruption [242].
4. RED WINE AND CARDIOVASCULAR RISK:
EVIDENCE FROM CLINICAL STUDIES
In the last decades, a large amount of epidemiologi-
cal studies from different countries showed lower CV
mortality associated with alcohol drinking (including
red-wine and partially beer). This effect has been as-
cribed to the high content of non-alcoholic phenolic
compounds with antioxidant and antithrombotic prop-
erties. Partially supporting this hypothesis, a meta-
analysis of 13 studies including 209,418 subjects
showed a reduced relative risk of CV disease associ-
ated with red/white wine intake as compared with ab-
stainers (RR 0.68 [95% CI 0.59-0.77]) [243]. In the
same study, reduced RR (0.78 [95% CI 0.70-0.86]) was
also associated with moderate beer consumption. How-
ever, a clear inverse relationship between the amount
of beer intake and vascular risk was not suggested. Co-
stanzo et al., who also reported a J-shaped relationship
between different amounts of wine intake [244], ob-
served the same results. However, a significantly lower
risk of CV disease has been also reported for beer and
spirits in other meta-analyses [245-247]. Of interest,
drinking pattern characterized by regular daily light
drinking has been also observed to have a substantial
protective role in diabetes [248], coronary artery dis-
ease [249], hypertension [250] and overall mortality
[251]. On the other hand, several epidemiological
pieces of evidence suggested that ethanol rather poly-
phenols might be the primary mechanism involved in
cardioprotection (Table 2) [252-265].
4.1. Red Wine and Coronary Heart Disease
In this context, the potential association between red
wine and CAD remains high controversial, also consid-
ering that no randomized controlled trials have been so
far performed. With this limitation, different clinical
studies suggested a beneficial effect of moderate red
wine intake on vascular function [252-255], whereas
only a small study failed to demonstrate a vaso-
protection induced by red wine intake [256] (Table 2).
However, those small clinical studies failed to clarify
the primary mechanisms involved in vascular protec-
tion. The lack of difference observed between red and
white wine administration, as reported by Whelan and
coll., might suggest a pivotal role of ethanol in vascular
prevention [252]. Conversely, Lekakis and coworkers
later emphasized the role of red grape polyphenol ex-
tract in increasing FMD [253], whereas Karatzi et al.
reported no difference between the administrations of
regular or dealcoholized red wine [254].
4.2. Red Wine and Metabolic Diseases
Some evidences supports a role of red wine in pre-
venting metabolic diseases (Table 2). As emerged from
PREDIMED trial, moderate red wine intake, assessed
by food-frequency questionnaire was associated with
reduced prevalence of metabolic syndrome and lower
10 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
Table 2. Randomized clinical trials investigating the effects of red wine consumption in cardiovascular prevention.
Author
Year
Sample Size
Treatment
Outcome
Coronary artery disease
Whelan et al.
[252]
2004
RCT
14 CAD patients
Red or white wine (4
mL/kg) with the meal
over a period of 30 min.
After 6 hours, FMD of the brachial artery improved
nearly threefold in both groups (p<0.01). No difference
was observed between the two treatment and no detect-
able change in plasma polyphenol levels after either
wine.
Lekakis et al.
[253]
2005
RCT
30 males with
CAD
Red grape polyphenol
extract (600 mg) di s-
solved in 20 ml of w ater
or placebo (water 20 ml).
Intake of the red grape polyphenol ex tract increased
FMD as compared to baseline and placebo (p<0.01 for
both).
Karatzi et al.
[254]
2005
cross-over trial
15 patients CAD
Regular or dealcoholized
red wine (250 ml).
Both regular and dealcoholized red wine decreased AIx
by 10.5% (p<0.01) and 6.1% (p=0.01), respectively.
Also a decrease in peripheral and central diastolic BP
was observed for both treatments (p=0.03 for both),
whereas peripheral systolic BP remained unaltered in
both groups.
Guarda et al.
[255]
2005
RCT
20 patients
treated with PCI
Red-wine (250 ml daily)
or abstinence from alco-
holic beverages.
The endothelium-dependent/independent dilatation ratio
significantly improved in both groups as compared to
baseline. In addition, red wine increase antioxidant activ-
ity as compared to control group (p<0.03).
Marinaccio et al.
[256]
2008
RCT
45 men with
stable CAD
Red wine (180 cc), or gin
(60 cc) or placebo (water
120 cc).
All treatment failed to improve IPC, as expressed by the
warm-up phenomenon on exercise stress testing
Metabolic syndrome
Tresserra-Rimbau
et al.
PREDIMED trial
[257]
2009
Cross-sectional
5801 high CV
subjects
Red wine intake was
recorded using a vali-
dated 137-item FFQ.
Compared with abstainers, moderate red wine drinkers
(1 drink/day) had redu ced MetS prevalence (OR 0.56
[95% CI 0 .45-0.68]; p<0.01), and lower waist circum-
ference (OR 0.59 [95% CI 0.46-0.77]; p<0.01), associ-
ated with increased HDL-c, low BP and low fasting
plasma glucose concentrations.
Chiva-Blanch et
al.
[258]
2013
Cross-over trial
67 patients at
high CV risk
Red wine (30 g alco-
hol/d), equivalent
amount of dealcoholized
red wine, and gin (30 g
alcohol/d) for 4 weeks in
a randomized order.
Red wine and dealcoholized red wine decreased mean
adjusted plasma insulin and HOMA-IR. Furthermore,
HDL cholesterol, ApoA-I, and A-II increased after red
wine and gin, whereas Lp(a) decreased after the red wine
administration
Gepner et al.
[259]
2015
RCT
224 patients with
well-controlled
T2DM
Red wine, white wine or
water with dinner for 2
years, associated with
MD.
Red wine significantly increased HDL-c (p<0.01) and
apoA-I (p=0.05) whereas decreased the total-c/HDL-c
ratio (p=0.04). In slow ethanol metabolizer, both red and
white improved glycemic control. Overall, red wine
significantly reduced the number of components of the
metabolic syndrome (p=0.05).
Myocardial infarction
Fernández-Jarne et
al.
[260]
2003
Case-control
study
171 treated sub-
jects vs. 171
controls
Alcohol intake assess-
ment during the previous
year (red win e, sweet
wine, other kinds of
wine, wine during meals,
beer, and spirits),
The risk of AMI was inversely associated with alcohol
intake. This was observed for both wine (OR 0.45 [95%
CI 0.22-0.93]) and other alcoholic beverages (OR 0.46
[95% CI 0 .22-0.94]).
(Table 2) contd….
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 11
Author
Year
Sample Size
Treatment
Outcome
Marfella et al.
[261]
2006
RCT
115 diabetic pa-
tients with AMI
Red wine
one glass (118-ml or 4-oz
[alcohol: 11.0 g])
After 1 year, red wine administration reduced LVM
and improved the synchrony between right and left
ventricle (p<0.05 for all). These findings were associ-
ated with reduced serum level of ROS (nitrotyrosine)
and inf lammatory biomarkers (CRP. TNF-α, IL-6).
Oliveira et al.
[262]
2010
Case-control study
820 patients with
AMI vs. 2196
healthy subjects
Adherence to SEAD
Adherence to SEAD was associated with 10% reduced
the risk of AMI (OR 0.90 [95% CI 0.85-0.96]).
Levantesi et al.
GISSI-Prevenzione
Trial
[263]
2013
RCT
11,323 patients
with recent AMI
Red wine up to 0.5 and
>0.5 to 1 l/day.
Wine intake was associated with lower mortality risk.
Consumption up to 0.5 L/day had an adjusted HR of
0.83 (95% CI 0.74-0.92), whereas an intake >0.5 to 1
l/day had an HR of 0.77 (95% CI 0.63-0.94). As com-
pared with non-drinkers, the p-value for trend was
<0.01.
Cosmi et al.
GISSI-Prevenzione
Trial
[264]
2013
RCT
6,973 patients
with recent AMI
Abstainers and occa-
sional drinkers, low to
moderate (1-2
glasses/day) and high (3
glasses/day) consump-
tion.
Moderate wine consumption was positively associated
with better NYHA class, whereas the incidence of
hospitalization and diabetes were lower.
Hernandez-
Hernandez et al.
[265]
2015
Prospective cohort
study
14,651 subjects
Conformity with the
MADP*
Better conformity with the MADP showed a trend
toward a reduction of CVD (AMI, IS CV death), espe-
cially CV death.
RCT: randomized clinical trial; CAD: coronary artery disease FMD: flow-mediated dilatation; Aix:!augmentation index; BP: blood pressure; PCI: percutaneous
coronary intervention; IPC: ischemic preconditioning; FFQ: food-frequency questionnaire; OR: odds ratio; CI: confidence interval; HDL-c: high-density lipo-
protein cholesterol; HOMA-IR: homeostasis model assessment for insulin resistance; APOA-I: apolipoprotein A-1; Lp(a): lipoprotein a; T2DM: type 2 diabetes
mellitus; MD: Mediterranean diet; AMI: acute myocardial infarction; LVM: left ventricular mass; ROS: reactive oxygen species; CRP: C-reactive protein;
TNF: tumor necrosis factor; IL: interleukin; SEAD: So uthern Euro pean Atlantic Diet; HR: hazard ratio; NYHA: New York Health Association; MADP: Medi-
terranean alcohol-drinking pattern; CVD: cardiovascular disease; IS: ischemic stroke; CV: cardiovascular.
*MADP: 0-to-9-points score used to identify the conformity to the traditional MADP. The score sum: (1) moderate total alcohol intake; (2) alcohol intake
spread out over the week; (3) preference for wine; (4) low consumption of spirits; (5) preference for red wine over other types of wines; (6) avoidance of excess
drinking occasions; (7) consuming win e prefe rably durin g meals.
waist circumference [257]. As additional finding, red
wine was also shown to improve glycemic control, as
later confirmed in other studies [258, 259].
4.3. Red Wine, Acute Myocardial Infarction and
Heart Failure
Finally, some interesting result has been observed in
secondary prevention of CV diseases (Table 2). Lower
risk of AMI was reported after moderate intake of red
wine or other alcoholic beverages, especially when as-
sociated with Mediterranean diet. A similar result was
also observed in secondary prevention of HF, assessed
by echocardiographic parameters or NYHA class, in
association with low circulating levels of pro-
inflammatory cytokines (CRP, TNF-α, IL-6). In addi-
tion, moderate red wine intake was shown to prevent
CV mortality.he concluding lines of the article may be
presented in a short section of conclusion.
CONCLUSION
Whereas heavy and binge drinking have been related
to increased mortality and morbidity, low-to-moderate
red wine consumption has shown beneficial effects on
metabolic and CV system. However, this intriguing
evidence is mainly based on epidemiological studies
and many unsolved questions remain. A first issue con-
cerns the need to better identify chemical compounds
conferring CV protection. This may explain why bene-
ficial effects of red over white wine have not been
clearly demonstrated, although white wine, which is
usually fermented without skin and seed, is missing
many of the polyphenols. Furthermore, it should con-
sider the grape type and the growing area/conditions as
a determinant of the chemical composition of wine, as
well as a potential synergistic effect with another
component of Mediterranean diet. Secondly, available
studies are frequently biased by methodological criti-
12 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
cisms. Standardization of red wine consumption could
be very difficult considering different drinking pat-
terns. In addition, others behavioral factors should be
considered: smoking, unhealthy diet and lack of physi-
cal exercise are more likely to be associated with heavy
drinking as compared to abstainers or moderate drink-
ers. It is expected that future studies will focus on bio-
chemical and pharmacological properties of red wine,
also including genetic research able to identified spe-
cific ethnic-age-gender groups that might potentially
benefit from low-moderate red wine consumption.
Meanwhile, given the global burden of alcohol-related
diseases, healthcare systems should maintain a cautious
attitude toward alcohol consumption. Furthermore, phy-
sicians should screen for excessive alcohol intake and
avoid encouraging alcohol consumption in abstainers.
CONSENT FOR PUBLICATION
Not applicable.
CONFLICT OF INTEREST
The authors declare no conflict of interest, financial
or otherwise.
ACKNOWLEDGEMENTS
FM, FD, and FC conceived the issue and the struc-
ture of the manuscript and supervised the overall work.
LL wrote paragraphs on flavonoid polyphenols. AB
wrote the part dealing with resveratrol. FC accounted
for the part dealing with ethanol and for conclusions.
This study was supported by the European Commis-
sion (FP7-INNOVATION I HEALTH-F2-2013-602114;
Athero-B-Cell: Targeting and exploiting B cell func-
tion for treatment of cardiovascular disease). This work
was supported by a Swiss National Science Foundation
Grant to Dr. F. Montecucco (#310030_152639/1).
REFERENCES
[1] Renaud, S.; de Lorgeril, M., Wine, alcohol, platelets, and
the French paradox for coronary heart disease. Lancet,
1992, 339, (8808), 1523-1526.
[2] Criqui, M.H.; Ringel, B.L., Does diet or alcohol explain the
French paradox? Lancet, 1994, 344, (8939-8940), 1719-1723.
[3] de Gaetano, G.; Cerletti, C.; European project, F.C.T.P.p.,
Wine and cardiovascular disease. Nutr Metab Cardiovasc
Dis, 2001, 11, (4 Suppl), 47-50.
[4] Dell'Agli, M.; Busciala, A.; Bosisio, E., Vascular effects of
wine polyphenols. Cardiovasc Res, 2004, 63, (4), 593-602.
[5] Shen, J.; Wilmot, K.A.; Ghasemzadeh, N.; Molloy, D.L.;
Burkman, G.; Mekonnen, G.; Gongora, M.C.; Quyyumi,
A.A.; Sperling, L.S., Mediterranean Dietary Patterns and
Cardiovascular Health. Annu Rev Nutr, 2015, 35, 425-449.
[6] European Association for Cardiovascular, P.; Rehabilita-
tion; Reiner, Z.; Catap ano, A.L.; De Backer, G.; Graham, I.;
Taskinen, M.R.; Wiklund, O.; Agewall, S.; Alegria, E.;
Chapman, M.J.; Durrington, P.; Erdine, S.; Halcox, J.;
Hobbs, R.; Kjekshus, J.; Filardi, P.P.; Riccardi, G.; Storey,
R.F.; Wood, D.; Guidelines, E.S.C.C.f.P.; Committees,
ESC/EAS Guidelines for the management of dyslipidae-
mias: the Task Force for the management of dyslipidaemias
of the European Society of Cardiology (ESC) and the Euro-
pean Atherosclerosis Society (EAS). Eur Heart J, 2011, 32,
(14), 1769-1818.
[7] Perk, J.; De Backer, G.; Gohlke, H.; Graham, I.; Reiner, Z.;
Verschuren, M.; Albus, C.; Benlian, P.; Boysen, G.; Cifk-
ova, R.; Deaton, C.; Ebrahim, S.; Fisher, M.; Germano, G.;
Hobbs, R.; Hoes, A.; Karadeniz, S.; Mezzani, A.; Prescott,
E.; Ryden, L.; Scherer, M.; Syvanne, M.; Scholte op Re-
imer, W.J.; Vrints, C.; Wood, D.; Zamorano, J.L.; Zannad,
F.; European Association for Cardiovascular, P.; Rehabilita-
tion; Guidelines, E.S.C.C.f.P., European Guidelines on car-
diovascular disease prevention in clinical practice (version
2012). The Fifth Joint Task Force of the European Society
of Cardiology and Other Societies on Cardiovascular Dis-
ease Prevention in Clinical Practice (constituted by repre-
sentatives of nine societies and by invited experts). Eur
Heart J, 2012, 33, (13), 1635-1701.
