Você está na página 1de 10

Cardiovascular Research 63 (2004) 593 602

www.elsevier.com/locate/cardiores

Review

Vascular effects of wine polyphenols


Mario DellAgli *, Alessandra Busciala`, Enrica Bosisio
Faculty of Pharmacy, Department of Pharmacological Sciences, Via Balzaretti, 9-20133 Milan, Italy
Received 30 January 2004; received in revised form 4 March 2004; accepted 23 March 2004
Available online 6 May 2004
Time for primary review 20 days

Abstract

Keywords: Red wine polyphenols; Atherosclerosis; Endothelial function; Signal transduction; Vasodilation

1. Introduction
For many numbers of years, considerable attention has
been directed towards human behavioural habits that could
either be considered risk factors or even protective elements
for developing chronic pathologies. In particular, much effort
has been devoted to elucidate the role of diet in preventing
cardiovascular diseases. A so-called Mediterranean diet is
thought to prevent cardiovascular diseases [1], as a consequence of its high content of antioxidants, which are crucial in
ameliorating oxidative events implicated in many diseases.
Similarly, moderate consumption of red wine has been
associated with a lowering of the risk of coronary heart
disease [2 4]. Red wine represents a rich source of polyphenols such as anthocyanosides (ACs), catechins, proanthocyanidins (PAs), stilbenes and other phenolics. Anthocyanosides (ACs) are flavonoids widely distributed in fruits
and vegetables. They provide colour to red wines and to skins
of red and black grapes [5]. Proanthocyanidins are another
class of plant phenol metabolites widely occurring in fruits
* Corresponding author. Tel.: +39-02-50318345; fax: +39-02-50318391.
E-mail address: mario.dellagli@unimi.it (M. DellAgli).

including grapes [6]. ACs and PAs are among the most
important compounds in determining the quality of the red
wine, because they greatly influence colour, bitterness, astringency, and chemical stability toward oxidation [7].
In addition to antioxidant/antiradical activity, red wine
polyphenols (RWPs) were shown to possess many biological
properties including the inhibition of platelet aggregation,
vasorelaxing activity, modulation of lipid metabolism, and
inhibition of low-density lipoprotein oxidation [2,3,8 11].
Thus, the health benefits of moderate consumption of red
wine are founded on a multiplicity of actions. The biological
properties of red wine stilbenes (trans-resveratrol) have been
extensively reviewed [12 14]. Because ACs and PAs exert
cardioprotective effects which are not only due to antioxidant/
antiradical activity, the present review will consider the
effects of ACs and PAs at vascular level with special emphasis
on the vasorelaxing and the antiaggregating properties.
2. Proanthocyanidins and anthocyanosides in red grapes
(Vitis vinifera) and wines
PAs (or so-called condensed tannins) are high-molecular
weight polymers, made of flavan-3-ol units (Fig. 1) linked

0008-6363/$ - see front matter D 2004 European Society of Cardiology. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.cardiores.2004.03.019

Downloaded from by guest on August 8, 2016

Moderate consumption of red wine has been putatively associated with lowering the risk of developing coronary heart disease. This
beneficial effect is mainly attributed to the occurrence of polyphenol compounds such as anthocyanosides (ACs), catechins,
proanthocyanidins (PAs), stilbenes and other phenolics in red wine. This review focuses on the vascular effects of red wine polyphenols
(RWPs), with emphasis on anthocyanosides and proanthocyanidins. From in vitro studies, the effect of red wine polyphenols on the vascular
tone is thought to be due to short- and long-term mechanisms. NO-mediated vasorelaxation represents the short-term response to wine
polyphenols, which exert the effect by increasing the influx of extracellular Ca2 +, and the mobilization of intracellular Ca2 + in endothelial
cells. Polyphenolic compounds may also have long-term properties, as they increase endothelial NO synthase expression acting on the
promoter activity. In addition, they decrease the expression of adhesion molecules and growth factors, involved in migration and proliferation
of vascular smooth muscle cells. Moreover, they inhibit platelet aggregation. However, a paucity of data as regards the bioavailability and
metabolism of these compounds in human studies is a limiting factor to proving their efficacy in vivo.
D 2004 European Society of Cardiology. Published by Elsevier B.V. All rights reserved.

594

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

Fig. 1. Basic structure of proanthocyanidins: R1, R2 = H, propelargonidins;


R1 = H, R2 = OH, procyanidins; R1, R2 = OH, prodelphinidins.

Fig. 3. Structure of procyanidins B5 B8.

tion of complex mixture of grape seed extracts and in


beverages [8,15 17].
ACs are water-soluble plant pigments, localized in the
skin of the berries of red grape cultivars. The term ACs (or
anthocyanins) refers to the glycosides of anthocyanidins
(free aglycons). ACs may be acylated on the sugar moiety
with aromatic and aliphatic acids. The 3-O-h-glucosides of
cyanidin, peonidin, delphinidin, petunidin and malvidin are
the most abundant in V. vinifera and red wines (Fig. 4).

Fig. 2. Structure of procyanidins B1 B4.

Downloaded from by guest on August 8, 2016

together by carbon carbon bonds. Oxidative condensation


occurs between carbon C4 of the heterocycle and C6 or C8
of the adjacent units. The flavan-3-ol unit may have a 3V,4V
dihydroxy substitution on ring B (catechin and epicatechin)
and 3V,4V,5V trihydroxy substitution (gallocatechin and epigallocatechin). The term procyanidins indicates the polymers of catechin and epicatechin, the term prodelphinidins
indicates the polymers made by gallocatechin and epigallocatechin. Dimers, trimers (among which procyanidin C1 is
the most abundant), tetramers and oligomers up to 8 units
are present in grape (V. vinifera L.) seeds, stems, and skins
[6]. The procyanidin dimers B1 B4, characterized by the
4 ! 8 linkage, are the most common in grapes (Fig. 2). The
corresponding 4 ! 6 linked isomers B5 B8, may also
occur (Fig. 3). Until recently, complete characterization of
PAs in grapes and wines has been impaired because of the
difficulty of analysing high-molecular weight compounds.
However, newly developed methods based on HPLC coupled with mass spectrometry has allowed the characteriza-

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

595

3. Red wine polyphenols as protective agents against


atherogenesis

Fig. 4. Structure of the ACs of red wines.