[8] WHO Global Status Report on Alcohol and Health 2014.
http://www.who.int/substance_abuse/publications/global_al
cohol_report/en/ (accessed 18 July, 2016),
[9] Wu, X.; Beecher, G.R.; Holden, J.M.; Haytowitz, D.B.;
Gebhardt, S.E.; Prior, R.L., Concentrations of anthocyanins
in common foods in the United States and estimation of
normal consumption. J Agric Food Chem, 2006, 54, (11),
4069-4075.
[10] Lin, J.; Rexrode, K.M.; Hu, F.; Albert, C.M.; Chae, C.U.;
Rimm, E.B.; Stampfer, M.J.; Manson, J.E., Dietary intakes
of flavonols and flavones and coronary heart disease in US
women. Am J Epidemiol, 2007, 165, (11), 1305-1313.
[11] Manach, C.; Scalbert, A.; Morand, C.; Remesy, C.; Jime-
nez, L., Polyphenols: food sources and bioavailability. Am J
Clin Nutr, 2004, 79, (5), 727-747.
[12] Duthie, G.G.; Pedersen, M.W.; Gardner, P.T.; Morrice,
P.C.; Jenkinson, A.M.; McPhail, D.B.; Steele, G.M., The
effect of whisky and wine consumption on total phenol con-
tent and antioxidant capacity of plasma from healthy volun-
teers. Eur J Clin Nutr, 1998, 52, (10), 733-736.
[13] Hollman, P.C.; Katan, M.B., Absorption, metabolism and
health effects of dietary flavonoids in man. Biomed Phar-
macother, 1997, 51, (8), 305-310.
[14] Bell, J.R.; Donovan, J.L.; Wong, R.; Waterhouse, A.L.;
German, J.B.; Walzem, R.L.; Kasim-Karakas, S.E., (+)-
Catechin in human plasma after ingestion of a single serv-
ing of reconstituted red wine. Am J Clin Nutr, 2000, 71, (1),
103-108.
[15] D'Archivio, M.; Filesi, C.; Di Benedetto, R.; Gargiulo, R.;
Giovannini, C.; Masella, R., Polyphenols, dietary sources
and bioavailability. Ann Ist Super Sanita, 2007, 43, (4),
348-361.
[16] Day, A.J.; Williamson, G., Biomarkers for exposure to die-
tary flavonoids: a review of the current evidence for identi-
fication of quercetin glycosides in plasma. Br J Nutr, 2001,
86 Suppl 1, S105-110.
[17] Notas, G.; Nifli, A.P.; Kampa, M.; Pelekanou, V.; Alexaki,
V.I.; Theodoropoulos, P.; Vercauteren, J.; Castanas, E.,
Quercetin accumulates in nuclear structures and triggers
specific gene expression in epithelial cells. J Nutr Biochem,
2012, 23, (6), 656-666.
[18] Kaindl, U.; Eyberg, I.; Rohr-Udilova, N.; Heinzle, C.;
Marian, B., The dietary antioxidants resveratrol and quer-
cetin protect cells from exogenous pro-oxidative damage.
Food Chem Toxicol, 2008, 46, (4), 1320-1326.
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 13
[19] Gorelik, S.; Ligumsky, M.; Kohen, R.; Kanner, J., A novel
function of red wine polyphenols in humans: prevention of
absorption of cytotoxic lipid peroxidation products. FASEB
J, 2008, 22, (1), 41-46.
[20] Boban, M.; Modun, D., Uric acid and antioxidant effects of
wine. Croat Med J, 2010, 51, (1), 16-22.
[21] Lu, J.; Wu, D.M.; Zheng, Y.L.; Hu, B.; Zhang, Z.F.; Shan,
Q.; Zheng, Z.H.; Liu, C.M.; Wang, Y.J., Quercetin activates
AMP-activated protein kinase by reducing PP2C expression
protecting old mouse brain against high cholesterol-induced
neurotoxicity. J Pathol, 2010, 222, (2), 199-212.
[22] Ahn, J.; Lee, H.; Kim, S.; Park, J.; Ha, T., The anti-obesity
effect of quercetin is mediated by the AMPK and MAPK
signaling pathways. Biochem Biophys Res Commun, 2008,
373, (4), 545-549.
[23] Eseberri, I.; Miranda, J.; Lasa, A.; Churruca, I.; Portillo,
M.P., Doses of Quercetin in the Range of Serum Concentra-
tions Exert Delipidating Effects in 3T3-L1 Preadipocytes by
Acting on Different Stages of Adipogenesis, but Not in Ma-
ture Adipocytes. Oxid Med Cell Longev, 2015, 2015,
480943.
[24] Davis, J.M.; Murphy, E.A.; Carmichael, M.D.; Davis, B.,
Quercetin increases brain and muscle mitochondrial bio-
genesis and exercise tolerance. Am J Physiol Regul Integr
Comp Physiol, 2009, 296, (4), R1071-1077.
[25] Nieman, D.C.; Williams, A.S.; Shanely, R.A.; Jin, F.;
McAnulty, S.R.; Triplett, N.T.; Austin, M.D.; Henson,
D.A., Quercetin's influence on exercise performan ce and
muscle mitochondrial biogenesis. Med Sci Sports Exerc,
2010, 42, (2), 338-345.
[26] Hung, C.H.; Chan, S.H.; Chu, P.M.; Tsai, K.L., Quercetin is
a potent anti-atherosclerotic compound by activation of
SIRT1 signaling under oxLDL stimulation. Mol Nutr Food
Res, 2015, 59, (10), 1905-1917.
[27] Dong, J.; Zhang, X.; Zhang, L.; Bian, H.X.; Xu, N.; Bao,
B.; Liu, J., Quercetin reduces obesity-associated ATM infil-
tration and inflammation in mice: a mechanism including
AMPKalpha1/SIRT1. J Lipid Res, 2014, 55, (3 ), 363-374.
[28] Virgili, F.; Acconcia, F.; Ambra, R.; Rinna, A.; Totta, P.;
Marino, M., Nutritional flavonoids modulate estrogen re-
ceptor alpha signaling. IUBMB Life, 2004, 56, (3), 145-151.
[29] Anter, E.; Chen, K.; Shapira, O.M.; Karas, R.H.; Keaney,
J.F., Jr., p38 mitogen-activated protein kinase activates
eNOS in endothelial cells by an estrogen recepto r alpha-
dependent pathway in response to black tea polyphenols.
Circ Res, 2005, 96, (10), 1072-1078.
[30] Chalopin, M.; Tesse, A.; Martinez, M.C.; Rognan, D.; Ar-
nal, J.F.; Andriantsitohaina, R., Estrogen receptor alpha as a
key target of red wine polyphenols action on the endothe-
lium. PLoS One, 2010, 5, (1), e8554.
[31] Simoncini, T.; Lenzi, E.; Zochling, A.; Gopal, S.; Goglia,
L.; Russo, E.; Polak, K.; Casarosa, E.; Jungbauer, A.;
Genazzani, A.D.; Genazzani, A.R., Estrogen-like effects of
wine extracts on nitric oxide synthesis in human endothelial
cells. Maturitas, 2011, 70, (2), 169-175.
[32] Kane, M.O.; Anselm, E.; Rattmann, Y.D.; Auger, C.;
Schini-Kerth, V.B., Role of gender and estrogen receptors
in the rat aorta endothelium -dependent relaxation to red wine
polyphenols. Vascul Pharmacol, 2009, 51, (2-3), 140-146.
[33] Chalopin, M.; Soleti, R.; Benameur, T.; Tesse, A.; Faure,
S.; Martinez, M.C.; Andriantsitohaina, R., Red wine poly-
phenol compounds favor neovascularisation through estro-
gen receptor alpha-independent mechanism in mice. PLoS
One, 2014, 9, (10), e110080.
[34] Duarte, J.; Andriambeloson, E.; Diebolt, M.; Andriantsito-
haina, R., Wine polyphenols stimulate superoxide anion
production to promote calcium signaling and endothelial-
dependent vasodilatation. Physiol Res, 2004, 53, (6), 595-
602.
[35] Andriambeloson, E.; Kleschyov, A.L.; Muller, B.; Beretz,
A.; Stoclet, J.C.; Andriantsitohaina, R., Nitric oxide produc-
tion and endothelium-dependent vasorelaxation induced by
wine polyphenols in rat aorta. Br J Pharmacol, 1997, 120,
(6), 1053-1058.
[36] de Moura, R.S.; Miranda, D.Z.; Pinto, A.C.; Sicca, R.F.;
Souza, M.A.; Rubenich, L.M.; Carvalho, L.C.; Rangel,
B.M.; Tano, T.; Madeira, S.V.; Resende, A.C., Mechanism
of the endothelium-dependent vasodilation and the anti-
hypertensive effect of Brazilian red wine. J Cardiovasc
Pharmacol, 2004, 44, (3), 302-309.
[37] Ndiaye, M.; Chataigneau, M.; Lobysheva, I.; Chataigneau,
T.; Schini-Kerth, V.B., Red wine polyphenol-induced, en-
dothelium-dependent NO-mediated relaxation is due to the
redox-sensitive PI3-kinase/Akt-dependent phosphorylation
of endothelial NO-synthase in the isolated porcine coronary
artery. FASEB J, 2005, 19, (3), 455-457.
[38] Anselm, E.; Chataigneau, M.; Ndiaye, M.; Chataigneau, T.;
Schini-Kerth, V.B., Grape juice causes endothelium-
dependent relaxation via a redox-sensitive Src- and Akt-
dependent activation of eNOS. Cardiovasc Res, 2007, 73,
(2), 404-413.
[39] Sarr, M.; Chataigneau, M.; Martins, S.; Schott, C.; El Bed-
oui, J.; Oak, M.H.; Muller, B.; Chataigneau, T.; Schini-
Kerth, V.B., Red wine polyphenols prevent angiotensin II-
induced hypertension and endothelial dysfunction in rats:
role of NADPH oxidase. Cardiovasc Res, 2006, 71, (4),
794-802.
[40] Bernatova, I.; Pechanova, O.; Babal, P.; Kysela, S.;
Stvrtina, S.; Andriantsitohaina, R., Wine polyphenols im-
prove cardiovascular remodeling and vascular function in
NO-deficient hypertension. Am J Physiol Heart Circ
Physiol, 2002, 282, (3), H942-948.
[41] Peng, N.; Clark, J.T.; Prasain, J.; Kim, H.; White, C.R.;
Wyss, J.M., Antihypertensive and cognitive effects of grape
polyphenols in estrogen-depleted, female, spontaneously
hypertensive rats. Am J Physiol Regul Integr Comp Physiol,
2005, 289, (3), R771-775.
[42] Jimenez, R.; Lopez-Sepulveda, R.; Kadmiri, M.; Romero,
M.; Vera, R.; Sanchez, M.; Vargas, F.; O'Valle, F.;
Zarzuelo, A.; Duenas, M.; Santos-Buelga, C.; Duarte, J.,
Polyphenols restore endothelial function in DOCA-salt hy-
pertension: role of endothelin-1 and NADPH oxidase. Free
Radic Biol Med, 2007, 43, (3), 462-473.
[43] Auger, C.; Gerain, P.; Laurent-Bichon, F.; Portet, K.; Bor-
net, A.; Caporiccio, B.; Cros, G.; Teissedre, P.L.; Rouanet,
J.M., Phenolics from commercialized grape extracts prevent
early atherosclerotic lesions in hamsters by mechanisms
other than antioxidant effect. J Agric Food Chem, 2004, 52,
(16), 5297-5302.
[44] Shanmuganayagam, D.; Warner, T.F.; Krueger, C.G.; Reed,
J.D.; Folts, J.D., Concord grape juice attenuates platelet ag-
gregation, serum cholesterol and development of atheroma
in hypercholesterolemic rabbits. Atherosclerosis, 2007, 190,
(1), 135-142.
[45] Pignatelli, P.; Pulcinelli, F.M.; Celestini, A.; Lenti, L.;
Ghiselli, A.; Gazzaniga, P.P.; Violi, F., The flavonoids
quercetin and catechin synergistically inhibit platelet func-
tion by antagonizing the in tracellular production of hydro-
gen peroxide. Am J Clin Nutr, 2000, 72, (5), 1150-1155.
[46] Freedman, J.E.; Parker, C., 3rd; Li, L.; Perlman, J.A.; Frei,
B.; Ivanov, V.; Deak, L.R.; Iafrati, M.D.; Folts, J.D., Select
flavonoids and whole juice from purple grapes inhibit plate-
let function and enhance nitric oxide release. Circulation,
2001, 103, (23), 2792-2798.
14 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
[47] Al-Awwadi, N.A.; Araiz, C.; Bornet, A.; Delbosc, S.; Cris-
tol, J.P.; Linck, N.; A zay, J.; Teissedre, P.L.; Cros, G., Ex-
tracts enriched in different polyphenolic families normalize
increased card iac NADPH oxid ase expression while having
differential effects on insulin resistance, hypertension, and
cardiac hypertrophy in high-fructose-fed rats. J Agric Food
Chem, 2005, 53, (1), 151-157.
[48] Ihm, S.H.; Lee, J.O.; Kim, S.J.; Seung, K.B.; Schini-Kerth,
V.B.; Chang, K.; Oak, M.H., Catechin prevents endothelial
dysfunction in the prediabetic stage of OLETF rats by re-
ducing vascular NADPH oxidase activity and expression.
Atherosclerosis, 2009, 206, (1), 47-53.
[49] Agouni, A.; Lagrue-Lak-Hal, A.H.; Mostefai, H.A.; Tesse,
A.; Mulder, P.; Rouet, P.; Desmoulin, F.; Heymes, C.;
Martinez, M.C.; Andriantsitohaina, R., Red wine polyphe-
nols prevent metabolic and cardiovascular alterations asso-
ciated with obesity in Zucker fatty rats (Fa/Fa). PLoS One,
2009, 4, (5), e5557.
[50] Terra, X.; Montagut, G.; Bustos, M.; Llopiz, N.; Ardevol,
A.; Blade, C.; Fernandez-Larrea, J.; Pujadas, G.; Salvado,
J.; Arola, L.; Blay, M., Grape-seed procyanidins prevent
low-grade inflammation by modulating cytokine expression
in rats fed a high-fat diet. J Nutr Biochem, 2009, 20, (3),
210-218.
[51] Rangel-Huerta, O.D.; Pastor-Villaescusa, B.; Aguilera,
C.M.; Gil, A., A Systematic Review of the Efficacy of Bio-
active Compounds in Cardiovascular Disease: Phenolic
Compounds. Nutrients, 2015, 7, (7), 5177-5216.
[52] Agewall, S.; Wright, S.; Doughty, R.N.; Whalley, G.A.;
Duxbury, M.; Sharpe, N., Does a glass of red wine improve
endothelial function ? Eur Heart J, 2000, 21, (1), 74-78.
[53] Hashimoto, M.; Kim, S.; Eto, M.; Iijima, K.; Ako, J.; Yo-
shizumi, M.; Akishita, M.; Kondo, K.; Itakura, H.; Hosoda,
K.; Toba, K.; Ouchi, Y., Effect of acute intake of red wine
on flow-mediated vasodilatation of the brachial artery. Am J
Cardiol, 2001, 88, (12), 1457-1460, A1459.