3.1. Red wine polyphenols and vasorelaxation


Red wines and grapes exhibit endothelium-dependent
relaxation of blood vessels via enhanced generation and/or
increased biological activity of NO, leading to the elevation
of cGMP levels [30,31,35,36]. In vivo RWPs were shown to
reduce blood pressure in normo and hypertensive rats
[33,34,37]. The administration of purple grape juice improved the endothelium dependent, flow-mediated vasodilation in coronary artery disease patients with impaired
endothelial function [38].
The amplitude of vasorelaxation changed as a function of
the variability of wine constituents according to grape
varieties, area of cultivation, and vinification methods.
Consequently, the vasodilatatory effect does not apply to
all wines and the degree of vasorelaxation is correlated to
the content and type of phenols. Endothelium-dependent
relaxation was greatest for red wines produced en barrique, a procedure leading to high concentration of phenolic
compounds [32]. A correlation between the phenolic content
with vasodilatatory effect was later confirmed by Burns et
al. [21]. Sixteen red wines were selected to provide a range
of origins, grape varieties, and vinification methods. The
ability of the wines to act both as vasodilators ex vivo, and
as antioxidants in vitro was strongly correlated with the
phenolic content of wines. In addition, while the antioxidant

Downloaded from by guest on August 8, 2016

The six-hydroxyl of the glucose may be acylated with


acetyl, coumaroyl and caffeoyl groups. During storage and
aging of red wines, ACs react with PAs to produce more
complex pigments which may partially precipitate [8]. The
AC content of red wines may be estimated using a pH
shift method adapted from Ribereau-Gayon and Stonestreet [18] where, at pH < 1, ACs are easily detected in
their red flavylium form. ACs are then quantified spectrophotometrically as malvidin-3-glucoside equivalents,
using the extinction coefficient e = 28,000. Alternatively,
ACs are separated from the other polyphenols and directly
quantified by their maximal absorbance in the visible
range (536 542 nm) [7]. Very recently, a capillary zone
electrophoresis method for the quantitative determination
of ACs in wine has been validated as alternative to HPLC
[19].
Because of their unstability, free aglycons of ACs are
never found in grape or wine, except in trace quantities. The
levels of ACs and PAs are highly variable due to differences
in fruit source (cultivars of V. vinifera) and wine processing
[20]. According to this author, a representative content of
ACs is 400 mg/l in a red wine less than 6 months of age, not
having fermented in oak barrels, and 90 mg/l in a red wine
aged for about 2 years or fermented in oak barrels. White
wines contain only trace amounts. Red wine contains greater
amounts of PAs (750 and 1000 mg/l, young and aged
respectively) vs. white wines (20 25 mg/l). The content
of extractable PAs (1600 4300 mg/kg of grape) and ACs
(300 1900 mg/kg of grape) in different cultivars grown in
Italy was reported by Mattivi et al. [7]. Burns et al. [21]
reported the total phenol content in 16 red wines selected to
provide a range of origins, grape varieties and vinification
methods. Total AC levels, measured using a colorimetric
assay ranged from 101 to 325 AM (expressed as malvidin 3glucoside equivalents).

Dysfunction of the vascular endothelial cells, and proliferation and migration of smooth muscle cells are among the
factors contributing to atherosclerosis. Vascular endothelium
synthesises and releases nitric oxide (NO), which in turns,
promotes vasorelaxation (endothelium-dependent), reduces
platelet aggregation, and limits the flux of atherogenic
plasma proteins into the artery wall (Fig. 5). The vasorelaxant effect of NO occurs through the activation of
guanyl cyclase leading to the accumulation of cGMP [22].
The interaction between wine polyphenols and cell metabolism in the vascular tissue has been extensively investigated. These studies aimed to support for the beneficial effects
of wine consumption in preventing cardiovascular diseases.
In several animal models, the administration of red wine,
grape juice, dealcoholized red wine and PAs extract from
grape seeds attenuated the development of atherosclerotic
lesions [23 26]. In a pig model, the consumption of RWPs
induced a partial reduction of thickness of intimal proliferation even if not statistically significant [27]. Several
processes of vascular reactivity seem to be influenced by
wine ingredients. Red wine supplementation modulated
haemostatic function and prevented thrombosis in experimental animals [10,28,29]. Wine polyphenols were shown
to modulate blood pressure, promote vasodilatation, inhibit
smooth muscle cell migration and proliferation, and inhibit
platelet aggregation.

596

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

Fig. 5. Protective effects of red wine polyphenols on the arterial wall.

intracellular stores. The authors suggested that Ca2 + signalling pathways leading to NO production could involve the
activation of multiple cellular targets (G-proteins, phospholipase C, tyrosine kinase), depending on the composition of
the polyphenol mixture. In addition to increased NO synthase activity, RWPs may prolong the half-life and increase
the bioavailability of NO, by reducing its degradation
mediated by reactive oxygen species [27].
3.2. Modulation of gene expression by red wine polyphenols
All events previously described are responsible for a
short-term response mediated by RWPs. Recent studies
pointed out that polyphenolic compounds from several
sources may also have long-term properties, as they are
able to modulate gene expression in different cancer cell
lines and in macrophages [48 50].
RWPs significantly increased NO synthase expression,
acting on the promoter activity [51]. Thus, polyphenols not
only activate the enzyme, but also increase its levels, which
might lead to a long-lasting effect. Adhesion of leukocytes,
monocytes and T-lymphocytes to the vascular endothelium
is among the earliest events in inflammatory response and
atherogenesis. Successively, accumulation of leukocytes in
the arterial intima can make atherosclerotic plaque worsening. Adhesion process is facilitated by the action of endothelial-leucocyte adhesion molecules, which include
intercellular adhesion molecule 1 (ICAM-1), E-selectin
and vascular cell adhesion molecule 1 (VCAM-1). The
expression of these molecules can be transcriptionally
regulated by inflammatory cytokines such as interleukin-1
alpha and tumor necrosis factor alpha (TNF-a). It was
reported that at 5 Ag/ml, PAs extracted from grape seeds
downregulated the TNF-a induced expression of VCAM-1