[54] Boban, M.; Modun, D.; Music, I.; Vukovic, J.; Brizic, I.;
Salamunic, I.; Obad, A.; Palada, I.; Dujic, Z., Red wine in-
duced modulation of vascular function: separating the role
of polyphenols, ethanol, and urates. J Cardiovasc Pharma-
col, 2006, 47, (5), 695-701.
[55] Karatzi, K.; Papamichael, C.; Karatzis, E.; Papaioannou,
T.G.; Voidonikola, P.T.; Vamvakou, G.D.; Lekakis, J.;
Zampelas, A., Postprandial improvement of endothelial
function by red wine and olive oil antioxidants: a synergis-
tic effect of components of the Mediterranean diet. J Am
Coll Nutr, 2008, 27, (4), 448-453.
[56] Engler, M.B.; Engler, M.M.; Chen, C.Y.; Malloy, M.J.;
Browne, A.; Chiu, E.Y.; Kwak, H.K.; Milbury, P.; Paul,
S.M.; Blumberg, J.; Mietus-Snyder, M.L., Flavonoid-rich
dark chocolate improves endothelial function and increases
plasma epicatechin concentrations in healthy adults. J Am
Coll Nutr, 2004, 23, (3), 197-204.
[57] Clifton, P.M., Effect of Grape Seed Extract and Quercetin
on Cardiovascular and Endothelial Parameters in High-Risk
Subjects. J Biomed Biotechnol, 2004, 2004, (5), 272-278.
[58] Schroeter, H.; Heiss, C.; Balzer, J.; Kleinbongard, P.; Keen,
C.L.; Hollenberg, N.K.; Sies, H.; Kwik-Uribe, C.; Schmitz,
H.H.; Kelm, M., (-)-Epicatechin mediates beneficial effects
of flavanol-rich coco a on vascular function in humans. Proc
Natl Acad Sci U S A, 2006, 103, (4), 1024-1029.
[59] Hampton, S.M.; Isherwood, C.; Kirkpatrick, V.J.; Lynne-
Smith, A.C.; Griffin, B.A., The influence of alcohol con-
sumed with a meal on endothelial function in healthy indi-
viduals. J Hum Nutr Diet, 2010, 23, (2), 120-125.
[60] Hall, W.L.; Formanuik, N.L.; Harnpanich, D.; Cheung, M.;
Talbot, D.; Chowienczyk, P.J.; Sanders, T.A., A meal en-
riched with soy isoflavones in creases n itric oxide-mediated
vasodilation in healthy postmenopausal women. J Nutr,
2008, 138, (7), 1288-1292.
[61] Papamichael, C.; Karatzis, E.; Karatzi, K.; Aznaouridis, K.;
Papaioannou, T.; Protogerou, A.; Stamatelopoulos, K.;
Zampelas, A.; Lekakis, J.; Mavrikakis, M., Red wine's anti-
oxidants counteract acute endothelial dysfunction caused by
cigarette smoking in healthy nonsmokers. Am Heart J,
2004, 147, (2), E5.
[62] Heiss, C.; Kleinbongard, P.; Dejam, A.; Perre, S.;
Schroeter, H.; Sies, H.; Kelm, M., Acute consumption of
flavanol-rich cocoa and the reversal of endothelial dysfunc-
tion in smokers. J Am Coll Cardiol, 2005, 46, (7), 1276-
1283.
[63] Papamichael, C.; Karatzi, K.; Karatzis, E.; Papaioannou,
T.G.; Katsichti, P.; Zampelas, A.; Lekakis, J., Combined
acute effects of red wine consumption and cigarette smok-
ing on haemodynamics of young smokers. J Hyperten s,
2006, 24, (7), 1287-1292.
[64] Coimbra, S.R.; Lage, S.H.; Brandizzi, L.; Yoshida, V.; da
Luz, P.L., The action of red wine and purple grape juice on
vascular reactivity is independent of plasma lipids in hyper-
cholesterolem ic patients. Braz J Med Biol Res, 2005, 38,
(9), 1339-1347.
[65] Faridi, Z.; Njike, V.Y.; Dutta, S.; Ali, A.; Katz, D.L., Acute
dark chocolate and cocoa ingestion and endothelial func-
tion: a randomized controlled crossover trial. Am J Clin
Nutr, 2008, 88, (1), 58-63.
[66] Balzer, J.; Rassaf, T.; Heiss, C.; Kleinbongard, P.; Lauer,
T.; Merx, M.; Heussen, N.; Gross, H.B.; Keen, C.L.;
Schroeter, H.; Kelm, M., Sustained benefits in vascular
function through flavanol-containing cocoa in medicated
diabetic patients a double-masked, randomized, controlled
trial. J Am Coll Cardiol, 2008, 51, (22), 2141-2149.
[67] Berry, N.M.; Davison, K.; Coates, A.M.; Buckley, J.D.;
Howe, P.R., Impact of cocoa flavanol consumption on
blood pressure responsiveness to exercise. Br J Nutr, 2010,
103, (10), 1480-1484.
[68] Aviram, M.; Dornfeld, L., Pomegranate juice consumption
inhibits serum angiotensin converting enzyme activity and
reduces systolic blood pressure. Atherosclerosis, 2001, 158,
(1), 195-198.
[69] Taubert, D.; Berkels, R.; Roesen, R.; Klaus, W., Chocolate
and blood pressure in elderly individuals with isolat ed sys-
tolic hypertension. JAMA, 2003, 290, (8), 1029-1030.
[70] Park, Y.K.; Kim, J.S.; Kang, M.H., Concord grape juice
supplem entation reduces blood pressure in Korean hyper-
tensive men: double-blind, placebo controlled intervention
trial. Biofactors, 2004, 22, (1-4), 145-147.
[71] Dohadwala, M.M.; Hamburg, N.M.; Holbrook, M.; Kim,
B.H.; Duess, M.A.; Levit, A.; Titas, M.; Chung, W.B.; Vin-
cent, F.B.; Caiano, T.L.; Frame, A.A .; Keaney, J.F., Jr.;
Vita, J.A., Effects of Concord grape juice on ambulatory
blood pressure in prehypertension and stage 1 hyperten sion.
Am J Clin Nutr, 2010, 92, (5), 1052-1059.
[72] Stein, J.H.; Keevil, J.G.; Wiebe, D.A.; Aeschlimann, S.;
Folts, J.D., Purple grape juice improves endothelial function
and reduces the susceptibility of LDL cholesterol to oxida-
tion in patients with coronary artery disease. Circulation,
1999, 100, (10), 1050-1055.
[73] Aviram, M.; Rosenblat, M.; Gaitini, D.; Nitecki, S.;
Hoffman, A.; Dornfeld, L.; Volkova, N.; Presser, D.; Attias,
J.; Liker, H.; Hay ek, T., Pomegran ate juice consumption for
3 years by patients with carotid artery stenosis reduces
common carotid intima-media thickness, blood pressure and
LDL oxidation. Clin Nutr, 2004, 23, (3), 423-433.
[74] Karatzi, K.; Papamichael, C.; Aznaouridis, K.; Karatzis, E.;
Lekakis, J.; Matsouka, C.; Boskou, G.; Chiou, A.; Sitara,
M.; Feliou, G.; Kontoyiannis, D.; Zampelas, A.; Mavrika-
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 15
kis, M., Constituents of red wine other than alcohol im-
prove endothelial function in patients with coronary artery
disease. Coron Artery Dis, 2004, 15, (8), 485-490.
[75] Widlansky, M.E.; Hamburg, N.M.; Anter, E.; Holbrook,
M.; Kahn, D.F.; Elliott, J.G.; Keaney, J.F., Jr.; Vita, J.A.,
Acute EGCG supplementation reverses endothelial dys-
function in patients with coronary artery disease. J Am Coll
Nutr, 2007, 26, (2), 95-102.
[76] Chou, E.J.; Keevil, J.G.; Aeschlimann, S.; Wiebe, D.A.;
Folts, J.D.; Stein, J.H., Effect of ingestion of purple grape
juice on endothelial function in patients with coronary heart
disease. Am J Cardiol, 2001, 88, (5), 553-555.
[77] Duffy, S.J.; Keaney, J.F., Jr.; Holbrook, M.; Gokce, N.;
Swerdloff, P.L.; Frei, B.; Vita, J.A., Short- and long-term
black tea consumption reverses endothelial dysfunction in
patients with coronary artery disease. Circulation, 2001,
104, (2), 151-156.
[78] Burns, J.; Yokota, T.; Ashihara, H.; Lean, M.E.; Crozier,
A., Plant foods and herbal sources of resveratrol. J Agric
Food Chem, 2002, 50, (11), 3337-3340.
[79] Das, D.K.; Maulik, N., Resveratrol in cardioprotection: a
therapeutic promise of alternative medicine. Mol Interv,
2006, 6, (1), 36-47.
[80] Goswami, S.K.; Das, D.K., Resveratrol and chemopreven-
tion. Cancer Lett, 2009, 284, (1), 1-6.
[81] Goldberg, D.M.; Tsang, E.; Karumanchiri, A.; Diamandis,
E.; Soleas, G.; Ng, E., Method to assay the concentrations
of phenolic constituents of biological interest in wines. Anal
Chem, 1996, 68, (10), 1688-1694.
[82] Pervaiz, S., Resveratrol: from grapevines to mammalian
biology. FASEB J, 2003, 17, (14), 1975-1985.
[83] Fornara, V.; Onelli, E.; Sparvoli, F.; Rossoni, M.; Aina, R.;
Marino, G.; Citterio, S., Localization of stilbene synthase in
Vitis vinifera L. during berry development. Protoplasma,
2008, 233, (1-2), 83-93.
[84] Wang, W.; Tang, K.; Yang, H.R.; Wen, P.F.; Zhang, P.;
Wang, H.L.; Huang, W.D., Distribution of resveratrol and
stilbene synthase in young grape plants (Vitis vin ifera L. cv.
Cabernet Sauvignon) and the effect of UV-C on its accumu-
lation. Plant Physiol Biochem, 2010, 48, (2-3), 142-152.
[85] Soleas, G.J.; Diamandis, E.P.; Goldberg, D.M., Resveratrol:
a molecule whose time has come? And gone? Clin Bio-
chem, 1997, 30, (2 ), 91-113.
[86] Kundu, J.K.; Surh, Y.J., Cancer ch emoprevent ive and
therapeutic potential of resveratrol: mechanistic perspec-
tives. Cancer Lett, 2008, 269, (2), 243-261.
[87] Bertelli, A.A.; Giovannini, L.; Bernini, W.; Migliori, M.;
Fregoni, M.; Bavaresco, L.; Bertelli, A., Antiplatelet acti-
vity of cis-resveratrol. Drugs Exp Clin Res, 1996, 22, (2),
61-63.
[88] Adrian, M.; Jeandet, P.; Veneau, J.; Weston, L.A.; Bessis,
R., Biological Activity of Resveratrol, a Stilbenic Com-
pound from Grapevines, Against Botrytis cinerea, the
Causal Agent for Gray Mold. Journal of Chemical Ecology,
1997, 23, (7), 1689-1702.
[89] Wong, R.H.; Howe, P.R.; Buckley, J.D.; Coates, A.M.;
Kunz, I.; Berry, N.M., Acute resveratrol supplementation
improves flow-mediated dilatation in overweight/obese in-
dividuals with mildly elevated blood pressure. Nutr Metab
Cardiovasc Dis, 2011, 21, (11), 851-856.
[90] Forstermann, U.; Li, H., Therapeutic effect of enhancing
endothelial nitric oxide synthase (eNOS) expression and
preventing eNOS uncoupling. Br J Pharmacol, 2011, 164,
(2), 213-223.
[91] Zhang, Q.J.; Wang, Z.; Chen, H.Z.; Zhou, S.; Zheng, W.;
Liu, G.; Wei, Y.S.; Cai, H.; Liu, D.P.; Liang, C.C., Endo-
thelium-specific overexpression of class III d eacetylase
SIRT1 decreases atherosclerosis in apolipoprotein E-
deficient mice. Cardiovasc Res, 2008, 80, (2), 191-199.
[92] Csiszar, A.; Labinskyy, N.; Pinto, J.T.; Ballabh, P.; Zhang,
H.; Losonczy, G.; Pearson, K.; de Cabo, R.; Pacher, P.;
Zhang, C.; Ungvari, Z., Resveratrol induces mitochondrial
biogenesis in endothelial cells. Am J Physiol Heart Circ
Physiol, 2009, 297, (1), H13-20.
[93] Klinge, C.M.; Blankenship, K.A.; Risinger, K.E.; Bh atna-
gar, S.; Noisin, E.L.; Sumanasekera, W.K.; Zhao, L.; Brey,
D.M.; Keynton, R.S., Resveratrol and estradiol rapidly acti-
vate MAPK signaling through estrogen receptors alpha and
beta in endothelial cells. J Biol Chem, 2005, 280, (9), 7460-
7468.
[94] Klinge, C.M.; Wickramasinghe, N.S.; Ivanova, M.M.;
Dougherty, S.M., Resveratrol stimulates nitric oxide pro-
duction by increasing estrogen receptor alpha-Src-caveolin-
1 interaction and phosphorylation in human umbilical vein
endothelial cells. FASEB J, 2008, 22, (7), 2185-2197.
[95] Mattagajasingh, I.; Kim, C.S.; Naqvi, A.; Yamamori, T.;
Hoffman, T.A.; Jung, S.B.; DeRicco, J.; Kasuno, K.; Irani,
K., SIRT1 promotes endothelium-dependent vascular re-
laxation by activating endothelial nitric oxide synthase.
Proc Natl Acad Sci U S A, 2007, 104, (37), 14855-14860.
[96] Frombaum, M.; Therond, P.; Djelidi, R.; Beaudeux, J.L.;
Bonnefont-Rousselot, D.; Borderie, D., Piceatannol is more
effective than resveratrol in restoring endothelial cell di-
methylarginine dimethylaminohydrolase expression and ac-
tivity after high-glucose oxidative stress. Free Radic Res,
2011, 45, (3), 293-302.
[97] Penumathsa, S.V.; Koneru, S.; Samuel, S.M.; Maulik, G.;
Bagchi, D.; Yet, S.F.; Menon, V.P.; Maulik, N., Strategic
targets to induce neovascularization by resveratrol in hyper-
cholesterolem ic rat myocardium: role of caveolin-1, endo-
thelial nitric oxide synthase, hem eoxygenase-1, and vascu-
lar endothelial growth factor. Free Radic Biol Med, 2008,
45, (7), 1027-1034.
[98] Xia, N.; Daiber, A.; Habermeier, A.; Closs, E.I.; Thum, T.;
Spanier, G.; Lu, Q.; Oelze, M.; Torzewski, M.; Lackner,
K.J.; Munzel, T.; Forstermann, U.; Li, H., Resveratrol re-
verses endothelial nitric-oxide synthase uncoupling in apol-
ipoprotein E knockout mice. J Pharmacol Exp Ther, 2010,
335, (1), 149-154.
[99] Li, H.; Xia, N.; Forstermann, U., Cardiovascular effects and
molecular targets of resveratrol. Nitric Oxide, 2012, 26, (2),
102-110.
[100] Howitz, K.T.; Bitterman, K.J.; Cohen, H.Y.; Lamming,
D.W.; Lavu, S.; Wood, J.G.; Zipkin, R.E.; Chung, P.;
Kisielewski, A.; Zhang, L.L.; Scherer, B.; Sinclair, D.A.,
Small molecule activators of sirtuins extend Saccharomyces
cerevisiae lifespan. Nature, 2003, 425, (6954), 191-196.
[101] Beher, D.; Wu, J.; Cumine, S.; Kim, K.W.; Lu, S.C.; Atan-
gan, L.; Wang, M., Resveratrol is not a direct activator of
SIRT1 enzyme activity. Chem Biol Drug Des, 2009, 74, (6),
619-624.
[102] Pacholec, M.; Bleasdale, J.E.; Chrunyk, B.; Cunningham,
D.; Flynn, D.; Garofalo, R.S.; Griffith, D.; Griffor, M.; Lou-
lakis, P.; Pabst, B.; Qiu, X.; Stockman, B.; Thanabal, V.;
Varghese, A.; Ward, J.; Withka, J.; Ahn, K., SRT1720,
SRT2183, SRT1460, and resveratrol are not direct activa-
tors of SIRT1. J Biol Chem, 2010, 285, (11), 8340-8351.
[103] Hu, Y.; Liu, J.; Wang, J.; Liu, Q., The controversial links
among calorie restriction, SIRT1, and resveratrol. Free
Radic Biol Med, 2011, 51, (2), 250-256.
[104] Dai, H.; Kustigian, L.; Carney, D.; Case, A.; Considine, T.;
Hubbard, B.P.; Perni, R.B.; Riera, T.V.; Szczepankiewicz,
B.; Vlasuk, G.P.; Stein, R.L., SIRT1 activation by small
molecules: kinetic and biophysical evidence for direct inter-
16 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
action of enzyme and activator. J Biol Chem , 2010, 285,
(43), 32695-32703.
[105] Mukherjee, S.; Lekli, I.; Gurusamy, N.; Bertelli, A.A.; Das,
D.K., Expression of the longevity proteins by both red and
white wines and their cardioprotective components, resvera-
trol, tyrosol, and hydroxytyrosol. Free Radic Biol Med,
2009, 46, (5), 573-578.
[106] Canto, C.; Gerhart-Hines, Z.; Feige, J.N.; Lagouge, M.;
Noriega, L.; Milne, J.C.; Elliott, P.J.; Puigserver, P.;
Auwerx, J., AMPK regulates energy expenditure by modu-
lating NAD+ metabolism and SIRT1 activity. Nature, 2009,
458, (7241), 1056-1060.
[107] Csiszar, A.; Smith, K.; Labinskyy, N.; Orosz, Z.; Rivera,
A.; Ungvari, Z., Resveratrol attenuates TNF-alpha-induced
activation of coronary arterial endothelial cells: role of NF-
kappaB inhibition. Am J Physiol Heart Circ Physiol, 2006,
291, (4), H1694-1699.
[108] Csiszar, A.; Labinskyy, N.; Podlutsky, A.; Kaminski, P.M.;
Wolin, M.S.; Zhang, C.; Mukhopadhyay, P.; Pacher, P.; Hu,
F.; de Cabo, R.; Ballabh, P.; Ungvari, Z., Vasoprotective ef-
fects of resveratrol and SIRT1: attenuation of cigarette
smoke-induced oxidative stress and proinflammatory phe-
notypic alterations. Am J Physiol Heart Circ Physiol, 2008,
294, (6), H2721-2735.
[109] Labinskyy, N.; Csiszar, A.; Veress, G.; Stef, G.; Pacher, P.;
Oroszi, G.; Wu, J.; Ungvari, Z., V ascular dysfunction in ag-
ing: potential effects of resveratrol, an anti-inflamm atory
phytoestrogen. Curr Med Chem, 2006, 13, (9), 989-996.
[110] Kundu, J.K.; Shin, Y.K.; Kim, S.H.; Surh, Y.J., Resveratrol
inhibits phorbol ester-induced expression of COX-2 and ac-
tivation of NF-kappaB in mouse skin by blocking IkappaB
kinase activity. Carcinogenesis, 2006, 27, (7), 1465-1474.
[111] Manna, S.K.; Mukhopadhyay, A.; Aggarwal, B.B., Res-
veratrol suppresses TNF-induced activation of nuclear tran-
scription factors NF-kappa B, activator protein-1, and apop-
tosis: potential role of reactive oxygen interm ediates and
lipid peroxidation. J Immunol, 2000, 164, (12), 6509-6519.
[112] Leonard, S.S.; Xia, C.; Jiang, B.H.; Stinefelt, B.; Klandorf,
H.; Harris, G.K.; Shi, X., Resveratrol scavenges reactive
oxygen species and effects radical-induced cellular re-
sponses. Biochem Biophys Res Commun, 2003, 309, (4),
1017-1026.
[113] Ungvari, Z.; Orosz, Z.; Rivera, A.; Labinskyy, N.; Xiang-
min, Z.; Olson, S.; Podlutsky, A.; Csiszar, A., Resveratrol
increases vascular oxidative stress resistance. Am J Physiol
Heart Circ Physiol, 2007, 292, (5), H2417-2424.
[114] Li, Y.; Cao, Z.; Zhu, H., Upregulation of endogenous anti-
oxidants and phase 2 enzymes by the red wine polyphenol,
resveratrol in cultured aortic smooth muscle cells leads to
cytoprotection against oxidative and electrophilic stress.
Pharmacol Res, 2006, 53, (1), 6-15.
[115] Spanier, G.; Xu, H.; Xia, N.; Tobias, S.; Deng, S.; Wo-
jnowski, L.; Forstermann, U.; Li, H., Resveratrol reduces
endothelial oxidative stress by modulating the gene expres-
sion of superoxide dismutase 1 (SOD1), glutathione peroxi-
dase 1 (GPx1) and NADPH oxidase subunit (Nox4). J
Physiol Pharmacol, 2009, 60 Suppl 4, 111-116.
[116] Tanno, M.; Kuno, A.; Yano, T.; Miura, T.; Hisahara, S.;
Ishikawa, S.; Shimamoto, K.; Horio, Y., Induction of man-
ganese superoxide dismutase by nuclear translocation and
activation of SIRT1 promotes cell survival in chronic hear t
failure. J Biol Chem, 2010, 285, (11), 8375-8382.
[117] Kaga, S.; Zhan, L.; Matsumoto, M.; Maulik, N., R esveratrol
enhances neovascularization in the infarcted rat myocar-
dium through the induction of thioredoxin-1, heme oxy-
genase-1 and vascular endothelial growth factor. J Mol Cell
Cardiol, 2005, 39, (5), 813-822.
[118] Juan, S.H.; Cheng, T.H.; Lin, H.C.; Chu, Y.L.; Lee, W.S.,
Mechanism of concentration-dependent induction of heme
oxygenase-1 by resveratrol in human aortic smooth muscle
cells. Biochem Pharmacol, 2005, 69, (1), 41-48.
[119] Dudley, J.; Das, S.; Mukherjee, S.; Das, D.K., Resveratrol,
a unique phytoalexin present in red win e, delivers either
survival signal or death signal to the ischemic myocardium
depending on dose. J Nutr Biochem, 2009, 20, (6), 443-452.
[120] Ungvari, Z.; Orosz, Z.; Labinskyy, N.; Rivera, A.; Xiang-
min, Z.; Smith, K.; Csiszar, A., Increased mitochondrial
H2O2 production promotes endothelial NF-kappaB activa-
tion in aged rat arteries. Am J Physiol Heart Circ Physiol,
2007, 293, (1), H37-47.
[121] Yu, H.P.; Hwang, T.L.; Hwang, T.L.; Yen, C.H.; Lau, Y.T .,
Resveratrol prevents endothelial dysfunction and aortic su-
peroxide production after trauma hemorrhage through es-
trogen receptor-dependent hemeoxygenase-1 pathway. Crit
Care Med, 2010, 38, (4), 1147-1154.
[122] Chow, S.E.; Hshu, Y.C.; Wang, J.S.; Chen, J.K., Resvera-
trol attenuates oxLDL-stimulated NADPH oxidase activity
and protects endothelial cells from oxidative functional
damages. J Appl Physiol (1985), 2007, 102, (4), 1520-1527.
[123] Shen, M.Y.; Hsiao, G.; Liu, C.L.; Fong, T.H.; Lin, K.H.;
Chou, D.S.; Sheu, J.R., Inhibitory mechanisms of resvera-
trol in platelet activation: pivotal roles of p38 MAPK and
NO/cyclic GMP. Br J Haematol, 2007, 139, (3), 475-485.
[124] Zordoky, B.N.; Robertson, I.M.; Dyck, J.R., Preclinical and
clinical eviden ce for the role of resveratrol in the treatment
of cardiovascular d iseases. Biochim Biophys Acta, 2015,
1852, (6), 1155-1177.
[125] Rivera, L.; Moron, R.; Zarzuelo, A.; Galisteo, M., Long-
term resveratrol administration reduces metabolic distur-
bances and lowers blood pressure in obese Zucker rats. Bio-
chem Pharmacol, 2009, 77, (6), 1053-1063.
[126] Dolinsky, V.W.; Chakrabarti, S.; Pereira, T.J.; Oka, T.;
Levasseur, J.; Beker, D.; Zordoky, B.N.; Morton, J.S.; Na-
gendran, J.; Lopaschuk, G.D.; Davidge, S.T.; Dyck, J.R.,
Resveratrol prevents hypertension and cardiac hypertrophy
in hypertensiv e rats and mice. Biochim Biophys Acta, 2013,
1832, (10), 1723-1733.
[127] Liu, Z.; Song, Y.; Zhang, X.; Liu, Z.; Zhang, W.; Mao, W.;
Wang, W.; Cui, W.; Zhang, X.; Jia, X.; Li, N.; Han, C.; Liu,
C., Effects of trans-resveratrol on hypertension-induced
cardiac hypertrophy using the partially nephrectomized rat
model. Clin Exp Pharmacol Physiol, 2005, 32, (12), 1049-
1054.
[128] Chan, V.; Fenning, A.; Iyer, A.; Hoey, A.; Brown, L., Res-
veratrol improves cardiovascular function in DOCA-salt
hypertensive rats. Curr Pharm Biotechnol, 2011, 12, (3),
429-436.
[129] Rimbaud, S.; Ruiz, M.; Piquereau, J.; Mateo, P.; Fortin, D.;
Veksler, V.; Garnier, A.; Ventura-Clapier, R., Resveratrol
improves survival, hemodynamics and energetics in a rat
model of hypertension leading to heart failure. PLoS One,
2011, 6, (10), e26391.
[130] Liu, Y.; Ma, W.; Zhang, P.; He, S.; Huang, D., Effect of
resveratrol on blood pressure: a meta-analysis of random-
ized controlled trials. Clin Nutr, 2015, 34, (1), 27-34.
[131] Cao, X.; Luo, T.; Luo, X.; Tang, Z., Resveratrol prevents
AngII-induced hypertension via AMPK activation and
RhoA/ROCK suppression in mice. Hypertens Res, 2014,
37, (9), 803-810.
[132] Wang, Z.; Zou, J.; Cao, K.; Hsieh, T.C.; Huang, Y.; Wu,
J.M., Dealcoholized red wine containing known amounts of
resveratrol suppresses atherosclerosis in hypercholes-
terolemic rabbits without affecting plasma lipid levels. Int J
Mol Med, 2005, 16, (4), 533-540.
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 17
[133] Gocmen, A.Y.; Burgucu, D.; Gumuslu, S., Effect of res-
veratrol on platelet activation in hypercholesterolemic rats:
CD40-CD40L system as a potential target. Appl Physiol
Nutr Metab, 2011, 36, (3), 323-330.
[134] Yashiro, T.; Nanmoku, M.; Shimizu, M.; Inoue, J.; S ato, R.,
Resveratrol increases the expression and activity of the low
density lipoprotein receptor in hepatocytes by the prote-
olytic activation of the sterol regulatory element-binding
proteins. Atherosclerosis, 2012, 220, (2), 369-374.
[135] Sahebkar, A., Effects of resveratrol supplementation on
plasma lipids: a systematic review and meta-analysis of
randomized controlled trials. Nutr Rev, 2013, 71, (12), 822-
835.
[136] Bhatt, J.K.; Thomas, S.; Nanjan, M.J., Resveratrol supple-
mentation improves glycemic control in type 2 diabetes
mellitus. Nutr Res, 2012, 32, (7), 537-541.
[137] Timmers, S.; Konings, E.; Bilet, L.; Houtkooper, R.H.; van
de Weijer, T.; Goossens, G.H.; Hoeks, J.; van der Krieken,
S.; Ryu, D.; Kersten, S.; Moonen-Kornips, E.; Hesselink,
M.K.; Kunz, I.; Schrauwen-Hinderling, V.B.; Blaak, E.E.;
Auwerx, J.; Schrauwen, P., Calorie restriction-like effects
of 30 days of resveratrol supplementation on energy me-
tabolism and metabolic profile in obese humans. Cell Me-
tab, 2011, 14, (5), 612-622.
[138] Tome-Carneiro, J.; Gonzalvez, M.; Larrosa, M.; Garcia-
Almagro, F.J.; Aviles-Plaza, F.; Parra, S.; Yanez-Gascon,
M.J.; Ruiz-Ros, J.A.; Garcia-Conesa, M.T.; Tomas-
Barberan, F.A.; Espin, J.C., Consumption of a grape extract
supplem ent containing resv eratrol decreases oxidized LDL
and ApoB in patients undergoing primary prevention of
cardiovascular disease: a triple-blind, 6-month follow-up,
placebo-controlled, randomized trial. Mol Nutr Food Res,
2012, 56, (5), 810-821.
[139] Khandelwal, A.R.; Hebert, V.Y.; Dugas, T.R., Essential
role of ER-alpha-dependent NO production in resveratrol-
mediated inhibition of restenosis. Am J Physiol Heart Circ
Physiol, 2010, 299, (5), H1451-1458.
[140] Belguendouz, L.; Fremont, L.; Linard, A., Resveratrol in-
hibits metal ion-dependent and independent peroxidation of
porcine low-density lipoproteins. Biochem Pharmacol,
1997, 53, (9), 1347-1355.
[141] Kim, J.W.; Lim, S.C.; Lee, M.Y.; Lee, J.W.; Oh, W.K.;
Kim, S.K.; Kang, K.W., Inhibition of neointimal formation
by trans-resveratrol: role of phosphatidyl inositol 3-kinase-
dependent Nrf2 activation in heme oxygenase-1 induction.
Mol Nutr Food Res, 2010, 54, (10), 1497-1505.
[142] Brito, P.M.; Devillard, R.; Negre-Salvayre, A.; Almeida,
L.M.; Dinis, T.C.; Salvayre, R.; Auge, N., Resveratrol in-
hibits the mTOR mitogenic signaling evoked by oxidized
LDL in smooth muscle cells. Atherosclerosis, 2009, 205,
(1), 126-134.
[143] Szkudelski, T.; Szkudelska, K., Anti-diabetic effects of
resveratrol. Ann N Y Acad Sci, 2011, 1215, 34-39.
[144] Arrick, D.M.; Sun, H.; Patel, K.P.; Mayhan, W.G., Chronic
resveratrol treatment restores vascular responsiveness of
cerebral art erioles in type 1 diabetic rats. Am J Physiol
Heart Circ Physiol, 2011, 301, (3), H696-703.
[145] Sulaiman, M.; Matta, M.J.; Sunderesan, N.R.; Gupta, M.P.;
Periasamy, M.; Gupta, M., Resveratrol, an activator of
SIRT1, upregulates sarcoplasmic calcium ATPase and im-
proves cardiac function in diabetic cardiomyopathy. Am J
Physiol Heart Circ Physiol, 2010, 298, (3), H833-843.