Downloaded from by guest on August 8, 2016

activity was associated with different classes of phenols


(gallic acid, resveratrol and catechins), vasodilatation activity was correlated only with the total content of ACs [21].
Investigations devoted to characterize the RWPs responsible
for the endothelium-dependent relaxation activity [35,39
42] agree that monomeric catechins and simple phenols
(benzoic acid, gallic acid and hydroxycinnamic acids) are
devoid of effect. On the contrary, AC enriched fractions and
oligomeric PAs (dimers, trimers and tetramers) were the
active compounds. Threshold for relaxation by PAs
oligomers was between 0.5 and 4 Ag/ml [35,42]. Much
higher concentration (>0.1 mg/l) were required for ACs
[41]. The endothelium-dependent relaxation activity was
lost when higher molecular weight polymers were assayed
[41]. RWPs enhanced NO synthesis and cGMP accumulation only in the presence of functional endothelium. In
denuded aortic rings, RWPs concentration 103-fold higher
was necessary to induce relaxation [43,45].
Besides NO, red wine affected the formation of other
mediators of vascular tone, such as endothelium-derived
hyperpolarizing factor [43] and prostacyclin [44]. In addition, the synthesis of a potent vasoconstrictor such as
endothelin-1 is reduced by red wine in bovine aortic
endothelial cells. The suppression occurred at transcriptional
level. The decreased synthesis of endothelin-1 was associated with the inhibition of tyrosine kinase family of phosphorylating enzymes. The effect was correlated with the
polyphenol content of red wine [45].
The mechanisms underlining NO-dependent vasorelaxation caused by RWPs were investigated [36,46,47]. All
studies support for an increase in intracellular Ca2 + as the
critical step for the activation of NO-synthase. RWPs
increased cytosolic free Ca2 + by enhancing extracellular
Ca2 + entry and by increasing Ca2 + mobilization from

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

in primary human umbilical endothelial cells, leading to a


reduced adherence with leukocytes and T-cells [52,53].
Similar results were reported in one human study [54]. In
systemic sclerosis patients, plasma levels of VCAM-1,
ICAM-1 and P-selectin were found to be higher than in
normal volunteers. The administration of 100 mg/day of
PAs derived from grape seeds for 1 month attenuated the
increased expression of these adhesion molecules.
Another study used ACs from blackberry administered
to carrageenan-treated rats, a model of acute lung inflammation. ACs reduced the upregulation of ICAM-1 in lung
tissue [55]. The AC mixture used in this study differs
from that found in wine, since cyanidin-3-O-glucoside
was predominant (80% of the total ACs); nevertheless, it
might be expected that a similar behaviour by ACs occurs
in red wines.
3.3. Red wine polyphenols and vascular smooth muscle
cells

protein kinase 1 and 2). The attenuation of the signals


leading to vascular smooth muscle cell proliferation and
migration could also be the consequence of the inhibition of
PDGF h receptor by RWPs [63].
Vascular endothelial growth factor (VEGF) released from
vascular smooth muscle cells is a powerful endothelial
mitogen and stimulates the expression of adhesion molecules and monocyte chemotactic protein-1 [64]. Red wine
administered to cholesterol-fed rabbits inhibited the expression of monocyte chemotactic protein-1, but this effect was
found also in animals receiving white wine [65]. The
exposure of vascular smooth muscle cells to RWPs ( z 3
Ag/ml) markedly reduced the overexpression of VEGF
induced by PDGF and other growth factors such as athrombin and transforming growth factor-h1 [66]. This
effect was due to a reduced phosphorylation of p38MAPK,
in agreement with previous findings [60]. Because the
phosphorylation of p38MAPK and VEGF expression in
vascular smooth muscle cells were stimulated in the presence of reactive oxygen species, the authors concluded that
the inhibition of VEGF expression by RWPs involves their
antioxidant properties.
3.4. Red wine polyphenols and platelet aggregation
Platelets contribute to the rate of development of atherosclerosis and coronary artery diseases through several
mechanisms [67]. While it is widely demonstrated the
antiaggregating effect of alcohol [68], flavonoids [69 72]
and resveratrol [12], the literature survey provided only
fragmentary and contradictory information about the contribution of ACs and PAs to the inhibition of platelet
aggregation.
Demrow et al. [10] compared the effect of red and white
wines (ml/kg), and grape juice (ml/kg) on platelet activity
and thrombus formation in an in vivo dog model. Red wine
and grape juice were effective as antiplatelet and antithrombotic compounds when administered intragastrically. Administration of white wine did not produce significant
effects. Because the amount of ethanol necessary to produce
the antiplatelet and antithrombotic activity is greatly reduced when red wine rather than pure ethanol was given, it
was concluded that red wine contained platelet inhibitors in
addition to ethanol. Rein et al. [73] reported the effects of
dealcoholized red wine on platelet function in vitro and in
vivo in healthy humans. In nonstimulated platelets, dealcoholized red wine added to whole blood induced platelet
activation while suppressed the activation in response to
epinephrine. A partial confirmation of these findings derives
from the study by Russo et al. [74], who found in vitro a
significant antiaggregating activity associated with a fraction of dealcoholized red wine containing procyanidins,
ACs and catechins. No activity was associated to the other
constituents: phenolic acids, flavonols, and polymeric PAs.
However, dealcoholized red wine showed no effects when
given orally to healthy subjects [73,75]. Other studies on

Downloaded from by guest on August 8, 2016

Vascular smooth muscle cells contribute to the pathogenesis of atherosclerotic lesions, in as much as their
proliferation and migration are critical events for progressive intimal thickening and development of arterial wall
sclerosis. At the site of the lesion formation, the most potent
mitogenic and chemotactic agent for vascular smooth muscle cells is platelet-derived growth factor (PDGF) released
by platelets, endothelial cells and vascular smooth muscle
cells themselves. PDGF exerts its biological effects via
activation of two subtypes of transmembrane receptor
tyrosine kinases, termed a and h PDGF receptor. Ligand
binding to the h receptor promotes the activation of signalling enzymes which are important for cell migration and
proliferation. Activation of phosphatidylinositol 3V-kinase
(PI3K) and mitogen-activated protein kinase (MAPK) pathways as response to PDGF is implicated in vascular smooth
muscle cells motility [56 58].
Then, the effect of RWPs on the proliferation and the
migration of cultured vascular smooth muscle cells has also
been investigated [59,60]. RWPs (1 100 Ag/ml) inhibited
rat aortic smooth muscle cells proliferation and DNA
synthesis. Polyphenol fractions of different molecular
weights, c 200 400 for monomeric components (ACs,
catechins and flavonoids) and 1600 2000 for oligomeric
PAs, showed similar antiproliferative effects. Two distinct
mechanisms have been postulated. One involves the downregulation of cyclin A gene expression, through the decreased expression of transcription factors ATF-1 and CREB
(cAMP-responsive element). The second is associated with
the downregulation of PI3K activity, which is known to
interfere with cell-cycle regulation via the upregulation of
p27kip1 acting as inhibitor of cyclin-dependent kinase
[61,62]. In addition to proliferation, RWPs (1 100 Ag/ml)
inhibited vascular smooth muscle cell migration through the
specific inhibition of PI3K and p38MAPK, but not through
other MAPKs and ERK 1/2 (extracellular signal-regulated

597

598

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

humans failed to show differences in platelet aggregation


after red and white wine intake [76,77]. Further studies are
needed to sort out these discrepancies.