[146] Deng, J.Y.; Hsi eh, P.S.; Hu ang, J.P.; Lu, L .S.; Hung, L.M.,
Activation of estrogen receptor is crucial for resveratrol-
stimulating muscular glucose uptake via both insulin-
dependent and -independent pathways. Diabetes, 2008, 57,
(7), 1814-1823.
[147] Palsamy, P.; Subramanian, S., Modulatory effects of res-
veratrol on attenuating the key enzymes activities of carbo-
hydrate metabolism in streptozotocin-nicotinamide-induced
diabetic rats. Chem Biol Interact, 2009, 179, (2-3), 356-362.
[148] Vetterli, L.; Brun, T.; Giovannoni, L.; Bosco, D.; Maechler,
P., Resveratrol potentiates glucose-stimulated insulin secre-
tion in INS-1E beta-cells and human islets through a
SIRT1-dependent mechanism. J Biol Chem, 2011, 286, (8),
6049-6060.
[149] Hardie, D.G.; Pan, D.A., Regulation of fatty acid synthesis
and oxidation by the AMP-activated protein kinase. Bio-
chem So c Tran s, 2002, 30, (Pt 6), 1064-1070.
[150] Sun, C.; Zhang, F.; Ge, X.; Yan, T.; Chen, X.; Shi, X.; Zhai,
Q., SIRT1 improves insulin sensitivity under insulin-
resistant conditions by repressing PTP1B. Cell Metab,
2007, 6, (4), 307-319.
[151] Um, J.H.; Park, S.J.; Kang, H.; Yang, S.; Foretz, M.;
McBurney, M.W.; Kim, M.K.; Viollet, B.; Chung, J.H.,
AMP-activated protein kinase-deficient mice are resistant to
the metabolic effects of resveratrol. Diabetes, 2010, 59, (3),
554-563.
[152] Olas, B.; Wachowicz, B.; Saluk-Juszczak, J.; Z ielinsk i, T.,
Effect of resveratrol, a natural polyphenolic compound, on
platelet activation induced by endotoxin or thrombin.
Thromb Res, 2002, 107, (3-4), 141-145.
[153] Abe, K.; Tawara, S.; Oi, K.; Hizume, T.; Uwatoku, T.; Fu-
kumoto, Y.; Kaibuchi, K.; Shimokawa, H., Long-term inhi-
bition of Rho-kinase ameliorates hypoxia-induced pulmo-
nary hypertension in mice. J Cardiovasc Pharmacol, 2006,
48, (6), 280-285.
[154] Bonaventura, A.; Montecucco, F.; Dallegri, F., Cellular
recruitm ent in myocardial ischem ia/reperfusion injury . Eur
J Clin Invest, 2016, 46(6), 590-601
[155] Montecucco, F.; Carbone, F.; Schindler, T.H., Pathophysi-
ology of ST-segment elevation myocardial infarction: novel
mechanisms and treatments. Eur Heart J, 2016, 37, (16),
1268-1283.
[156] Gurusamy, N.; Ray, D.; Lekli, I.; Das, D.K., Red wine anti-
oxidant resveratrol-modified cardiac stem cells regenerate
infarcted myocardium. J Cell Mol Med, 2010, 14, (9), 2235-
2239.
[157] Hung, L.M.; Su, M.J.; Chen, J.K., Resveratrol protects
myocardial ischemia-reperfusion injury through both NO-
dependent and NO-independent mechanisms. Free Radic
Biol Med, 2004, 36, (6), 774-781.
[158] Petrovski, G.; Gurusamy, N.; Das, D.K., Resveratrol in
cardiovascular health and disease. Ann N Y Acad Sci, 2011,
1215, 22-33.
[159] Das, S.; Cordis, G.A.; Maulik, N.; Das, D.K., Pharmacol-
ogical preconditioning with resveratrol: role of CREB-
dependent Bcl-2 signaling via adenosine A3 receptor acti-
vation. Am J Physiol Heart Circ Physiol, 2005, 288, (1),
H328-335.
[160] Chen, C.J.; Yu, W.; Fu, Y.C.; Wang, X.; Li, J.L.; Wang,
W., Resveratrol protects cardiomyocytes from hypoxia-
induced apoptosis through the SIRT1-FoxO1 pathway. Bio-
chem Biophys Res Commun, 2009, 378, (3), 389-393.
[161] Gurusamy, N.; Lekli, I.; Mukherjee, S.; Ray, D.; Ahsan,
M.K.; Gherghiceanu, M.; Popescu, L.M.; Das, D.K., Car-
dioprotection by resveratrol: a novel mechanism via auto-
phagy involving the mTORC2 pathway. Cardiovasc Res,
2010, 86, (1), 103-112.
[162] Chen, Y.R.; Yi, F.F.; Li, X.Y.; Wang, C.Y.; Chen, L.;
Yang, X.C.; Su, P.X.; Cai, J., Resveratrol attenuates ven-
tricular arrhythmias and improves the long-term survival in
rats with myocardial infarction. Cardiovasc Drugs Ther,
2008, 22, (6), 479-485.
18 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
[163] Gu, X.S.; Wang, Z.B.; Ye, Z.; Lei, J.P.; Li, L.; Su, D.F.;
Zheng, X., Resveratrol, an activator of SIRT1, upregulates
AMPK and improves cardiac function in heart failure.
Genet Mol Res, 2014, 13, (1), 323-335.
[164] Kanamori, H.; Takemura, G.; Goto, K.; Tsujimoto, A.; Og-
ino, A.; Takeyama, T.; Kawaguchi, T.; Watanabe, T.; Mor-
ishita, K.; Kawasaki, M.; Mikami, A.; Fujiwara, T.; Fuji-
wara, H.; Seishima, M.; Minatoguchi, S., Resveratrol re-
verses remodeling in hearts with large, old myocardial in-
farctions th rough enhanced autophagy-activ ating AMP
kinase pathway. Am J Pathol, 2013, 182, (3), 701-713.
[165] Gupta, P.K.; DiPette, D.J.; Supowit, S.C., Protective effect
of resveratrol against pressure overload-induced heart fail-
ure. Food Sci Nutr, 2014, 2, (3), 218-229.
[166] Magyar, K.; Halmosi, R.; Palfi, A.; Feher, G.; Czopf, L.;
Fulop, A.; Battyany, I.; Sumegi, B.; Toth, K.; Szabados, E.,
Cardioprotection by resveratrol: A human clinical trial in
patients with stable coronary artery disease. Clin Hemor-
heol Microcirc, 2012, 50, (3), 179-187.
[167] Chan, A.Y.; Dolinsky, V.W.; Soltys, C.L.; Viollet, B.;
Baksh, S.; Light, P.E.; Dyck, J.R., Resveratrol inhibits car-
diac hypertrophy via AMP-activated protein kinase and
Akt. J Biol Chem , 2008, 283, (35), 24194-24201.
[168] Carbone, F.; Teixeira, P.C.; Braunersreuther, V.; Mach, F.;
Vuilleumier, N.; Montecucco, F., Pathophysiology and
Treatments of Oxidative Injury in Ischemic Stroke: Focus
on the Phagocytic NADPH Oxidase 2. Antioxid Redox Sig-
nal, 2015, 23, (5), 460-489.
[169] Clark, D.; Tuor, U.I.; Thompson, R.; Institoris, A.; Ku-
lynych, A.; Zhang, X.; Kinniburgh, D.W.; Bari, F.; Busija,
D.W.; Barber, P.A., Protection against recurrent stroke with
resveratrol: endothelial protection. PLoS One, 2012, 7, (10),
e47792.
[170] Huang, S.S.; Tsai, M.C.; Chih, C.L.; Hung, L.M.; Tsai,
S.K., Resveratrol reduction of infarct size in Long-Evans
rats subjected to focal cerebral ischemia. Life Sci, 2001, 69,
(9), 1057-1065.
[171] Singh, N.; Agrawal, M.; Dore, S., Neuroprotective proper-
ties and mechanisms of resveratrol in in vitro and in vivo
experimental cerebral stroke models. ACS Chem Neurosci,
2013, 4, (8), 1151-1162.
[172] Wan, D.; Zhou, Y.; Wang, K.; Hou, Y.; Hou, R.; Ye, X.,
Resveratrol provides neuroprotection by inhibiting phos-
phodiesterases and regulating the cAMP/AMPK/SIRT1
pathway after stroke in rats. Brain Res Bull, 2016, 121, 255-
262.
[173] de la Lastra, C.A.; Villegas, I., Resveratrol as an anti-
inflammatory and anti-aging agent: mechanisms and clini-
cal implications. Mol Nutr Food Res, 2005, 49, (5), 405-
430.
[174] Jin, F.; Wu, Q.; Lu, Y.F.; Gong, Q.H.; Shi, J.S., Neuropro-
tective effect of resv eratrol on 6-OHDA-induced Parkin-
son's disease in rats. Eur J Pharmacol, 2008, 600, (1-3), 78-
82.
[175] Kennedy, D.O.; Wightman, E.L.; Reay, J.L.; Lietz, G.;
Okello, E.J.; Wilde, A.; Haskell, C.F., Effects of resveratrol
on cerebral blood flow variables and cognitive performance
in humans: a double-blind, placebo-controlled, crossover
investigation. Am J Clin Nutr, 2010, 91, (6), 1590-1597.
[176] Wightman, E.L.; Reay, J.L.; Haskell, C.F.; Williamson, G.;
Dew, T.P.; Kennedy, D.O., Effects of resveratrol alone or in
combination with piperine on cerebral blood flow parame-
ters and cognitive performance in human subjects: a ran-
domised, double-blind, placebo-controlled, cross-over in-
vestigation. Br J Nutr, 2014, 112, (2), 203-213.
[177] Evans, H.M.; Howe, P.R.; Wong, R.H., Clinical Evaluation
of Effects of Chronic Resveratrol Supplementation on
Cerebrovascular Function, Cognition, Mood, Physical
Function and General Well-Being in Postmenopausal
Women-Rationale and Study Design. Nutrients, 2016, 8,
(3), 150.
[178] Chen, J.; Bai, Q.; Zhao, Z.; Sui, H.; Xie, X., Resveratrol
improves delayed r-tPA treatment outcome by reducing
MMPs. Acta Neurol Scand, 2016, 134, (1), 54-60.
[179] Shao, A.W.; Wu, H.J.; Chen, S.; Ammar, A.B.; Zhang,
J.M.; Hong, Y., Resv eratrol attenuates early brain injury af-
ter subarachnoid hemorrh age through inhibition of NF-
kappaB-dependent inflammatory/MMP-9 pathway. CNS
Neurosci Ther, 2014, 20, (2), 182-185.
[180] Gao, D.; Huang, T.; Jiang, X.; Hu, S.; Zhang, L.; Fei, Z.,
Resveratrol protects primary cortical neuron cultures from
transient oxygen-glucose deprivation by inhibiting MMP-9.
Mol Med Rep, 2014, 9, (6), 2197-2204.
[181] Wei, H.; Wang, S.; Zhen, L.; Yang, Q.; Wu, Z.; Lei, X.; Lv,
J.; Xiong, L.; Xue, R. , Resveratrol attenu ates the blood-
brain barrier dysfunction by regulation of the MMP-
9/TIMP-1 balance after cerebral ischemia reperfusion in
rats. J Mol Neurosci, 2015, 55, (4), 872-879.
[182] Krenz, M.; Korthuis, R.J., Moderate ethanol ingestion and
cardiovascular protection: from epidemiologic associations
to cellular mechanisms. J Mo l Cell Cardiol, 2012, 52, (1),
93-104.
[183] Yusuf, S.; Hawken, S.; Ounpuu, S.; Dans, T.; Avezum, A.;
Lanas, F.; McQueen, M.; Budaj, A.; Pais, P.; Varigos, J.;
Lisheng, L.; Investigators, I.S., Effect of potentially modifi-
able risk factors associated with myocardial infarction in 52
countries (the INTERHEART study): case-control study.
Lancet, 2004, 364, (9438), 937-952.
[184] Mukamal, K.J.; Jensen, M.K.; Gronbaek, M.; Stampfer,
M.J.; Manson, J.E.; Pischon, T.; Rimm, E.B., Drinking fre-
quency, mediating biomarkers, and risk of myocardial in-
farction in women and men. Circulation, 2005, 112, (10),
1406-1413.
[185] Mukamal, K.J.; Chiuve, S.E.; Rimm, E.B., Alcohol con-
sumption and risk for coronary heart disease in men with
healthy lifestyles. Arch Intern Med, 2006, 166, (19), 2145-
2150.
[186] de Leiris, J.; Besse, S.; Boucher, F., Diet and heart health:
moderate wine drinking strengthens the cardioprotective ef-
fects of fish consumption. Curr Pharm Biotechnol, 2010,
11, (8), 911-921.
[187] Patra, J.; Taylor, B.; Irving, H.; Roerecke, M.; Baliunas, D.;
Mohapatra, S.; Rehm, J., Alcohol consumption and the risk
of morbidity and mortality for different stroke types--a sys-
tematic review and meta-analysis. BMC Public Health,
2010, 10, 258.
[188] Djousse, L.; Gaziano, J.M., Alcohol consumption and heart
failure: a systematic review. Curr Atheroscler Rep, 2008,
10, (2), 117-120.
[189] Puddey, I.B.; Zilkens, R.R.; Croft, K.D.; Beilin, L.J., Alco-
hol and endothelial function: a brief review. Clin Exp
Pharmacol Physiol, 2001, 28, (12), 1020-1024.
[190] Abou-Agag, L.H.; Khoo, N.K.; Binsack, R.; White, C.R.;
Darley-Usmar, V.; Grenett, H.E.; Booyse, F.M.; Digerness,
S.B.; Zhou, F.; Parks, D.A., Evidence of cardiovascular
protection by moderate alcohol: role of nitric oxide. Free
Radic Biol Med, 2005, 39, (4), 540-548.
[191] Gazzieri, D.; Trevisan i, M.; Tarantini, F.; Bechi, P.; Ma-
sotti, G.; Gensin i, G.F.; Castellani, S.; Marchionni, N.;
Geppetti, P.; Harrison, S., Ethanol dilates coronary arteries
and increases coronary flow via transient receptor potential
vanilloid 1 and calcitonin gene-related peptide. Cardiovasc
Res, 2006, 70, (3), 589-599.
[192] Carbone, F.; Montecucco, F., Inflammation in arterial dis-
eases. IUBMB Life, 2015, 67, (1 ), 18-28.
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 19
[193] Badia, E.; Sacanella, E.; Fernandez-Sola, J.; Nicolas, J.M.;
Antunez, E.; Rotilio, D.; de Gaetano, G.; Urbano-Marquez,
A.; Estruch, R., Decreased tumor necrosis factor-induced
adhesion of human monocytes to endothelial cells after
moderate alcohol consumption. Am J Clin Nutr, 2004, 80,
(1), 225-230.
[194] Kiviniemi, T.O.; Saraste, A.; Toikka, J.O.; Saraste, M.;
Raitakari, O.T.; Parkka, J.P.; Lehtimaki, T.; Hartiala, J.J.;
Viikari, J.; Koskenvuo, J.W., A moderate dose of red wine,
but not de-alcoholized red wine increases coronary flow re-
serve. Atherosclerosis, 2007, 195, (2), e176-181.
[195] Tousoulis, D.; Ntarladimas, I.; Antoniades, C.; Vasiliadou,
C.; Tentolouris, C.; Papageorgiou, N.; Latsios, G.; Stefa-
nadis, C., Acute effects of different alcoholic beverages on
vascular endothelium, inflammatory markers and thrombo-
sis fibrinolysis system. Clin Nutr, 2008, 27, (4), 594-600.