4. Bioavailability of red wine polyphenols

Downloaded from by guest on August 8, 2016

Two-thirds of the total intake of polyphenols in the


human diet has been estimated to be accounted by flavonoids, the most abundant being flavan-3-ols (catechins and
PAs), ACs and their oxidation products [78]. Information
about the absorption, distribution, metabolism and excretion
of individual flavonoids in humans is scarce. Furthermore,
most studies have been designed using flavonoids in the
aglycon form rather than their glycosides, which predominate in plants. Because the term flavonoid is applied in
general to a multitude of compounds, in this section we will
focus on absorption and metabolism of ACs and PAs
(dimers, trimers, oligomers and long chain polymers).
The question whether ACs are absorbed as free aglycon
or in their glycoside form was approached for the first time
by Paganga and Rice-Evans [79] who demonstrated the
presence of ACs in human plasma, but without quantification. Later, several other studies confirmed that ACs are
absorbed in the glycosilated form after oral consumption
[80 85].
All studies performed in animals and humans agree that
the bioavailability of ACs is limited [86 90]. In humans,
plasma levels of total ACs are low, the order of magnitude
being around 0.1 AM or less [82,83,91 93] and most of
the ACs are excreted in urines during the first 4 h [83].
Between 1.5% and 5.1% of the ingested ACs were
recovered in the urine of healthy subjects, within 12 h after
consumption of 300 ml of wine containing around 218 mg
of anthocyanosides. The AC levels in the urine reached a
peak within 6 h from consumption [87]. More recent
studies indicated that the amount of ACs excreted in urine
was in much lower range, 0.04 0.1% of the dose [84,85].
After the ingestion of 400 ml of Lemberger wine (containing 270 mg of ACs) by healthy volunteers, red wine
ACs were found in human plasma (Cmax: 42 F 8 Ag/l) and
urine (cumulative excretion: 0.20 0.23% of the oral dose)
as unchanged glucosides [93,94]. Other peaks detected in
the urine samples showed similar absorption spectra as the
ACs. These additional AC-like compounds may be metabolites of ACs, but the nature of the compounds was not
elucidated [85,93]. Neither the aglycon nor glucoronide
and/or sulphate derivatives were found in plasma and
urines from subjects taking wine or other sources of ACs
[80,83,88,91], but in some studies plasma and urines were
not enzymatically treated. Recently, Wu et al. [95] reported
methylated and glucoronidated AC metabolites in elderly
women after ingestion of eldberry extract. Then the issue
whether ACs undergo metabolism pre/post absorption
requires further investigations. As ACs are bulky and polar
molecules, it is expected that they are absorbed at gastro-

intestinal level through a carrier-mediated system. Indeed,


Passamonti et al. [90,96] reported that ACs are competitive
inhibitors of bilitranslocase (Ki ranging from 1.4 to 22
AM), a carrier protein expressed on gastric mucosa [97]
and in vivo experiments proved the involvement of the
stomach in the absorption of grape anthocyanosides in rats.
The structure activity relationship revealed that monoglucosyl and diglucosyl ACs were better substrates than the
corresponding aglycon [96]. The latter observation represents a further confirmation that glycosides are the preferred form for absorption. Furthermore, the ability of ACs
to penetrate the gastric mucosa could be the explanation
for the rapid appearance ( c 5 20 min) of ACs in systemic plasma [83,88,89,91]. At the same time, the contribution of the upper part of the intestine in the process of
absorption cannot be excluded.
Other questions that remain open concern the influence
of alcohol, tartaric acid, a major organic acid in wine, and
PAs to the efficiency of absorption. Results from the study
by Bub et al. [88] tend to exclude the effect of alcohol on
AC bioavailability, but the low number of subjects (6) is a
limit to draw clearcut conclusions. On the other hand,
tartaric acid increased the AUC of (+)-catechin metabolites
in plasma of rats treated orally with a dose of 100 mg/kg of
RWPs, suggesting that it can enhance the absorption and the
bioavailability of polyphenols in red wine [98]. This observation needs to be confirmed in human studies.
In comparison with ACs, the bioavailability of PAs has
received much less attention, maybe partially due to the
complexity of structures, which makes their analysis difficult, and for the lack of commercially available pure standards. Few data are available in animals and in humans. Early
studies were performed with radiolabelled PA extracts
which did not allow discrimination between monomers,
oligomers and/or metabolites. Feeding rats and mice with
a radiolabelled mixture of grape PAs [largely dimers, plus
(+)-catechin and (  )-epicatechin] permitted the identification in urines and faeces of some ill-defined PAs and of
hippuric acid, ethylcatechol and various phenolic acids. The
radioactivity was found in all the tissues of the animals
within few hours after the ingestion [99,100]. Other data
obtained from studies in chickens and sheep indicated that
the radioactivity associated with PA polymers was completely recovered in the bowel and the faeces of the treated
animals, thus showing a lack of absorption of these large
molecules [101,102].
Procyanidin polymers were also shown to be degraded in
vitro into low-molecular weight phenolic acids by a human
colonic microflora grown anaerobically [103,104]. Dimers
and trimers were uptaken at similar extent into Caco-2 cell
layer, a well-established model for human absorption, while
polymers adhered partially to the cell surface [105]. The fate
of procyanidin dimers has been studied recently: procyanidin B2 was found in plasma of subjects receiving 0.375 g
cocoa/kg body weight, as beverage. The compound
appeared in plasma as early as 0.5 h (16 nM). The maximal

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

concentration was reached by 2 h (41 nM). The single


monomers and their metabolites were also detected indicating cleavage of the dimer in the digestive tract and biotransformation [106]. The cleavage of procyanidin dimers
B2 and B5 to the corresponding monomers at gastrointestinal level has been confirmed by other studies [107 109].
Monomers are then O-methylated and conjugated as glucoronides as described for other flavonoids [110]. In contrast,
Donovan et al. [111] found that procyanidin B3, which is
also present in wine, was not bioavailable: no trace of the
compound was detected in plasma and urine after treating
rats with a grape seed extract or procyanidin B3 alone. Nor
procyanidins were cleaved to the bioavailable monomers.
The reason for these discrepancies might reside in the
experimental protocols and in the use of animal models.
Rats were fasted before receiving procyanidin B2 [109]; in
the study by Holt et al. [106], subjects consumed a flavanolrich cocoa; in the study by Donovan et al. [111], rats were
fed a standard semipurified diet supplemented with procyanidin B3 or the grape seed extract.
No studies thus far have addressed the issue of evaluating
absorption and metabolism of PAs after wine drinking. It is
then of utmost priority to eliminate this deficiency.