[196] Suzuki, K.; Elkind, M.S.; Boden-Albala, B.; Jin, Z.; Berry,
G.; Di Tullio, M.R.; Sacco, R.L.; Homma, S., Moderate al-
cohol consumption is associated with better endothelial
function: a cross sectional study. BMC Cardiovasc Disord,
2009, 9, 8.
[197] Vasdev, S.; Ford, C.A.; Longerich, L.; Parai, S.; Gadag, V.,
Antihypertensive effect of low ethanol intake in spontane-
ously hypertensive rats. Mol Cell Biochem, 1999, 200, (1-
2), 85-92.
[198] Briasoulis, A.; Agarwal, V.; Messerli, F.H., Alcohol con-
sumption and the risk of hyperten sion in men and women: a
systematic review and meta-analysis. J Clin Hypertens
(Greenwich), 2012, 14, (11), 792-798.
[199] Gardner, C.D.; Tribble, D.L.; Young, D.R.; Ahn, D.; Fort-
mann, S.P., Associations o f HDL, HDL(2), and HDL (3)
cholesterol and apolipoproteins A-I and B with lifestyle fac-
tors in healthy women and men: the Stanford Five City Pro-
ject. Prev Med, 2000, 31, (4), 346-356.
[200] Brinton, E.A., Effects of ethanol intake on lipoproteins.
Curr Atheroscler Rep, 2012, 14, (2), 108-114.
[201] Perret, B.; Ruidavets, J.B.; Vieu, C.; Jaspard, B.; Cambou,
J.P.; Terce, F.; Collet, X., Alcohol consumption is associ-
ated with en richment of high-density lipoprotein particles in
polyunsaturated lipids and increased cholesterol esterifica-
tion rate. Alcohol Clin Exp Res, 2002, 26, (8), 1134-1140.
[202] Sierksma, A.; Vermunt, S.H.; Lankhuizen, I.M.; van der
Gaag, M.S.; Scheek, L.M.; Grobbee, D.E.; van Tol, A.;
Hendriks, H.F., Effect of moderate alcohol consumption on
parameters of reverse cholesterol transport in postmeno-
pausal women. Alcohol Clin Exp Res, 2004, 28, (4), 662-
666.
[203] Beulens, J.W.; Sierksma, A.; van Tol, A.; Fournier, N.; van
Gent, T.; Paul, J.L.; Hendriks, H.F., Moderate alcohol con-
sumption increases cholesterol efflux mediated by ABCA1.
J Lipid Res, 2004, 45, (9), 1716-1723.
[204] Hoang, A.; Tefft, C.; Duffy, S.J.; Formosa, M.; Henstridge,
D.C.; Kingwell, B.A.; Sviridov, D., ABCA1 expression in
humans is associated with physical activity and alcohol
consumption. Atherosclerosis, 2008, 197, (1), 197-203.
[205] Liisanantti, M.K.; Savolainen, M.J., Phosphatidylethanol
mediates its effects on the vascular endothelial growth fac-
tor via HDL receptor in endothelial cells. Alcohol Clin Exp
Res, 2009, 33, (2), 283-288.
[206] Hannuksela, M.L.; Ramet, M.E.; Nissinen, A.E.; Liisan-
antti, M.K.; Savolainen, M.J., Effects of ethanol on lipids
and atherosclerosis. Pathophysiology, 2004, 10, (2), 93-103.
[207] Wakabayashi, I.; Groschner, K., Modification of the asso-
ciation between alcohol drinking and non-HDL cholesterol
by gender. Clin Chim Acta, 2009, 404, (2), 154-159.
[208] Tolstrup, J.S.; Gronbaek, M.; Nordestgaard, B.G., Alcohol
Intake, Myocardial Infarction, Biochemical Risk Factors,
and Alcohol Dehydrogenase Genotypes. Circ-Cardiovasc
Gene, 2009, 2, (5), 507-514.
[209] Perissinotto, E.; Buja, A.; Maggi, S.; Enzi, G.; Manzato, E.;
Scafato, E.; Mastrangelo, G.; Frigo, A.C.; Coin, A.; Cre-
paldi, G.; Sergi, G.; Group, I.W., Alcohol consumption and
cardiovascular risk factors in older lifelong wine drinkers:
the Italian Longitudinal Study on Aging. Nutr Metab Car-
diovasc Dis, 2010, 20, (9), 647-655.
[210] Yin, R.X.; Li, Y.Y.; Liu, W.Y.; Zhang, L.; Wu, J.Z., Inter-
actions of the apolipoprotein A5 gene polymorphisms and
alcohol consumption on serum lipid levels. PLoS One,
2011, 6, (3), e17954.
[211] Corella, D.; Portoles, O.; Arriola, L.; Chirlaque, M.D.;
Barrricarte, A.; Frances, F.; Huerta, J.M.; Larranaga, N.;
Martinez, C.; Martinez-Camblor, P.; Molina, E.; Navarro,
C.; Quiros, J.R.; Rodriguez, L.; Sanchez, M.J.; Ros, E.;
Sala, N.; Gonzalez, C.A.; Moreno-Iribas, C., Saturated fat
intake and alcohol consumption modulate the association
between the APOE polymorphism and risk of future coro-
nary heart disease: a nested case-control study in the Span-
ish EPIC cohort. J Nutr Biochem, 2011, 22, (5), 487-494.
[212] Wang, Z.; Yao, T.; Song, Z., Chronic alcohol consumption
disrupted cholesterol homeostasis in rats: down-regulation
of low-density lipoprotein receptor and enhancement of
cholesterol biosynthesis pathway in the liver. Alcohol Clin
Exp Res, 2010, 34, (3), 471-478.
[213] Ruixing, Y.; Yiyang, L.; Meng, L.; Kela, L.; Xingjiang, L.;
Lin, Z.; Wanying, L.; Jinzhen, W.; Dezhai, Y.; Weixiong,
L., Interactions of the apolipoprotein C-III 3238C>G poly-
morphism and alcohol consumption on serum triglyceride
levels. Lipids Health Dis, 2010, 9, 86.
[214] Pietraszek, A.; Gregersen, S.; Hermansen, K., Alcohol and
type 2 diabetes. A review. Nutr Metab Cardiovasc Dis,
2010, 20, (5), 366-375.
[215] Baliunas, D.O.; Taylor, B.J.; Irving, H.; Roerecke, M.; Pa-
tra, J.; Mohapatra, S.; Rehm, J., Alcohol as a Risk Factor
for Type 2 Diabetes A systematic review and meta-analy sis.
Diabetes Care, 2009, 32, (11), 2123-2132.
[216] Schrieks, I.C.; Heil, A.L.J.; Hendriks, H.F.J.; Mukamal,
K.J.; Beulens, J.W.J., The Effect of Alcohol Consumption
on Insulin Sensitivity and Glycemic Status: A Systematic
Review and Meta-analysis of Intervention Studies. Diabetes
Care, 2015, 38, (4), 723-732.
[217] Lieber, C.S., Alcohol and the liver: 1984 update. Hepatol-
ogy, 1984, 4, (6), 1243-1260.
[218] Sierksma, A.; Patel, H.; Ouchi, N.; Kihara, S.; Funahashi,
T.; Heine, R.J.; Grobbee, D.E.; Kluft, C.; Hendriks, H.F.,
Effect of moderate alcohol consumption on adiponectin,
tumor necrosis factor-alpha, and insulin sensitivity. Diabe-
tes Care, 2004, 27, (1), 184-189.
[219] Carbone, F.; Mach, F.; Montecucco, F., The role of adipo-
cytokines in atherogenesis and atheroprogression. Curr
Drug Targets, 2015, 16, (4), 295-320.
[220] Ruan, H.; Dong, L.Q., Adiponectin signaling and function
in insulin target tissues. J Mol Cell Biol, 2016, 8(2):101-9
[221] Imhof, A.; Froehlich, M.; Brenner, H.; Boeing, H.; Pepys,
M.B.; Koenig, W., Effect of alcohol consumption on sys-
temic markers of inflammation. Lancet, 2001, 357, (9258),
763-767.
[222] Sierksma, A.; van der Gaag, M.S.; Kluft, C.; Hendriks,
H.F., Moderate alcohol consumption reduces plasma C-
reactive protein and fibrinogen levels; a randomized, diet-
controlled intervention study. Eur J Clin Nutr, 2002, 56,
(11), 1130-1136.
[223] Albert, M.A.; Glynn, R.J.; Ridker, P.M., Alcohol consump-
tion and plasma concentration of C-reactive protein. Circu-
lation, 2003, 107, (3), 443-447.
20 Current Medicinal Chemistry, 2017, Vol. 24, No. 00 Liberale et al.
[224] Wandler, A.; Bruun, J.M.; Nielsen, M.P.; Richelsen, B.,
Ethanol exerts anti-inflammatory effects in human adipose
tissue in vitro. Mol Cell Endocrinol, 2008, 296, (1-2), 26-
31.
[225] Miceli, M.; Alberti, L.; Bennardini, F.; Di Simplicio, P.;
Seghieri, G.; Rao, G.H.; Franconi, F., Effect of low doses of
ethanol on platelet function in long-life abstainers and mod-
erate-wine drinkers. Life Sci, 2003, 73, (12), 1557-1566.
[226] Estruch, R.; Sacanella, E.; Mota, F.; Chiva-Blanch, G.;
Antunez, E.; Casals, E.; Deulofeu, R.; Rotilio, D.; Andres-
Lacueva, C.; Lamuela-Raventos, R.M.; de Gaetano, G.; Ur-
bano-Marquez, A., Moderate consumption of red wine, but
not gin, decreases erythrocyte superoxide dismutase activ-
ity: a randomised cross-over trial. Nutr Metab Cardiovasc
Dis, 2011, 21, (1), 46-53.
[227] Jensen, T.; Retterstol, L.J.; Sandset, P.M.; Godal, H.C.;
Skjonsberg, O.H., A daily glass of red wine induces a pro-
longed reduction in plasma viscosity: a randomized con-
trolled trial. Blood Coagul Fibrinolysis, 2006, 17, (6), 471-
476.
[228] Pieters, M.; Vorster, H.H.; Jerling, J.C.; Venter, C.S.;
Kotze, R.C.; Bornman, E.; Malfliet, J.J.; Rijken, D.C., The
effect of ethanol and its metabolism on fibrinolysis. Thromb
Haemost, 2010, 104, (4), 724-733.
[229] Reynolds, K.; Lewis, B.; Nolen, J.D.; Kinney, G.L.; Sathya,
B.; He, J., Alcohol consumption and risk of stroke: a meta-
analysis. JAMA, 2003, 289, (5), 579-588.
[230] Pomp, E.R.; Rosendaal, F.R.; Doggen, C.J., Alcohol con-
sumption is associated with a decreased risk of venous
thrombosis. Thromb Haemost, 2008, 99, (1), 59-63.
[231] Miyamae, M.; Rodriguez, M.M.; Camacho, S.A.; Diamond,
I.; Mochly-Rosen, D.; Figueredo, V.M., Activation of epsi-
lon protein kinase C correlates with a cardioprotective ef-
fect of regular ethanol consumption. Proc Natl Acad Sci U
S A, 1998, 95, (14), 8262-8267.
[232] Zhou, H.Z.; Karliner, J.S.; Gray, M.O., Moderate alcohol
consumption induces sustained cardiac protection by acti-
vating PKC-epsilon and Akt. Am J Physiol Heart Circ
Physiol, 2002, 283, (1), H165-174.
[233] Zhu, P.; Zhou, H.Z.; Gray, M.O., Chronic ethanol-induced
myocardial protection requires activation of mitochondrial
K(ATP) channels. J Mol Cell Cardiol, 2000, 32, (11), 2091-
2095.
[234] Yang, X.; Cohen, M.V.; Downey, J.M., Mechanism of car-
dioprotection by early ischemic preconditioning. Cardio-
vasc Drugs Ther, 2010, 24, (3), 225-234.
[235] Hausenloy, D.J.; Yellon, D.M., The second window of pre-
conditioning (SWOP) where are we now ? Cardiovasc
Drugs Ther, 2010, 24, (3), 235-254.
[236] Chen, C.H.; Budas, G.R.; Churchill, E.N.; Disatnik, M.H.;
Hurley, T.D.; Mochly-Rosen, D., Activation of aldehyde
dehydrogenase-2 reduces ischemic damage to the heart.
Science, 2008, 321, (5895), 1493-1495.
[237] Dayton, C.; Yamaguchi, T.; Kamada, K.; Carter, P.;
Korthuis, R.J., Antecedent ethanol ingestion prevents
postischemic leukocyte adhesion and P-selectin expression
by a protein kinase C-dependent mechanism. Dig Dis Sci,
2005, 50, (4), 684-690.
[238] Yamaguchi, T.; Dayton, C.; Shigematsu, T.; Carter, P.;
Yoshikawa, T.; Gute, D.C.; Korthuis, R.J., Preconditioning
with ethanol prevents postischemic leukocyte-endothelial
cell adhesive interactions. Am J Physiol Heart Circ Physiol,
2002, 283, (3), H1019-1030.
[239] Gaskin, F.S.; Kamada, K.; Yusof, M.; Durante, W.; Gross,
G.; Korthuis, R.J., AICAR preconditioning prevents
postischemic leukocyte rolling and adhesion: role of
K(ATP) channels and heme oxygenase. Microcirculation,
2009, 16, (2), 167-176.
[240] Rakotovao, A.; Berthonneche, C.; Guiraud, A.; de Lorgeril,
M.; Salen, P.; de Leiris, J.; Boucher, F., Ethanol, wine, and
experimental cardioprotection in isch emia/ reperfusion: role
of the prooxidant/antioxidant balance. Antioxid Redox Sig-
nal, 2004, 6, (2), 431-438.
[241] Churchill, E.N.; Disatnik, M.H.; Mochly-Rosen, D., Time-
dependent and ethanol-induced cardiac protection from
ischemia mediated by mitochondrial translocation of varep-
silonPKC and activation of aldehyde dehydrogenase 2. J
Mol Cell Cardiol, 2009, 46, (2), 278-284.
[242] Yamaguchi, T.; Kamada, K.; Dayton, C.; Gaskin, F.S.;
Yusof, M.; Yoshikawa, T.; Carter, P.; Korthuis, R.J., Role
of eNOS-derived NO in the postischemic anti-inflammatory
effects of anteceden t ethanol ingestion in murine small in-
testine. Am J Physiol Heart Circ Physiol, 2007, 292, (3),
H1435-1442.
[243] de Gaetano, G.; Di Castelnuovo, A.; Rotondo, S.; Ia-
coviello, L.; Donati, M.B., A meta-analysis of studies on
wine and beer and cardiovascular disease. Pathophysiol
Haemost Thromb, 2002, 32, (5-6), 353-355.
[244] Costanzo, S.; Di Castelnuovo, A.; Donati, M.B.; Iacoviello,
L.; de Gaetano, G., Wine, beer or spirit drinking in relation
to fatal and non-fatal cardiovascu lar events: a meta-
analysis. Eur J Epidemiol, 2011, 26, (11), 833-850.
[245] Rimm, E.B.; Klatsky, A.; Grobbee, D.; Stampfer, M.J.,
Review of moderate alcohol consumption and reduced risk
of coronary heart disease: is the effect due to beer, wine, or
spirits. BMJ, 1996, 312, (7033), 731-736.