The interest in compounds present in red wine was


stimulated by the claim that the regular consumption of
this beverage could help in preventing cardiovascular
diseases. In comparison with white wine, which does not
share the same putative beneficial effects as red wine, the
latter contains ACs and PAs in greater amounts. Then most
of the studies reported in this review aimed to ascertain the
contribution of these two classes of polyphenolic compounds to the overall effect of red wine. As shown by the
large number of publications, ACs and PAs possess in
vitro relevant biological activities and interact with many
targets of the cell signalling pathways. However, the
physiological significance of these findings depends on
the assumption that ACs and PAs are bioavailable in the
same chemical form as that used for the in vitro studies
and can attain plasma levels as high as those used for in
vitro studies. As reported in the previous paragraph, both
assumptions are questionable. Even if most of the data
were not obtained after wine drinking, few studies available about the plasma levels of ACs and PAs report values
as 100 nM or lower, while concentrations used in vitro
exceeded these values. In addition, no data are available to
assess the distribution of ACs and PAs at the site of action
in the unmetabolized form, as it would be required to
mimic the situation of in vitro experiments. Among PAs, it
is apparent that only dimers appear in plasma as the native
compounds, but this observation needs to be confirmed in
humans drinking wine. It is then imperative to acquire
knowledge on the fate of red wine polyphenols after single

dosing and in conditions simulating the regular consumption of this beverage.

Acknowledgements
This review is dedicated to the memory of Professor
Giovanni Galli. We gratefully acknowledge Dr. Germana
Galli for her help during the preparation of this manuscript.

References
[1] Ulbright TLV, Southgate DAT. Coronary heart disease: seven dietary
factors. Lancet 1991;338:985 92.
[2] Frankel EN, Kanner J, German JB, Parks E, Kinsella JE. Inhibition
of oxidation of human low-density lipoprotein by phenolic substances in red wine. Lancet 1993;341(8843):454 7.
[3] Maxwell S, Cruickshank A, Thorpe G. Red wine antioxidant activity
in serum. Lancet 1994;344(8916):193 4.
[4] Aviram M, Fuhram B. Wine flavonoids protect against LDL oxidation and atherosclerosis. Ann N Y Acad Sci 2002;957:146 61.
[5] Francis FJ. Food colorants: anthocyanins. Crit Rev Food Sci Nutr
1989;28:273 314.
[6] Bombardelli E, Morazzoni P. Vitis vinifera L. Fitoterapia 1995;66(4):
291 317.
[7] Mattivi F, Zulian C, Nicolini G, Valenti N. Wine, biodiversity, technology and antioxidants. Ann N Y Acad Sci 2002;957:37 56.
[8] Santos-Buelga C, Scalbert A. Proanthocyanidins and tannin-like
compounds-nature, occurrence, dietary intake and effects on nutrition and health. J Sci Food Agric 2000;80:1094 117.
[9] Bravo L. Polyphenols: chemistry, dietary sources, metabolism, and
nutritional significance. Nutr Rev 1998;56:317 33.
[10] Demrow HS, Slane PR, Folts JD. Coronary artery disease/platelets:
administration of wine and grape juice inhibits in vivo platelet activity and thrombosis in stenosed canine coronary arteries. Circulation 1995;91:1182 8.
[11] Tedesco II, Russo M, Russo P, et al. Antioxidant effect of red
wine polyphenols on red blood cells. J Nutr Biochem 2000;11(2):
114 9.
[12] Fremont L. Biological effects of resveratrol. Life Sci 2000;66:
663 73.
[13] Bhat KPL, Kosmeder JW, Pezzuto JM. Biological effects of resveratrol. Antioxid Redox Signal 2001;3(6):1041 64.
[14] Kimura Y. Pharmacological studies on resveratrol. Methods Find
Exp Clin Pharmacol 2003;25(4):297 310.
[15] Vivas N, Bourgeois G, Vitry C, Glories Y, de Freitas V. Determination of the composition of commercial tannin extracts by liquid
secondary ion mass spectrometry (LSIMS). J Sci Food Agric
1996;72:309 17.
[16] De Freitas VAP, Glories Y, Bourgeois G, Vitry C. Characterisation
of oligomeric and polymeric procyanidins from grape seeds by
liquid secondary ion mass spectrometry. Phytochemistry 1998;49:
1435 41.
[17] Gabetta B, Fuzzati N, Griffini A, et al. Characterization of proanthocyanidins from grape seeds. Fitoterapia 2000;71:162 75.
[18] Ribereau-Gayon P, Stonestreet E. Le dosage des anthocyanes dans
les vins rouges. Bull Soc Chim Fr 1965;9:2649 52.
[19] Sa`en-Lo`pez R, Fernandez-Zurbano P, Tena MT. Development and
validation of a capillary zone electrophoresis method for the quantitative determination of anthocyanins in wine. J Chromatogr, A
2003;990:247 58.
[20] Waterhouse AL. Wine phenolics. Ann N Y Acad Sci 2002;957:
21 36.
[21] Burns J, Gardner PT, ONeil J, et al. Relationship among antioxidant

Downloaded from by guest on August 8, 2016

5. Conclusions

599

600

[22]
[23]

[24]

[25]

[26]

[27]
[28]

[29]

[30]

[32]

[33]

[34]

[35]

[36]

[37]

[38]

[39]

[40]

[41]

activity, vasodilation capacity and phenolic content of red wines.