[246] Brien, S.E.; Ronksley, P.E.; Turner, B.J.; Mukamal, K.J.;
Ghali, W.A., Effect of alcohol consumption on biological
markers associated with risk of coronary heart disease: sys-
tematic review and meta-analy sis of intervention al studies.
BMJ, 2011, 342, d636.
[247] Ronksley, P.E.; Brien, S.E.; Turner, B.J.; Mukamal, K.J.;
Ghali, W.A., Association of alcohol consumption with se-
lected cardiovascular disease outcomes: a system atic review
and meta-analysis. BMJ, 2011, 342, d671.
[248] Conigrave, K.M.; Hu, B.F.; Camargo, C.A., Jr.; Stampfer,
M.J.; Willett, W.C.; Rimm, E.B., A prospective study of
drinking patterns in relation to risk of type 2 diabetes
among men. Diabetes, 2001, 50, (10), 2390-2395.
[249] Mukamal, K.J.; Conigrave, K.M.; Mittleman, M.A.;
Camargo, C.A., Jr.; Stampfer, M.J.; Willett, W.C.; Rimm,
E.B., Roles of drinking pattern and type of alcohol con-
sumed in coronary heart disease in men. N Engl J Med,
2003, 348, (2), 109-118.
[250] Stranges, S.; Wu, T.; Dorn, J.M.; Freudenheim, J.L.; Muti,
P.; Farinaro, E.; Russell, M.; Nochajski, T.H.; Trevisan, M.,
Relationship of alcohol drinking pattern to risk of hyperten-
sion: a population-based study. Hypertension, 2004, 44, (6),
813-819.
[251] Della Valle, E.; Stranges, S.; Trevisan, M.; Krogh, V.; Fus-
coni, E.; Dorn, J.M.; Farinaro, E., Drinking habits and
health in Northern Italian and American men. Nutr Metab
Cardiovasc Dis, 2009, 19, (2), 115-122.
[252] Whelan, A.P.; Sutherland, W.H.; McCormick, M.P.; Yeo-
man, D.J.; de Jong, S.A.; Williams, M.J., Effects of white
and red wine on endothelial function in subjects with co ro-
nary artery disease. Intern M ed J, 2004, 34, (5), 224-228.
[253] Lekakis, J.; Rallid is, L.S.; Andreadou, I.; Vamvakou, G.;
Kazantzoglou, G.; Magiatis, P.; Skaltsounis, A.L.; Kre-
mastinos, D.T., Polyphenolic compounds from red grapes
acutely improve endothelial function in patients with coro-
nary heart disease. Eur J Cardiovasc Prev Rehabil, 2005,
12, (6), 596-600.
[254] Karatzi, K.N.; Papamichael, C.M.; Karatzis, E.N.; Pa-
paioannou, T.G.; Aznaouridis, K.A.; Katsichti, P.P.; Sta-
matelopoulos, K.S.; Zampelas, A.; Lekakis, J.P.; Mavrika-
Impact of Red Wine Consumption on Cardiovascular Health Current Medicinal Chemistry, 2017, Vol. 24, No. 00 21
kis, M.E., Red wine acutely induces favorable effects on
wave reflections and central pressures in coronary artery
disease patients. Am J Hypertens, 2005, 18, (9 Pt 1), 1161-
1167.
[255] Guarda, E.; Godoy, I.; Foncea, R.; Perez, D.D.; Romero, C.;
Venegas, R.; Leighton, F., Red wine reduces oxidative
stress in patients with acute coronary syndrome. Int J Car-
diol, 2005, 104, (1), 35-38.
[256] Marinaccio, L.; Lanza, G.A.; Niccoli, G.; Fabretti, A.;
Lamendola, P.; Barone, L.; Di Monaco, A.; Di Clemente,
F.; Crea, F., Effect of low doses of alcohol on the warm-up
phenomenon in patients with stable angina pectoris. Am J
Cardiol, 2008, 102, (2), 146-149.
[257] Tresserra-Rimbau, A.; Medina-Remon, A.; Lamuela-
Raventos, R.M.; Bullo, M.; Salas-Salvado, J.; Corella, D.;
Fito, M.; Gea, A.; Gomez-Gracia, E.; Lapetra, J.; Aros, F.;
Fiol, M.; Ros, E.; Serra-Majem, L.; Pinto, X.; Munoz,
M.A.; Estruch, R.; Investigators, P.S., Moderate red wine
consumption is associated with a lower prevalence of the
metabolic syndrome in the PREDIMED population. Br J
Nutr, 2015, 113 Suppl 2, S121-130.
[258] Chiva-Blanch, G.; Urpi-Sarda, M.; Ros, E.; Valderas-
Martinez, P.; Casas, R.; Arranz, S.; Guillen, M.; Lamuela-
Raventos, R.M.; Llorach, R.; Andres-Lacueva, C.; Estruch,
R., Effects of red wine polyphenols and alcohol on glucose
metabolism and the lipid profile: a randomized clinical trial.
Clin Nutr, 2013, 32, (2), 200-206.
[259] Gepner, Y.; Golan, R.; Harman-Boehm, I.; Henkin, Y.;
Schwarzfuchs, D.; Shelef, I.; Durst, R.; Kovsan, J.; Bolotin,
A.; Leitersdorf, E.; Shpitzen, S.; Balag, S.; Shemesh, E.;
Witkow, S.; Tangi-Rosental, O.; Chassidim, Y.; Liberty,
I.F.; Sarusi, B.; Ben-Avraham, S.; Helander, A.; Ceglarek,
U.; Stumvoll, M.; Bluher, M.; Thiery, J.; Rudich, A.;
Stampfer, M.J.; Shai, I., Effects of Initiating Moderate Al-
cohol Intake on Cardiometabolic Risk in Adults With Type
2 Diabetes: A 2-Year Randomized, Controlled Trial. Ann
Intern Med, 2015, 163, (8), 569-579.
[260] Fernandez-Jarne, E.; Martinez-Losa, E.; Serrano-Martinez,
M.; Prado-Santamaria, M.; Brugarolas-Brufau, C.;
Martinez-Gonzalez, M.A., Type of alcoholic beverage and
first acute myocardial infarction: a case-control study in a
Mediterranean country. Clin Cardiol, 2003, 26, (7), 313-
318.
[261] Marfella, R.; Cacciapuoti, F.; Siniscalchi, M.; Sasso, F.C.;
Marchese, F.; Cinone, F.; Musacchio, E.; Marfella, M.A.;
Ruggiero, L.; Chiorazzo, G.; Liberti, D.; Chiorazzo, G.; Ni-
coletti, G.F.; Saron, C.; D'Andrea, F.; A mmendola, C.; Ver-
za, M.; Coppola, L., Effect of moderate red wine intake on
cardiac prognosis after recent acute myocardial infarction of
subjects with Type 2 diabetes mellitus. Diabet Med, 2006,
23, (9), 974-981.
[262] Oliveira, A.; Lopes, C.; Rodriguez-Artalejo, F., Adherence
to the Southern European Atlantic Diet and occurrence of
nonfatal acute myocardial infarction. Am J Clin Nutr, 2010,
92, (1), 211-217.
[263] Levantesi, G.; Marfisi, R.; Mozaffarian, D.; Franzosi, M.G.;
Maggioni, A.; Nicolosi, G.L.; Schweiger, C.; Silletta, M.;
Tavazzi, L.; Tognoni, G.; Marchioli, R., Wine consumption
and risk of cardiovascular events after myocardial infarc-
tion: results from the GISSI-Prevenzione trial. Int J Car-
diol, 2013, 163, (3), 282-287.
[264] Cosmi, F.; Di Giulio, P.; Masson, S.; Finzi, A.; Marfisi,
R.M.; Cosmi, D.; Scarano, M.; Tognoni, G.; Maggioni,
A.P.; Porcu, M.; Boni, S.; Cutrupi, G.; Tav azzi, L.; Latini,
R.; Investigators, G.-H., Regular wine consumption in
chronic heart failure: impact on outcomes, quality of life,
and circulating biomarkers. Circ Heart Fail, 2015, 8, (3),
428-437.
[265] Hernandez-Hernandez, A.; Gea, A.; Ruiz-Canela, M.;
Toledo, E.; Beunza, J.J.; Bes-Rastrollo, M.; Martinez-
Gonzalez, M.A., Mediterranean Alcohol-Drinking Pattern
and the Incidence of Cardiovascular Disease and Cardio-
vascular Mortality: The SUN Project. Nutrients, 2015, 7,
(11), 9116-9126.
DISCLAIMER: The above article has been published in Epub (ahead of print) on the basis of the materials provided by the author. The
Editorial Department reserves the right to make minor modifications for further improvement of the manuscript.
PMID: 2852 1683
... Recent studies have suggested that the health benefits of RW can be attributed to ethanol rather than reactive molecules. [6] Moderate alcohol consumption has been found to impede the permeability of monolayers caused by serum amyloid A1 (SAA1), as well as the expression of ICAM-1, VCAM-1, and monocyte adhesion. [7] Several studies also revealed that moderate and stable alcohol consumption is associated with a lower risk of coronary heart disease. ...
... [8,9] Imhof suggested that moderate ethanol intake for a short duration was more effective at inhibiting MCP-1 expression in monocytes than dealcoholized RW. [10] On the other hand, research has indicated that the benefits of moderate consumption of RW are due to the presence of a variety of polyphenolic compounds in the beverage. [6] Janega et al [11] found that the cardioprotective effect of alcohol-free RW extract was related to its anti-inflammatory properties. Additionally, Lombardo et al [12] evaluated the correlations between acute and chronic RW consumption and health to uncover the benefits of moderate RW consumption in patients with type 2 diabetes mellitus and indicated that RW has a potential effect on antioxidant stress, thrombosis and inflammation markers, lipid profile, and gut microbiota. ...
Article
Full-text available
Background Moderate red wine (RW) consumption is associated with a low risk of cardiovascular disease (CVD). However, few studies have evaluated the effects of RW and white wine (WW) on inflammatory markers related to atherosclerosis in healthy individuals and high-risk subjects for CVD. This study aimed to assess the effect of RW on inflammatory markers in healthy individuals and high-risk subjects for CVD compared with moderate alcohol consumption. Methods The Preferred Reporting Items for Systematic Reviews and Meta-Analyses 2020 (PRISMA) was followed in this study. The PubMed, Embase, Cochrane, Web of Science, SinoMed, EbscoHost, and ScienceDirect databases were searched. The risk of bias and quality of the included trials were assessed using the Cochrane Handbook. The main results are summarized in Stata 12. Results Twelve studies were included in the meta-analysis. The results demonstrated that RW significantly decreased circulating intercellular cell adhesion molecule-1, vascular cell adhesion molecule-1 (VCAM-1), tumor necrosis factor-alpha (TNF-α), lymphocyte function-associated antigen-1, and Sialyl-Lewis X expression on the surface of monocytes in healthy subjects, but not in patients with CVD. Additionally, RW significantly decreased Sialyl-Lewis X but increased clusters of differentiation 40 (CD40) expressed on the surface of T lymphocytes and significantly decreased C-C chemokine receptor type 2 (CCR2) and very late activation antigen 4 (VLA-4) expressed on the surface of monocytes. Interestingly, subgroup analysis also found that RW significantly decreased circulating interleukin-6 (IL-6) in Spain but not in other countries, and significantly increased αMβ2 (Mac-1) in the group that had an intervention duration of less than 3 weeks. Conclusions Moderate consumption of RW is more effective than WW in alleviating atherosclerosis-related inflammatory markers in healthy people rather than high-risk subjects for CVD, but this needs to be further confirmed by studies with larger sample sizes.
... There is a strong ambiguity surrounding RW consumption and health [3]. Guidelines for the prevention of cardiovascular and neoplastic diseases advise against alcohol consumption, but drinking low-to-moderate amounts of wine may have some beneficial effects on cardiovascular disease risk (CVD) in certain populations [4]. Prospective cohort studies have demonstrated that any form of alcohol increases the risk of cancer [5]. ...
... In fact, the European Code Against Cancer advises limiting or eliminating alcohol consumption [6], and the International Agency for Research on Cancer has classified the consumption of alcoholic beverages as carcinogenic to humans (Group 1) in a dose-response manner. High alcohol consumption has been correlated with an increased risk of cancers of mouth, pharynx and larynx, oesophagus (squamous cell carcinoma), liver, colorectum, breast (before and after menopause), and stomach, in addition to many other diseases, such as cirrhosis, infectious diseases, CVD, diabetes, neuropsychiatric conditions, and early dementia [3][4][5]. ...
Article
Full-text available
A strong controversy persists regarding the effect of red wine (RW) consumption and health. Guidelines for the prevention of cardiovascular diseases (CVD) and cancers discourage alcohol consumption in any form, but several studies have demonstrated that low RW intake may have positive effects on CVD risk. This review evaluated randomised controlled trials (RCTs), examining the recent literature on the correlations between acute and chronic RW consumption and health. All RCTs published in English on PubMed from 1 January 2000 to 28 February 2023 were evaluated. Ninety-one RCTs were included in this review, seven of which had a duration of more than six months. We assessed the effect of RW on: (1) antioxidant status, (2) cardiovascular function, (3) coagulation pathway and platelet function, (4) endothelial function and arterial stiffness, (5) hypertension, (6) immune function and inflammation status, (7) lipid profile and homocysteine levels, (8) body composition, type 2 diabetes and glucose metabolism, and (9) gut microbiota and the gastrointestinal tract. RW consumption mostly results in improvements in antioxidant status, thrombosis and inflammation markers, lipid profile, and gut microbiota, with conflicting results on hypertension and cardiac function. Notably, beneficial effects were observed on oxidative stress, inflammation, and nephropathy markers, with a modest decrease in CVD risk in five out of seven studies that evaluated the effect of RW consumption. These studies were conducted mainly in patients with type 2 diabetes mellitus, and had a duration between six months and two years. Additional long-term RCTs are needed to confirm these benefits, and assess the potential risks associated with RW consumption.
... Owing to its elevated nutritional value, exceptional flavor profile, and substantial profitability, grapes have garnered unprecedented popularity [1,2]. Wine, being the most paramount product of grape processing, exhibits certain effects in anti-aging [3], cerebrovascular protection [4,5], anti-cancerous [6], and so on. As the realization of wine's significance grew, its demand in China has experienced a rapid increase, propelling China to the helm of the fastest-growing wine consumer nation globally [7]. ...
Article
Full-text available
The application of fertilizers and soil quality are crucial for grape fruit quality. However, the molecular data linking different fertilizer (or soil conditioner [SC]) treatments with grape fruit quality is still lacking. In this study, we investigated three soil treatments, namely inorganic fertilizer (NPK, 343.5 kg/hm² urea [N ≥ 46%]; 166.5 kg/hm² P2O5 [P2O5 ≥ 64%]; 318 kg/hm² K2O [K2O ≥ 50%]), organic fertilizer (Org, 9 t/hm² [organic matter content ≥ 35%, N + P2O5 + K2O ≥ 13%]), and SC (SC, 3 t/hm² [humic acid ≥ 38.5%; C, 56.1%; H, 3.7%; N, 1.5%; O, 38%; S, 0.6%]), on 4-year-old Cabernet Sauvignon grapevines. Compared with the NPK- and Org-treated groups, the SC significantly improved the levels of soluble solids, tannins, anthocyanins, and total phenols in the grape berries, which are important biochemical indicators that affect wine quality. Furthermore, we conducted RNA-seq analysis on the grapevine roots from each of the three treatments and used weighted gene co-expression network analysis to identify five hub genes that were associated with the biochemical indicators of the grape berries. Furthermore, we validated the expression levels of three hub genes (ERF, JP, and SF3B) and five selected genes related to anthocyanin biosynthesis (UFGT1, UFGT2, UFGT3, GST, and AT) by using quantitative reverse transcription-polymerase chain reaction. Compared to the NPK and Org treatment groups, the SC treatment resulted in a significant increase in the transcription levels of three hub genes as well as VvUFGT1, VvUFGT3, VvGST, and VvAT. These results suggest that the SC can improve grape fruit quality by altering gene transcription patterns in grapevine roots and further influence the biochemical indices of grape fruits, particularly anthocyanin content. This study reveals that the application of SC can serve as an important measure for enhancing vineyard SC and elevating grape quality.