J Agric Food Chem 2000;48:220 30.
Li H, Forstemann U. Nitric oxide in the pathogenesis of vascular
disease. J Pathol 2000;190:244 54.
Vinson JA, Teufel K, Wu N. Wine Red wine, dealcoholized red
wine, and especially grape juice, inhibit atherosclerosis in a hamster
model. Atherosclerosis 2001;156:67 72.
Klurfeld DM, Kritchevsky DM. Differential effects of alcoholic
beverages on experimental atherosclerosis in rabbits. Exp Mol
Pathol 1981;34:62 71.
da Luz PL, Serano JCV, Chacra AP, et al. The effect of red wine on
experimental atherosclerosis: lipid-independent protection. Exp Mol
Pathol 1999;65:150 9.
Yamakoshi J, Kataoka S, Koga T, Ariga T. Proanthocyanidin-rich
extract from grape seeds attenuates the development of aortic atherosclerosis in cholesterol-fed rabbits. Atherosclerosis 1999;142:
139 49.
de aetano G, Cerletti C. Wine and cardiovascular disease. Nutr
Metab Cardiovasc Dis 2001;11:47 50.
Wollny T, Aiello L, Di Tommaso D, et al. Modulation of haemostatic
function and prevention of experimental thrombosis by red wine in
rats: a role for increased nitric oxide production. Br J Pharmacol
1999;127(3):747 55.
De Curtis A, Porto AL, Vischetti M, Amore C, Donati MB, Iacoviello L. Effects of alcohol-free red wine on experimental thrombosis in hyperlipidemic rats. Thromb Haemost 2001;86:P1617.
Fitzpatrick DF, Hirschfield SL, Coffey RG. Endothelium-dependent
vasorelaxing activity of wine and other grape products. Am J Physiol: Heart Circ Physiol 1993;265(34):H774 8.
Fitzpatrick DF, Hirschfield SL, Ricci T, Jantzen P, Coffey RG. Endothelium-dependent vasorelaxation caused by various plant
extracts. J Cardiovasc Pharmacol 1995;26(1):90 5.
Flesch M, Schwarz A, Bohm M. Effects of red and white wine on
endothelium-dependent vasorelaxation of rat aorta and human coronary arteries. Am J Physiol: Heart Circ Physiol 1998;275(44):
H1183 90.
Mizutani K, Ikeda K, Kawai Y, Yamori Y. Extract of wine phenolics
improves aortic biomechanical properties in stroke-prone spontaneously hypertensive rats (SHRSP). J Nutr Sci Vitaminol (Tokyo)
1999;45(1):95 106.
Diebolt M, Bucher B, Andriantsitohaina R. Wine polyphenols decrease blood pressure, improve NO vasodilatation, and induce gene
expression. Hypertension 2001;38:159 65.
Fitzpatrick DF, Fleming RC, Bing B, Maggi DA, OMalley R. Isolation and characterization of endothelium-dependent vasorelaxing
compounds from grape seeds. J Agric Food Chem 2000;48(12):
6384 90.
Zenebe W, Pechao`ova O, Andriantsitohaina R. Red wine polyphenols induce vasorelaxation by increased nitric oxide bioactivity.
Physiol Res 2003;52:425 32.
Bernatova I, Pechao`ova O, Babal P, Kysela S, Stvrtina S, Andriantsitohaina R. Wine polyphenols improve cardiovascular remodelling
and vascular function in NO-deficient hypertension. Am J Physiol:
Heart Circ Physiol 2002;282:H942 8.
Stein JH, Keevil JG, Wiebe DA, Aeschlimann S, Folts JD. Purple
grape juice improves endothelial function and reduces the susceptibility of LDL cholesterol to oxidation in patients with coronary
artery disease. Circulation 1999;100(10):1050 5.
Andriambeloson E, Kleschyov AL, Muller B, Beretz A, Stoclet JC,
Andriantsitohaina R. Nitric oxide production and endothelium-dependent vasorelaxation induced by wine polyphenols in rat aorta. Br
J Pharmacol 1997;120:1053 8.
Freslon JL, Mendes A, Desgranges C, Vercauteren J. Procyanidin
oligomers from grape seed induce NO-dependent relaxation of aortic
rings via P2Y1 receptor activation. Br J Pharmacol 1997;122:200
[Abstract].
Andriambeloson E, Magnier C, Haan-Archipoff G, et al. Lobstein

[42]

[43]

[44]

[45]
[46]

[47]

[48]

[49]

[50]

[51]

[52]

[53]

[54]

[55]

[56]

[57]

[58]

[59]

[60]

Natural dietary polyphenolic compounds cause endothelium-dependent vasorelaxation in rat thoracic aorta. J Nutr 1998;128:2324 33.
Fitzpatrick DF, Bing B, Maggi DA, Fleming RC, OMalley R. Vasodilating procyanidins derived from grape seeds. Ann N Y Acad Sci
2002;957:78 89.
Ndiaye M, Chataigneau T, Andriantsitohaina JCS, Schini-Kerth VB.
Red wine polyphenols cause endothelium-dependent EDHF-mediated relaxations in porcine coronary arteries via a redox-sensitive
mechanism. Biochem Biophys Res Commun 2003;310:371 7.
Derek D, Pearson DA, German JB. Endothelial cell basal PGI2 release is stimulated by wine in vitro: one mechanism that may mediate
the vasoprotective effects of wine. J Nutr Biochem 1997;8:647 51.
Corder R, Douthwaite JA, Lees DM, et al. Khan Endothelin-1 synthesis reduced by red wine. Nature 2001;414(6866):863 4.
Andriambeloson E, Stoclet JC, Andriantsitohaina R. Mechanism of
endothelial nitric oxide-dependent vasorelaxation induced by wine
polyphenols in rat thoracic aorta. J Cardiovasc Pharmacol 1999;33:
248 54.
Martin S, Andriambeloson E, Takeda K, Andriantsitohaina R. Red
wine polyphenols increase calcium in bovine aortic endothelial cells:
a basis to elucidate signalling pathways leading to nitric oxide production. Br J Pharmacol 2002;135:1579 87.
Lin YL, Lin JK. (  )-Epigallocatechin-3-gallate blocks the induction of nitric oxide synthase by down-regulating lipopolysaccharideinduced activity of transcription factor nuclear factor-nB. Mol Pharmacol 1997;52:465 72.
Okabe S, Fujimoto N, Sueoka N, Suganuma M, Fujiki H. Modulation
of gene expression by (  )-epigallocatechin gallate in PC-9 cells
using a cDNA expression array. Biol Pharm Bull 2001;24:883 6.
Bellosta S, DellAgli M, Canavesi M, et al. Inhibition of metalloproteinase-9 activity and gene expression by polyphenolic compounds
isolated from the bark of Tristaniopsis calobuxus (Myrtaceae). Cell
Mol Life Sci 2003;60:1440 8.
Leikert JF, Rathel TR, Wohlfart P, Cheynier V, Vollmar AM, Dirsch
VM. Red wine polyphenols enhance endothelial nitric oxide synthase expression and subsequent nitric oxide release from endothelial cells. Circulation 2002;106:1614 7.
Sen CK, Bagchi D. Regulation of inducible adhesion molecule expression in human endothelial cells by grape seed proanthocyanidin
extract. Mol Cell Biochem 2001;216(1 2):1 7.
Bagchi D, Bagchi M, Sthos SJ, Ray SD, Sen CK, Preuss HG. Cellular protection with proanthocyanidins derived from grape seeds.
Ann N Y Acad Sci 2002;957:260 70.
Kalin R, Righi A, Del Rosso A, et al. Activin, a grape seed-derived
proanthocyanidin extract, reduces plasma levels of oxidative stress
and adhesion molecules (ICAM-1, VCAM-1 and E-selectin) in systemic sclerosis. Free Radic Res 2002;36(8):819 25.
Rossi A, Serraino I, Dugo P, et al. Protective effects of anthocyanins
from blackberry in a rat model of acute lung inflammation. Free
Radic Res 2003;37(8):891 900.
Knall C, Worthen GS, Johnson GL. Interleukin 8-stimulated phosphatidylinositol-3-kinase activity regulates the migration of human
neutrophils independent of extracellular signal-regulated kinase and
p38 mitogen-activated protein kinases. Proc Natl Acad Sci U S A
1997;94(7):3052 7.
Imai Y, Clemmons DR. Roles of phosphatidylinositol-3-kinase and
mitogen-activated protein kinase pathways in stimulation of vascular smooth muscle cell migration and deoxyribonucleic acid synthesis by insulin-like growth factor-I. Endocrinology 1999;140(9):
4228 35.
Hedges JC, Dechert MA, Yamboliev IA, et al. A role for
p38(MAPK)/HSP27 pathway in smooth muscle cell migration. J
Biol Chem 1999;274:24211 9.
Iijima K, Yoshizumi M, Ouchi Y, et al. Red wine polyphenols inhibit
proliferation of vascular smooth muscle cells and downregulate expression of cyclin A gene. Circulation 2000;101:805 11.
Iijima K, Yoshizumi M, Hashimoto M, et al. Red wine polyphenols