... In contrast, there is some evidence that red wine consumption has a protective effect against hypertension. Red wine polyphenols (RWPs), such as resveratrol, catechin, epicatechin, quercetin and anthocyanin, have a positive effect on hypertension, dyslipidemia and metabolic diseases [51]. Possible mechanisms for the protective effect of RWPs on hypertension may include the following. ...
Article
Full-text available
Purpose: Previous studies have indicated that the prevalence rate of hypertension in adolescents is high, but it has not received much attention and the influencing factors are unclear, especially in Yunnan Province, China. Materials and methods: A cluster sampling method was used to investigate 4781 freshmen in a college in Kunming, Yunnan Province from November to December. Demographic and lifestyle data were collected using questionnaires, and height, weight and blood pressure were measured. Decision tree model of hypertension in college students was established by Chi-square automatic interactive detection method. Results: Prevalence of prehypertension of systolic blood pressure (SBP) and diastolic blood pressure (DBP) were detected in 33.9% and 32.1%, respectively. Prevalence of hypertension of SBP and DBP was detected in 1.2% and 7.2%, respectively. The hypertension and prehypertension decision tree of SBP has gender (χ2 = 728.64, p < .001) at the first level and body mass index (BMI) (boys: χ2 = 55.98, p < .001; girls: χ2 = 79.58, p < .001) at the second level. The hypertension and prehypertension decision tree of DBP has gender (χ2 = 381.83, p < .001) at the first level, BMI (boys: χ2 = 40.54, p < .001; girls: χ2 = 48.79, p < .001) at the second level, only children (χ2 = 6.43, p = .04) and red wine consumption (χ2 = 8.17, p = .017) at the third level. Conclusions: The present study suggests that gender, BMI, only children and red wine consumption were the main factors affecting hypertension in college students in southwest border areas of China.
... Despite its alcohol content, a low-to-moderate wine consumption has been associated with beneficial health effects, due to its high abundance in bioactive phytochemicals (4,5). More specifically, previous studies have demonstrated that wine is a rich source of polyphenols, plant secondary metabolites that possess potent antioxidant (6), anti-inflammatory (7) and cardioprotective (8,9) properties. ...
... Several studies have demonstrated the relationship between moderate red wine consumption and a lower prevalence of cardiovascular diseases and metabolic syndromes [1][2][3][4]. The impact of red wine consumption on human health is associated with factors such as lower consumption of sweetened beverages in Mediterranean countries [5], but its benefits are also attributed to the phenolic compounds in grapes. ...
Article
Full-text available
The content of minerals and bioactive compounds in wine depends on various factors, among which are the origin of the grapes, their phenolic composition, and the winemaking process. This study monitored the physicochemical parameters, phenolic compound contents, and antioxidant capacity of the red grape Vitis vinifera L. “Cabernet Sauvignon” harvested in three Mexican vineyards during the first nine days of the fermentation process. The bioactive compounds and elemental composition (determined by inductively coupled plasma–optical emission spectrometry, ICP-OES) were correlated. The fermentation process decreased from 22 to 5 °Bx in all cases, while the acidity increased from 6.5 to 8 g of tartaric acid/L, decreasing the pH. The phenolic compounds extracted during the winemaking ranged from 1400 to 1600 gallic acid equivalent/L, while the antioxidant capacity was 9 mmol Trolox equivalent. The bioactive compounds identified by HPLC were resveratrol, piceid, catechin, and epicatechin. The presence of Na, Mg, and Fe was correlated with antioxidant capacity, while higher Mn, Pb, Zn, and Cu contents were related to the presence of resveratrol, piceid, and catechin in Cabernet Sauvignon wine. Thus, certain minerals present in the soil that were transferred to the V. vinifera grapes can influence the amount and type of bioactive compounds present in the wine. The phenolic content and, therefore, the organoleptic characteristics of the wine are related to the mineral composition of the vine-growing soil (origin).
Article
Full-text available
NAFLD is the most common chronic liver disease worldwide, characterized by lipid accumulation in the liver, and usually evolves from steatohepatitis to fibrosis, cirrhosis, or even HCC. Its incidence is rapidly rising in parallel with the increasing prevalence of obesity and metabolic syndrome. Current therapies are limited to lifestyle changes including dietary intervention and exercise, in which dietary modification exerts an important part in losing weight and preventing NAFLD. In this review, we briefly discuss the roles and mechanisms of dietary components including fructose, non-nutritive sweeteners, fat, proteins, and vitamins in the progression or prevention of NAFLD. We also summarize several popular dietary patterns such as calorie-restricted diets, intermittent fasting, ketogenic diets, Mediterranean diets, and dietary approach to stop hypertension diets and compare the effects of low-fat and low-carbohydrate diets in preventing the development of NAFLD. Moreover, we summarize the potential drugs targeting metabolic-related targets in NAFLD.
Article
Cardiovascular diseases (CVDs) pose a serious threat to human health and incidence is increasing gradually. Nutrition has an important impact on the prophylaxis and progression of CVD. In this article, general attention is drawn to the possible positive effects of berries on CVD. Polyphenols have beneficial effects on the vascular system by inhibiting low-density lipoprotein oxidation and platelet aggregation, lowering blood pressure, improving endothelial dysfunction, and attenuating antioxidant defense and inflammatory responses. This review provides an overview of the effects of berries for the prevention and treatment of CVDs. Berries contain several cardioprotective antioxidants, vitamins, and numerous phytochemicals, such as phenolic compounds, that have antioxidant properties and antiplatelet activity. Phytochemical compounds in their structures can modulate dissimilar signaling pathways related to cell survival, differentiation, and growth. Important health benefits of berries include their antioxidant roles and anti-inflammatory impacts on vascular function. The effectiveness and potential of polyphenols primarily depend on the amount of bioavailability and intake. Although circulating berry metabolites can improve vascular function, their biological activities, mechanisms of action, and in vivo interactions are still unknown. Analyzing human studies or experimental studies to evaluate the bioactivity of metabolites individually and together is essential to understanding the mechanisms by which these metabolites affect vascular function.
Article
Full-text available
Obesity-linked type 2 diabetes is one of the paramount causes of morbidity and mortality worldwide, posing a major threat on human health, productivity, and quality of life. Despite great progress made towards a better understanding of the molecular basis of diabetes, the available clinical counter-measures against insulin resistance, a defect that is central to obesity-linked type 2 diabetes, remain inadequate. Adiponectin, an abundant adipocyte-secreted factor with a wide-range of biological activities, improves insulin sensitivity in major insulin target tissues, modulates inflammatory responses, and plays a crucial role in the regulation of energy metabolism. However, adiponectin as a promising therapeutic approach has not been thoroughly explored in the context of pharmacological intervention, and extensive efforts are being devoted to gain mechanistic understanding of adiponectin signaling and its regulation, and reveal therapeutic targets. Here, we discuss tissue- and cell-specific functions of adiponectin, with an emphasis on the regulation of adiponectin signaling pathways, and the potential crosstalk between the adiponectin and other signaling pathways involved in metabolic regulation. Understanding better just why and how adiponectin and its downstream effector molecules work will be essential, together with empirical trials, to guide us to therapies that target the root cause(s) of type 2 diabetes and insulin resistance.
Article
Full-text available
Background: This methodological paper presents both a scientific rationale and a methodological approach for investigating the effects of resveratrol supplementation on mood and cognitive performance in postmenopausal women. Postmenopausal women have an increased risk of cognitive decline and dementia, which may be at least partly due to loss of beneficial effects of estrogen on the cerebrovasculature. We hypothesise that resveratrol, a phytoestrogen, may counteract this risk by enhancing cerebrovascular function and improving regional blood flow in response to cognitive demands. A clinical trial was designed to test this hypothesis. Method: Healthy postmenopausal women were recruited to participate in a randomised, double-blind, placebo-controlled (parallel comparison) dietary intervention trial to evaluate the effects of resveratrol supplementation (75 mg twice daily) on cognition, cerebrovascular responsiveness to cognitive tasks and overall well-being. They performed the following tests at baseline and after 14 weeks of supplementation: Rey Auditory Verbal Learning Test, Cambridge Semantic Memory Battery, the Double Span and the Trail Making Task. Cerebrovascular function was assessed simultaneously by monitoring blood flow velocity in the middle cerebral arteries using transcranial Doppler ultrasound. Conclusion: This trial provides a model approach to demonstrate that, by optimising circulatory function in the brain, resveratrol and other vasoactive nutrients may enhance mood and cognition and ameliorate the risk of developing dementia in postmenopausal women and other at-risk populations.
Article
Full-text available
Background: We assessed the still unclear effect of the overall alcohol-drinking pattern, beyond the amount of alcohol consumed, on the incidence of cardiovascular clinical disease (CVD). Methods: We followed 14,651 participants during up to 14 years. We built a score assessing simultaneously seven dimensions of alcohol consumption to capture the conformity to a traditional Mediterranean alcohol-drinking pattern (MADP). It positively scored moderate alcohol intake, alcohol intake spread out over the week, low spirit consumption, preference for wine, red wine consumption, wine consumed during meals and avoidance of binge drinking. Results: During 142,177 person-years of follow-up, 127 incident cases of CVD (myocardial infarction, stroke or cardiovascular mortality) were identified. Compared with the category of better conformity with the MADP, the low-adherence group exhibited a non-significantly higher risk (HR) of total CVD ((95% CI) = 1.55 (0.58-4.16)). This direct association with a departure from the traditional MADP was even stronger for cardiovascular mortality (HR (95% CI) = 3.35 (0.77-14.5)). Nevertheless, all these associations were statistically non-significant. Conclusion: Better conformity with the MADP seemed to be associated with lower cardiovascular risk in most point estimates; however, no significant results were found and more powered studies are needed to clarify the role of the MADP on CVD.
Article
Background: Myocardial infarction is strictly linked to atherosclerosis. Beyond the mechanical narrowing of coronary vessels lumen, during MI a great burden of inflammation is carried out. One of the crucial events is represented by the ischemia/reperfusion injury, a complex event involving inflammatory cells (such as neutrophils, platelets, monocytes/macrophages, lymphocytes, and mast cells) and key activating signals (such as cytokines, chemokines, and growth factors). Cardiac repair following myocardial infarction is dependent on a finely regulated response involving sequential recruitment and clearance of different subsets of inflammatory cells. Materials and methods: This narrative review was based on the works detected on PubMed and MEDLINE up to November 2015. Results: Infarct healing classically follows three overlapping phases: the inflammatory phase, in which the innate immune pathways are activated and inflammatory leukocytes are recruited in order to clear the wound from dead cells; the proliferative phase, characterized by the suppression of pro-inflammatory signaling and infiltration of "repairing" cells secreting matrix proteins in the injured area; and the maturation phase, that is associated with quiescence and elimination of the reparative cells together with cross-linking of the matrix. All these phases are timely regulated by production soluble mediators, such as cytokines, chemokines, and growth factors. Conclusion: Targeting inflammatory cell recruitment early during reperfusion and healing might be promising to selectively abrogate injury and favoring repair. This approach might substantially improve adverse post-ischemic left ventricle remodeling, characterized by dilation, hypertrophy of viable segments, and progressive dysfunction. This article is protected by copyright. All rights reserved.
Article
Dysfunction of energy metabolism can be a significant and fundamental pathophysiological basis for strokes. In studies of both humans and rodents, resveratrol, a natural polyphenol, has been reported to provide protection from cerebral ischemic injury by regulating expression of silent mating type information regulation 2 homolog 1 (SIRT1). However, direct evidence demonstrating that resveratrol exerts neuroprotection from cerebral ischemia injury by decreasing energy consumption is still lacking. Therefore, the aim of this study was to elucidate the mechanisms and signaling pathways through which resveratrol regulates energy metabolism in the ischemic brain, and to identify potential targets of resveratrol. ATP levels in brain tissues were detected by high performance liquid chromatography. SIRT1 and the phosphorylation of adenosine-monophosphate-activated protein kinase (P-AMPK) expressiones were evaluated by western blot. Levels of phosphodiesterase (PDEs) and cAMP were quantitated by real-time PCR and ELISA, respectively. Results showed that resveratrol significantly reduced the harmful effects of cerebral ischemic injury in vivo. Moreover, levels of ATP, p-AMPK, SIRT1, and cAMP were increased by resveratrol and PDE inhibitors. In conclusion, our findings indicate that resveratrol provides neuroprotection by inhibiting PDEs and regulating the cAMP/AMPK/SIRT1 pathway, which reduces ATP energy consumption during ischemia.
Article
Despite major advances in mechanical and pharmacological reperfusion strategies to improve acute myocardial infarction (MI) injury, substantial mortality, morbidity, and socioeconomic burden still exists. To further reduce infarct size and thus ameliorate clinical outcome, the focus has also shifted towards early detection of MI with high-sensitive troponin assays, imaging, cardioprotection against pathophysiological targets of myocardial reperfusion injury with mechanical (ischaemic post-conditioning, remote ischaemic pre-conditioning, therapeutic hypothermia, and hypoxemia) and newer pharmacological interventions (atrial natriuretic peptide, cyclosporine A, and exenatide). Evidence from animal models of myocardial ischaemia and reperfusion also demonstrated promising results on more selective anti-inflammatory compounds that require additional validation in humans. Cardiac stem cell treatment also hold promise to reduce infarct size and negative remodelling of the left ventricle that may further improves symptoms and prognosis in these patients. This review focuses on the pathophysiology, detection, and reperfusion strategies of ST-segment elevation MI as well as current and future challenges to reduce ischaemia/reperfusion injury and infarct size that may result in a further improved outcome in these patients.
Article
Objectives: Although recombinant tissue plasminogen activator (r-tPA) is currently the most effective treatment for brain ischemic stroke, the 3-h narrow therapeutic windows severely limits its clinical efficacy. We aim to investigate the effect of resveratrol on improving treatment outcomes of delayed r-tPA administration. Materials & methods: Patients were randomly divided according to their onset-to-treatment time (OTT), as early OTT or delayed OTT. Then, they were either treated with r-tPA + placebo or with r-tPA + resveratrol. Twenty-four hours after the treatment, outcomes were assessed with NIH stroke scale (NIHSS), and plasma levels of MMP-2 and MMP-9 were also examined with ELISA. Results: In patients receiving delayed r-tPA treatment, co-administration of resveratrol significantly improves their treatment outcomes compared with those receiving placebo, as indicated by improved NIHSS scores. This improved outcome was be caused by resveratrol-induced reduction in plasma levels of both matrix metalloproteinase (MMP)-2 and MMP-9, as a positive correlation was observed between reductions in both MMPs and patient NIHSS scores. Conclusions: Resveratrol could be potentially administered as an adjuvant with r-tPA treatment, which extends the clinical therapeutic window of r-tPA, therefore improving the outcome of patients receiving late stroke treatment.