Downloaded from by guest on August 8, 2016

[31]

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

[61]

[62]

[63]

[64]
[65]

[66]

[67]

[69]

[70]

[71]

[72]

[73]

[74]

[75]

[76]

[77]

[78]
[79]
[80]

[81] Tsuda T, Horio F, Osawa T. Absorption and metabolism of cyanidin


3-O-h-D-glucoside in rats. FEBS Lett 1999;449:179 82.
[82] Matsumoto H, Inaba H, Kishi M, Tominaga S, Hirayama M, Tsuda
T. Orally administered delphinidin-3-rutinoside and cyanidin-3-rutinoside are directly absorbed in rats and humans and appear in the
blood as the intact forms. J Agric Food Chem 2001;49(3):1546 51.
[83] Cao G, Muccitelli HU, Sanchez-Moreno C, Prior RL. Anthocyanins
are absorbed in glycated forms in elderly women: a pharmacokinetic
study. Am J Clin Nutr 2001;73:920 6.
[84] Nielsen ILF, Dragsted LO, Ravn-Haren G, Freese R, Rasmussen SE.
Absorption and excretion of black currant anthocyanins in humans
and watanabe heritable hyperlipidemic rabbits. J Agric Food Chem
2003;51:2813 20.
[85] McGhie TK, Ainge GD, Barnett LE, Cooney JM, Jensen D. Anthocyanins glycosides from berry fruits are absorbed and excreted unmetabolized by both humans and rats. J Agric Food Chem 2003;
51:4539 48.
[86] Morazzoni P, Livio S, Scilingo A, Malandrino S. S. Vaccinium myrtillus anthocyanosides pharmacokinetics in rats. Arzneimittelforschung 1991;41(2):128 31.
[87] Lapidot T, Harel S, Granit R, Kanner J. Bioavailability of red wine
anthocyanins as detected in human urine. J Agric Food Chem
1998;46:4297 302.
[88] Bub A, Watzl B, Heeb D, Rechkemmer G, Briviba K. Malvidin-3glucoside bioavailability in humans after ingestion of red wine, dealcoholized red wine and grape juice. Eur J Nutr 2001;40:113 20.
[89] Milbury PE, Cao G, Prior RI, Blumberg J. Bioavailability of elderberry anthocyanins. Mech Ageing Dev 2002;123(8):997 1006.
[90] Passamonti S, Vrhovsek U, Vanzo A, Mattivi F. The stomach as a
site for anthocyanins absorption from food. FEBS Lett 2003;544:
210 3.
[91] Miyazawa T, Nakagawa K, Kudo M, Muraishi K, Someya K. Direct
intestinal absorption of red fruits anthocyanins, cyanidin-3-glucoside
and cyanidin-3,5-diglucoside, into rats and humans. J Agric Food
Chem 1999;47(3):1083 91.
[92] Mazza G, Kay CD, Cottrell T, Holub BJ. Absorption of anthocyanins from blueberries and serum antioxidant status in human subjects. J Agric Food Chem 2002;50:7731 7.
[93] Netzel M, Strass G, Bitsch I, Konitz R, Christmann M, Bitsch R.
Red wine anthocyanins and their bioavailability in humans. Ernaehrung 2002;26(2):58 63.
[94] Frank T, Netzel M, Strass G, Bitsch R, Bitsch I. Bioavailability of
anthocyanidin-3-glucosides following consumption of red wine and
red grape juice. Can J Physiol Pharmacol 2003;81(5):423 35.
[95] Wu X, Cao G, Prior RL. Absorption and metabolism of anthocyanins in elderly women after consumption of elderberry or blueberry.
J Nutr 2002;132:1865 71.
[96] Passamonti S, Urhovsek U, Mattivi F. The interaction of anthocyanins with bilitranslocase. Biochem Biophys Res Commun 2002;269:
631 6.
[97] Battiston L, Macagno A, Passamonti S, Micali F, Sottocasa GL.
Specific sequence-directed anti-bilitranslocase antibodies as a tool
to detect potentially bilirubin-binding proteins in different tissues of
the rat. FEBS Lett 1999;453(3):351 5.
[98] Yamashita S, Sakane T, Harada M, et al. Absorption and metabolism
of antioxidative polyphenolic compounds in red wine. Ann N Y
Acad Sci 2002;957:325 8.
[99] Laparra J, Michaud J, Masquelier J. Etude pharmacocinetique
des oligome`res flavanoliques (Pharmacokinetic study of flavonic
oligomers). Plant Med Phytother 1977;11:133 42.
[100] Harmand MF, Blanquet P. The fate of total flavanolic oligomers
(OFT) extracted from Vitis vinifera L., in the rat. Eur J Drug Metab
Pharmacokinet 1978;3:15 30.
[101] Jimenez-Ramsey LM, Rogler JC, Housley TL, Butler LG, Elkin RG.
Absorption and distribution of C-14-labelled condensed tannins and
related sorghum phenolics in chickens. J Agric Food Chem 1994;42:
963 7.

Downloaded from by guest on August 8, 2016

[68]

inhibit vascular smooth muscle cell migration through two distinct


signalling pathways. Circulation 2002;105:2404 10.
Collado M, Medema RH, Garcia-Cao I, et al. Inhibition of the phosphoinositide 3-kinase pathway induces a senescence-like arrest mediated by p27Kip1. J Biol Chem 2000;275(29):21960 8.
Suzuki E, Nagata D, Yoshizumi M, et al. Reentry into the cell cycle
of contact -inhibited vascular endothelial cells by a phosphatase
inhibitor. Possible involvement of extracellular signal-regulated kinase and phosphatidylinositol 3-kinase. J Biol Chem 2000;275(5):
3637 44.
Rosenkranz S, Knirel D, Dietrich H, Flesch M, Erdmann E, Bohm
M. Inhibition of the PDGF receptor by red wine flavonoids provides
a molecular explanation for the French paradox. FASEB J
2002;16(14):1958 60.
Ferrara N. Molecular and biological properties of vascular endothelial growth factor. J Mol Med 1999;77(7):527 43.
Feng AN, Chen YL, Chen YT, Ding YZ, Lin SJ. Red wine inhibits
monocyte chemotactic protein-1 expression and modestly reduces
neointimal hyperplasia after balloon injury in cholesterol-fed rabbits.
Circulation 1999;100:2254 9.
Oak MH, Chataigneau M, Keravis T, et al. Red wine polyphenolic
compounds inhibit vascular endothelial growth factor expression in
vascular smooth muscle cells by preventing the activation of the p38
mitogen-activated protein kinase pathway. Arterioscler Thromb Vasc
Biol 2003;23:1001 7.
Fuster V, Badimon L, Badimon JJ, Chesebro JH. The pathogenesis
of coronary artery disease and acute coronary syndromes. N Engl J
Med 1992;326:242 50.
Renaud SC, Ruf JC. Effects of alcohol on platelet functions. Clin
Chim Acta 1996;246:77 89.
Landolfi R, Mower R, Steiner M. Modification of platelet function
and arachidonic acid metabolism by bioflavonoids. Biochem Pharmacol 1984;33:1525 30.
Gryglewski R, Korbut R, Robak J, Swies J. On the mechanism of
antithrombotic action of flavonoids. Biochem Pharmacol 1987;36:
317 22.
Beretz A, Stierle A, Anton R, Cazenave JP. Role of cAMP in the
inhibition of human platelet aggregation by quercetin, a flavonoid
that potentiates the effect of prostacyclin. Biochem Pharmacol
1981;31:3597 600.
Slane PR, Qureshi AA, Folts JD. Platelet inhibition in stenosed
canine arteries by quercetin and rutin, polyphenolic flavonoids found
in red wine. Clin Res 1994;42:162 [Abstract].
Rein D, Paglieroni TG, Pearson DA, et al. Cocoa and wine polyphenols modulate platelet activation and function. J Nutr 2000;
130:2120S 6S.
Russo P, Tedesco I, Russo M, Russo GL, Venezia A, Cicala C.
Effects of de-alcoholated red wine and its phenolic fractions on
platelet aggregation. Nutr Metab Cardiovasc Dis 2001;11(1):25 9.
Pellegrini N, Pareti FI, Stabile F, Brusamolino A, Simonetti P.
Effects of moderate consumption of red wine on platelet aggregation
and haemostatic variables in healthy volunteers. Eur J Clin Nutr
1996;50(4):209 13.
Lavy A, Fuhrman B, Market A, et al. Effect of dietary supplementation of red or white wine on human blood chemistry, hematology
and coagulation: favorable effect of red wine on plasma high-density
lipoprotein. Ann Nutr Metab 1994;38(5):287 94.
Pace-Asciak CR, Rounova O, Hahn SE, Diamandis EP, Goldberg
DM. Wines and grape juices as modulators of platelet aggregation in
healthy human subjects. Clin Chim Acta 1996;246(1 2):163 82.
Scalbert A, Williamson G. Dietary intake and bioavailability of polyphenols. J Nutr 2000;130(8S):2073S 85S.
Paganga G, Rice-Evans CA. The identification of flavonoids as
glycosides in human plasma. FEBS Lett 1997;401:78 82.
Cao G, Prior RL. Anthocyanins are detected in human plasma after
oral administration of an elderberry extract. Clin Chem 1999;45(4):
574 6.

601

602

M. DellAgli et al. / Cardiovascular Research 63 (2004) 593602

[102] Terril TH, Waghorn GC, Wooley DJ, McNabb WC, Barry TN. Assay and digestion of 14C-labelled condensed tannins in the gastrointestinal tract of sheep. Br J Nutr 1994;72:467 77.
[103] Groenewoud G, Hundt HKL. The microbial metabolism of condensed (+)-catechins by rat-caecal microflora. Xenobiotica 1986;
16:99 107.
[104] Deprez S, Brezillon C, Rabot S, et al. Polymeric proanthocyanidins
are catabolized by human colonic microflora into low-molecular
weight phenolic acids. J Nutr 2000;130(11):2733 8.
[105] Deprez S, Mila I, Scalbert A, Huneau JF, Tome D. Transport of
proanthocyanidin dimer, trimer and polymer across monolayers of
human intestinal epithelial Caco-2 cells. Antioxid Redox Signal
2001;3(6):957 67.
[106] Holt RR, Lazarus SA, Sullards MC, et al. Procyanidin dimer B2
[epicatechin-(4h-8)-epicatechin] in human plasma after the consumption of a flavanol-rich cocoa. Am J Clin Nutr 2002;76:
798 804.
[107] Spencer JP, Chaudry F, Pannala AS, Srai SK, Debnam E, Rice-Evans

[108]

[109]

[110]

[111]

C. Decomposition of cocoa procyanidins in the gastric milieu. Biochem Biophys Res Commun 2000;272(1):236 41.
Spencer JP, Schroeter H, Shenoy B, Srai SK, Debnam E, Rice-Evans
C. Epicatechin is the primary bioavailable form of the procyanidin
dimers B2 and B5 after transfer across the small intestine. Biochem
Biophys Res Commun 2001;285(3):588 93.
Baba S, Osakabe N, Natsume M, Terao J. Absorption and urinary
excretion of procyanidin B2 [epicatechin-(4h-8)-epicatechin] in rats.
Free Radic Biol Med 2002;33(1):142 8.
Spencer JP, Schroeter H, Crossthwaithe AJ, Kuhnle G, Williams
RJ, Rice-Evans C. Contrasting influences of glucuronidation and
O-methylation of epicatechin on hydrogen peroxide-induced cell
death in neurons and fibroblasts. Free Radic Biol Med 2001;31(9):
1139 46.
Donovan JL, Manach C, Rios L, Morand C, Scalbert A, Remesy C.
Procyanidins are not bioavailable in rats fed a single meal containing
a grapeseed extract or the procyanidin dimer B3. Br J Nutr 2002;
87:299 306.

Downloaded from by guest on August 8, 2016

Você também pode gostar